Tel Aviv University The Lester and Sally Entin Faculty of Humanities The Chaim Rosenberg School of Jewish Studies and Archaeology The Jacob M. Alkow Department of Archaeology and Ancient Near Eastern Cultures

The Intermediate Bronze Age (c. 2500–1950 BCE) in the Negev Highlands: The Geoarchaeological Perspective

Thesis submitted for the degree “Doctor of Philosophy” (Ph.D.) by Zachary Clark Dunseth

This work was carried out under the supervision of Prof. Finkelstein (Tel Aviv University) and Prof. Ruth Shahack-Gross (University of Haifa)

Submitted to the Senate of Tel Aviv University 2019 1

Table of Contents

Acknowledgements ...... 6

Abstract ...... 9

List of Figures ...... 11

List of Tables ...... 14

1 Introduction ...... 15

2 Background: Geology and Environment of the Negev Highlands ...... 22

2.1 General Geology of the Negev Highlands ...... 22

2.2 Modern Environmental Conditions ...... 22

2.3 Paleoenvironment ...... 24

3 Settlement history of the Negev Highlands ...... 25

3.1 Results of archaeological surveys in the Negev Highlands ...... 25

3.2 Origin and nature of IBA Negev settlement ...... 27

3.2.1 Kochavi (1967, 2009) ...... 27

3.2.2 Dever (1970, 1971, 1973, 1980, 1985, 1992, 1995, 2014): ...... 28

3.2.3 Cohen (1983, 1986, 1992, 1999): ...... 29

3.2.4 Finkelstein (1989, 1991a, 1995a,b) ...... 30

3.2.5 Haiman (1996) ...... 30

3.2.6 Rosen (2011a,b, 2016, among others) ...... 32

3.2.7 Gidding (Gidding 2016; also Ben-Yosef et al. 2016) ...... 32

3.3 Types of archaeological sites in this study ...... 33

2

3.4 Chronology of the Intermediate Bronze Age ...... 35

3.4.1 Traditional chronology of the Intermediate Bronze Age ...... 35

3.4.2 Radiocarbon investigations at IBA sites in the Negev and surrounding regions .... 35

4 Methodology ...... 47

4.1 Geo-ethnoarchaeology and reconstructing ancient subsistence practices ...... 47

4.1.1 Dung spherulites ...... 48

4.1.2 Phytoliths ...... 49

4.1.3 Ash pseudomorphs ...... 51

5 Materials and methods ...... 52

5.1 Description of sites, their microenvironments and previous excavations ...... 52

5.1.1 Mashabe Sade ...... 52

5.1.2 Ein Ziq ...... 55

5.1.3 Nahal Boqer 66 ...... 57

5.1.4 Nahal Nizzana 332 ...... 59

5.2 Analysis...... 60

5.2.1 Fourier transform infrared (FTIR) spectroscopy ...... 60

5.2.2 X-Ray fluorescence (XRF) ...... 61

5.2.3 Microremains: phytoliths, dung spherulites and ash pseudomorphs ...... 61

5.2.4 Spatial analysis and statistics ...... 65

6 General description of the articles and their contribution ...... 66

6.1 Methodological contribution ...... 66

3 6.2 Chronology of IBA settlement in the Negev ...... 66

6.3 Subsistence economies and trading systems of IBA Negev settlement ...... 67

6.4 Preliminary results from Nizzana 332 ...... 68

7. Articles and preliminary results from Nahal Nizzana ...... 68

7.1 Methodological contribution (Dunseth and Shahack-Gross 2018) ...... 69

7.2 Chronology of IBA settlement in the Negev (Dunseth et al. 2017) ...... 77

7.3 Subsistence economies and trading systems of IBA Negev settlement (Dunseth et al.

2016, 2018) ...... 92

7.4 Preliminary results from Nizzana 332 ...... 153

7.4.1 Excavation Summary ...... 154

7.4.2 Field Observations ...... 157

7.4.3 Macroarchaeological data ...... 158

7.4.4 Chronometric data: radiocarbon and optically stimulated luminescence ...... 169

7.4.5 Microarchaeological data ...... 171

7.4.6 Summary of Nahal Nizzana ...... 179

8 Discussion ...... 181

8.1 Methodological contributions ...... 181

8.2 Subsistence strategies at IBA Negev sites: updated model ...... 182

8.2.1 Small sites: Microarchaeological data as direct indicators of pastoral nomadism183

8.2.2 Large sites: Microarchaeological data as direct indicators of fire activities ...... 183

8.2.3 Pottery assemblages of nomadic pastoralists ...... 184

8.2.4 Pottery assemblages of large sites as indicators of trade ...... 185

4 8.2.5 Site layout, activity areas and use of space at small and large sites ...... 185

8.2.6 Trade: Copper and desert goods ...... 187

8.2.7 A note on copper exchange vectors ...... 190

8.2.8 Lithic assemblages as indirect indicators of subsistence ...... 191

8.3 Chronology and settlement history of the Negev Highlands ...... 192

9 Conclusion ...... 194

10 Bibliography ...... 196

11 Appendix A: Mineralogical and microremain data from Nahal Nizzana 332.1...... 223

קתציר...... 225

5

Acknowledgements

Long acknowledgements are boring. So here’s a short one.

Above all, special thanks to my supervisors and mentors Israel Finkelstein and Ruth Shahack- Gross for their unwavering support, education and guidance for the last eight (!) years. I owe you everything.

Generous scholarships from the Israel Institute (2012-2014), Tel Aviv University (2014-2015) and the Dan David Foundation (Archaeology and the Natural Sciences, 2017-2018) supported my studies, for which I am grateful. Grants from the European Research Council (No. 229418, PIs: Israel Finkelstein and Steve Weiner) and the German-Israeli Foundation for Scientific Research and Development (No. I-1244-107.4/2014, PIs: Ruth Shahack-Gross and Markus Fuchs) sustained me and this research.

To my mentors and friends at the Kimmel Center for Archaeological Science, Weizmann Institute of Science (2012-2015); the Department of Geological Sciences, the George Washington University (2014), the Intensive Course on Soil Micromorphology in Zagreb, University of Croatia (2015), and the Laboratory for Sedimentary Archaeology, University of Haifa (2015-2019), thank you for teaching me far too much about geosciences in general, and dung in particular. Special thanks to Steve Weiner for laboratory access and support in the beginning of this research; to Richard Tollo for a rigorous introduction to mineralogy; to George Stoops, Rosa Poch, Vera Marcelino, L. Galović, H. Posilović, M. Brlek for an intensive primer in micromorphology.

To the people in the Megiddo Office, thank you for your friendship and help with organizing too many excavations (Ma’ayan Mor, Naama Walzer and Sivan Einhorn in particular). Sorry for making you bear this with me.

To the dozens (!) of volunteers willing to join unpaid excavations in the desert (the infamous ‘Zachscavations’), thank you for making this dissertation possible. There’s pictures of all you wonderful people in the next few pages. Special thanks also to Alon Shavit and Boaz Gross of the Israeli Institute of Archaeology for logistical support throughout fieldwork; and to the Israel Antiquities Authority and National Parks Authority for their permission and enthusiasm.

To Aaron Gidding and Karen Covello-Paran, thank you for great conversations and a long- suffering interest in the IBA.

To my scattered family, sorry for missing birthdays, holidays and life events. I miss you all.

To Rachel, thank you for your love and endless support from afar. Let’s never do this again.

Finally, to all my friends in Tel Aviv and Haifa—and those who’ve scattered elsewhere—you know who you are. You made this worth it. Thank you most of all.

6 Mashabe Sade, January 2013 (from left): Tomi, Ma‘ayan Mor, Adam Kaplan, Salman Abu-Jelidan, Israel Finkelstein, the author, Dafna Langgut, Raphaela Primus [not shown: Ruth Shahack-Gross, Shirly Ben-Dor Evian, Julia Fridman, Elon Heymans].

Mashabe Sade (and Nahal Boqer), March 2013 (clockwise from left): Tomi, Ma‘ayan Mor, Ruth Shahack-Gross, Israel Finkelstein, Mandy Morrow, Mike Millman, Erin Hall, Eli Itkin, Idan Jonish, George Mavronanos, Dan Devery.

Ein Ziq 2014 (from back to front, left to right): Unknown, Mathilde Forget, Andrew Pleffer, Abra Spiciarich, Mark Cavanagh, Joe Pacheco, Eli Itkin, Erin Hall, Gennadiy Shoykhedbrod, Raphaela Primus, Dana Ackerfeld, [not shown: Ma’ayan Mor, Ruth Shahack-Gross, Israel Finkelstein, Dafna Langgut, Kei Ishii, Yoav Weingarten, Alon Arad, Andrea Junge, Johanna Lomax, Markus Fuchs].

7 Ein Ziq, 2015 (from left): The author, Logan Hunt, Batel Hunt, Adam Kaplan, Naama Walzer, Joe Pacheco, Abra Spiciarich, Mathilde Forget, Bar Efrati, Erin Hall, Vanessa Workman, Arnauld Puy, Dana Ackerfeld, Mark Cavanagh, Ruth Shahack-Gross, Lee Oz [not shown: Israel Finkelstein, Assaf Kleiman, Nadav Nir, Sarah Richardson, Astrid Hasday, Sharon Hasday].

Nahal Boqer 2016 (from left): Ruth Shahack-Gross, Assaf Kleiman, Sabine Kleiman, Naama Walzer, the author, Israel Finkelstein, Omer Ze’evi, Alex Wrathall, Abra Spiciarich.

Nahal Nizzana 2017 (from left): Dana Ackerfeld, Danilo Giordano Rabell, Ronnie Avidov, Eli Itkin, Erin Hall, Angie Hodson, the author, Alex Wrathall, Bar Efrati, Mark Cavanagh, Maya Hadash, David Krouwer [not shown: Omer Ze’evi, Andrea Junge, Tillman Wolpert, Israel Finkelstein, Ruth Shahack-Gross].

8

Abstract

This dissertation presents the results of a detailed macro- and microarchaeological

investigation into subsistence practices, chronology and settlement history of the Negev

Highlands during the Intermediate Bronze Age (IBA, c. 2500-1950 BCE). Four sites were

excavated and investigated in the scope of this research: Ein Ziq and Mashabe Sade, two of the

largest central sites in the Negev Highlands, as well as Nahal Boqer 66 and Nahal Nizzana 332,

two small sites typical of the period. This is the first major investigation into IBA subsistence

strategies and chronology in several decades, and the first ever to look at the subject using

high-resolution microarchaeological sampling and analysis.

First, this dissertation details a new method developed based on the study of sediments from

Nahal Boqer 66 to rapidly identify ancient animal dung at archaeological sites using Fourier

transform infrared spectroscopy (FTIR). This is important for addressing subsistence practices

questions that are associated with herding and livestock grazing habits, and can be carried out

in the field during excavation.

Second, the macro- and microarchaeological data from the sites excavated are presented and

interpreted to explore questions regarding subsistence strategies and ancient activity areas. The

microarchaeological evidence is unequivocal and indicates that the small sites of Nahal Boqer

66 and Nahal Nizzana 332 were primarily engaged in livestock management, with herds

grazing on the local environment. No evidence for opportunistic cereal agriculture was

identified. The ceramic assemblages are dominated by cooking and serving vessels, and the high-resolution collection and sampling of Nizzana 332 shows activity and refuse disposal patterns common to both ancient and 20th century CE pastoral sites.

In contrast, the large central sites of Ein Ziq and Mashabe Sade show no evidence for any food

production—no livestock management or cereal farming. The ceramic assemblages are

9

dominated by non-local closed storage vessels of various sizes. Copper ingots, implements and scrap are also common at both sites; however, a detailed XRF investigation showed no evidence for copper production or processing activities, contrary to earlier ideas. Based on the macro- and microarchaeological data, large central sites are reconstructed as trading outposts, situated in defensible locations (Mashabe Sade) or near permanent water sources (Ein Ziq).

Finally, the chronological aspects of the radiocarbon and optically stimulated luminescence

(OSL) ages from our excavations are discussed. This research indicates that small sites are a long-term phenomenon—Nahal Boqer 66 shows continuous radiocarbon determinations from c. 3300-2300 BCE—while large central sites such as Ein Ziq appear limited to only the first half of the IBA, c. 2500-2200 BCE. Both large and small sites are abandoned c. 2200 BCE, paralleling the evidence from Faynan, and in particular Khirbet Hamra Ifdan. The collapse of the Negev and Faynan copper system is linked primarily with the demise of the Egyptian Old

Kingdom. The persistence of the small sites throughout the Early Bronze (EB) into the

Intermediate Bronze Age rekindles old discussions about the EBII-III division in the Negev and surrounding regions, and also suggests that during the third millennium BCE the southern desert copper system is insulated from—or resilient to—the general failure of the urban areas to the north.

10

List of Figures

Figure 1: Map of region and location of four sites presented in this research.

Figure 2: Histogram of sites in the Negev Highlands by period according to survey. Survey maps included: 160, 163, 164, 166, 167, 168, 169, 170, 173, 174, 177, 178, 193, 194, 195, 196, 198, 199, 200, 201, 203, 204, 224, 225. All data collected from the Archaeological Survey of Israel website: https://survey.antiquities.org.il/ .

Figure 3: Plan of Mashabe Sade, redrawn after Cohen 1999: Fig. 71. Survey structures indicated as plain ovals; excavated structures detailed. Scale is in meters.

Figure 4: A) Aerial photo of Ein Ziq, taken in 2015 after new excavations. B) Ein Ziq site plan, with new areas excavated and sampled in red. Redrawn after Cohen 1999: Fig. 88.

Figure 5: A) Aerial photo of Nahal Boqer 66 taken in 2018 (courtesy Omer Ze’evi). B) Nahal Boqer 66 site plan redrawn after Cohen 1999: Fig. 41. New excavation areas shown in red.

Figure 6: Georectified orthophoto of Nahal Nizzana 332.1, produced using Agisoft Photoscan 1.42.

Figure 7: Coefficient of variation (CoV) of samples analyzed in duplication to microremain concentrations. All concentrations reported in millions per g of sediment. A) Phytolith concentrations to CoV. B) Dung spherulite concentrations to CoV. C) Ash pseudomorph concentrations to CoV. Note trendlines leveling off around 30-40% in the phytolith and ash pseudomorph plots, approximately the same error as reported in Albert and Weiner 2001; Katz et al. 2010 and Gur-Arieh et al. 2013. The samples exceptionally rich in dung spherulites (orders of magnitude greater than the other microremains) show lower error.

Figure 8: Schematic site plan redrawn after Haiman 1991: 128, with Nizzana 332.1 and Nizzana 332.2 outlined in red. Note that Nizzana 332.3 is beyond the limit of the site plan approximately 250 m to the west.

11

Figure 9: Georectified orthophoto of Nahal Nizzana 332.1 with 1 x 1 m grid letters and numbers indicated. Red lines indicate vertical sections studied. Note scale in Squares M-N12 is 0.8 m.

Figure 10: Nahal Nizzana 332.1 pottery analysis. A) presents total assemblage (all sherds); B) presents diagnostic sherds only. Note that surface sherds are not included in the locus totals.

Figure 11: Pottery distribution normalized to weight in g of ceramic material by excavated area (m2). Weights presented include both diagnostic and non-diagnostic sherds.

Figure 12: Distribution of lithics at Nahal Nizzana 332.1 according to quantity/area. A) Tool distribution (includes all tool types); B) Debitage distribution; C) Debris distribution.

Figure 13: Location and age ranges of radiocarbon determinations and OSL samples. Later intrusions in red.

Figure 14: Calcite grinding curves for archaeological and control sediments from Nahal Nizzana 332.1. Local sediment controls are plotted (NIZ-C1, alluvial sediment from north of the site) and limestone (NIZ-C8). Note all archaeological sediments plot between alluvial controls, ash and plaster. Courtyard sediments plot closer to the plaster trendline, due to elevated concentrations of dung spherulites (see below and cf. Dunseth and Shahack-Gross 2018). Reference trendlines are from Regev et al. 2010 and Dunseth and Shahack-Gross 2018.

Figure 15: A) Boxplots of phytolith and dung spherulite concentrations from surface sediments and controls. Minor enrichment of phytolith concentrations can be observed in all archaeological contexts. Enrichment in dung spherulite concentrations is apparent only in the surface sediments of the courtyard. B) Boxplots of phytolith and dung spherulite concentrations from the archaeological accumulation on bedrock. Phytolith concentrations are essentially the same inside and outside the structures (within error) in comparison to the surface sediments, while concentrations in the courtyard are enriched. Dung spherulite concentrations

12

are markedly enriched in the courtyard and in a few samples from inside the structure. Note different scales for phytolith and dung spherulite concentrations.

Figure 16: Inverse distance weighted interpolation of logarithmically normalized microremain concentrations. Point locations are marked as white circles. A) Dung\ spherulite concentration interpolation. Clear enrichment in dung spherulites is interpolated inside Courtyard 17/NIZ/10, and to a lesser extent in the eastern part of Structure 17/NIZ/11. B) Phytolith concentration interpolation. Enriched phytolith concentrations are only clearly interpolated in the northern part of Courtyard 17/NIZ/10. Note in general, higher values indicate higher microremain concentrations.

Figure 17: Phytolith morphologies of samples from the courtyard (NIZ-L11.2) and the open area outside the structure (NIZ-M19.2) at Nahal Nizzana 332.1. Note the dominance of dicotyledonous morphotypes in both samples.

13

List of Tables

Table 1: Radiocarbon determinations from Early and Intermediate Bronze Age sites in the Negev and surrounding regions.

Table 2: Coefficient of variation data for all samples counted in duplicate. Note high error for samples with low concentrations of microremains.

Table 3: Whole lithic assemblage from Nizzana 332.1.

Table 4: Lithic assemblage from Nahal Nizzana 332.1 by locus.

Table 5: Radiocarbon determinations from Nahal Nizzana 332.1. Note Beta- 499151 is a modern intrusion.

14

1 Introduction

After nearly a century of study, the Intermediate Bronze Age (IBA, also known as the Early

Bronze Age [EB] IV or the Middle Bronze Age [MB] I, see Palumbo 2008: 227–228 for detailed discussion on terminology) remains an enigmatic period in the settlement history of the southern Levant. Conventionally dated to approximately 2300–2000 BCE (Stern 2008), the

IBA was traditionally characterized by the systemic collapse of urban sites in the north and corresponding to a unique wave of new settlement in the arid Negev Highlands (Dever 1985).

The impetus behind this dramatic settlement oscillation was and remains debated, with scholars variously associating the phenomenon with changing political (Kochavi 1967), demographic

(Kenyon 1951, 1966; Prag 1985), socio-economic (Dever 1980), and climatic conditions (Bell

1971; Rosen 1989, 1995; Weiss et al. 1993; Frumkin 2009) throughout the southern Levant

(see Palumbo 2008 for conventional review; for an update, see Kennedy 2016).

Until the last decade, the debate over the nature of the IBA in the southern Levant had mellowed into a general consensus that the period was characterized in the north by a shift from urban life to rural agropastoralism (Palumbo 1990; Esse 1991; Dever 1995), and in the

Negev, to subsistence practices characterized by animal husbandry and opportunistic agriculture (Finkelstein 1991, 1995a; Cohen 1992, 1999; Dever 1995). Hoards of copper ingots found at Negev sites (as well as at sites in the north), and the then-recent discovery of an IBA copper production center at Hamra Ifdan complicated this view, suggesting that a developed copper economy sustained much of the Negev settlement system (Haiman 1996; Adams 2000;

Levy et al. 2002). Some debate arose over the role of copper and local subsistence economies

(Haiman 1996 contra Cohen 1999; also, Dever 2014), but generally all settlements, large and small, were assumed to have practiced a form of mixed semi-sedentary agropastoralism

(Finkelstein 1991; Cohen 1999). However, to date, these assumptions have been based primarily on (often unstated) ethnographic and historical parallels to pre-modern and recent

15

Bedouin (e.g. Palmer 1871; Musil 1908; Marx 1967; Ginguld et al. 1997), historical documents

from later periods (e.g., Kraemer 1958; Mayerson 1962) and a rather small collection of archaeological (lithic, faunal and botanical) assemblages (e.g. Kochavi 1967; Cohen 1999;

Vardi et al. 2007; Dever 2014).

Recently, the collation and reanalysis of radiocarbon determinations from excavations across the southern Levant redated the IBA to 2500–1950 BCE (Regev et al. 2012). This backdates and extends the period by two centuries and places the IBA contemporary both with the large

EB IV city-states in and the Egyptian Old Kingdom (3rd–6th Dynasties; Dunseth et al.

2016; Ben-Yosef et al. 2016; Kennedy 2016). This chronological correction forces a

reevaluation of the historical framework of the IBA, and the role of Negev settlement systems

within it.

Simultaneously, developments in microarchaeology have given us new tools to investigate

aspects of human culture and ancient systems (see Weiner 2010; Karkanas and Goldberg 2019).

High-resolution paleoclimate reconstructions, in conjunction with the new chronology, has shown that the relationship between settlement and environment is complicated and period- specific (Langgut et al. 2014, 2015; Kagan et al. 2015; Laugomer et al. forthcoming).

Geoethnoarchaeological work on modern nomadic pastoralists and semi-sedentary agropastoralists provide foundations for interpreting subsistence practices and settlement organization, site formation, preservation and post-depositional processes at various archaeological sites (see Friesem 2016; Shahack-Gross 2017; specifically, Shahack-Gross et al. 2003, 2014; Mallol et al. 2007; Shahack-Gross and Finkelstein 2008; Friesem et al. 2014a,b,

2016; Gur-Arieh et al. 2013, 2014, 2018; Portillo et al. 2014, 2017). Having generally exhausted the interpretive potential of the limited macroarchaeological assemblage from the

16

Negev Highlands, microarchaeological methodologies provide new approaches to reassess old

assumptions.

This dissertation builds upon the recent chronological and microarchaeological developments

to reassess the Negev Highland settlement phenomenon during the IBA. Primarily, this work

utilized macro- and microarchaeological data including ceramic analysis, phytoliths and

calcitic microremains, mineralogy and elemental analysis together with high resolution chronometric (‘absolute’) dating methods to reconstruct 1) site use, 2) subsistence strategies and 3) chronology of four archaeological sites across a transect of different micro- environments in the Negev Highlands.

This dissertation presents four published articles and unpublished material from new macro- and microarchaeological investigations at two of the largest IBA sites, Mashabe Sade and Ein

Ziq, and two smaller sites, Nahal Boqer 66 and Nahal Nizzana 332 (Fig. 1). First, I present a new method to identify degraded animal dung at archaeological sites using Fourier transform infrared spectroscopy (FTIR) (Dunseth and Shahack-Gross 2018). This method was applied to archaeological sediments at Nahal Boqer 66 (and subsequently Nahal Nizzana 332) and found to reliably differentiate degraded dung sediments from other sediment samples at both sites.

As such, it proved to be an effective technique for rapidly evaluating sediment samples from open air archaeological sites in calcareous environments such as the Negev Highlands.

Second, I present the chronological data, utilizing radiocarbon and optically stimulated luminescence (OSL) ages (Dunseth et al. 2017). Small sites are shown to be a long-term phenomenon, existing from mid-fourth through most of the third millennium BCE without significant interruption (~3600-2200 BCE, Dunseth et al. 2017). Importantly, this data shows there is no Early Bronze III lacuna in Negev settlement, assumed in archaeological surveys published in the last few decades. In contrast, large sites are unique to the first half of the IBA,

17

c. 2500–2200 BCE (Junge et al. 2016; Dunseth et al. 2017). For the first time, this research shows that large and small sites coexisted in the first half of the IBA, and both site types were abandoned thereafter. This parallels the trajectory of the Egyptian Old Kingdom—especially the 3rd through 6th Dynasties—and reflects the demand for Arabah copper originating in

Faynan.

Third, I present the macro- and microarchaeological data from the sites investigated during the scope of this research (Dunseth et al. 2016, 2018). The small sites of Nahal Boqer 66 and Nahal

Nizzana 332 show overwhelming microarchaeological evidence for the stabling of free-grazing animals and subsistence based on livestock management. No evidence for cereal agriculture, nor the foddering with cereal byproducts, was identified in any contexts. This data has ethnographic parallels in premodern Bedouin sites and archaeological parallels in the Iron IIA

(Shahack-Gross and Finkelstein 2008; Shahack-Gross et al. 2014; Dunseth et al. 2016). These microarchaeological assemblages are in direct contrast to those found at later Byzantine/Early

Islamic sites (Shahack-Gross et al. 2014; Dunseth et al. 2019), known to be actively engaged in agriculture from textual records (Kraemer 1958; Mayerson 1965) and OSL investigations of agricultural terraces (Avni et al. 2013, 2019). Features typical of ephemeral sites (throughout history) including limited ceramic assemblages dominated by domestic (cooking and serving) forms, spatial patterning of deposited macro- and microremains, are also attested at both IBA sites (cf. Rosen and Avni 1997; Saidel 2002a,b, 2004; Saidel and Haiman 2015; Palmer et al.

2007).

In contrast to the smaller sites, the large sites of Mashabe Sade and Ein Ziq show no macro- or microarchaeological evidence for any sort of animal or cereal production. This follows earlier lithic evidence (e.g., Vardi 2005; Vardi et al. 2007) that discounted the assumed role of farming at large sites. Instead, the sites are characterized by an overwhelming assemblage of non-local

18 materials and desert goods, including imported ceramics dominated by storage vessels, Red

Sea shells, bitumen, as well as copper ingots and finished objects (Goren 1996; Cohen 1999;

Dunseth et al. 2018). Although copper ingots, finished objects and scrap were common at Ein

Ziq and Mashabe Sade, detailed X-Ray Fluorescence (XRF) studies of archaeological sediments indicate that earlier suggestions of copper processing/producing industries at large sites (Haiman 1996) are generally unfounded (Dunseth et al. 2016, 2018). Based on the accumulated micro- and macroarchaeological data, as well as their position in defensible localities or near water sources (Haiman 1996), I propose in this research that large sites can best be understood as trading hubs or waystations along the copper exchange network, and are almost exclusively supported by trade.

Finally, the Negev systems are interpreted in their historical context (Dunseth et al. 2017). As mentioned above, small and large IBA sites track the rise and fall of the Egyptian Old

Kingdom. Small sites are a long-term phenomenon, existing from at least the Early Bronze I through the first half of the IBA without significant interruption (Dunseth et al. 2017). This suggests that the southern settlement system, previously interpreted as reacting to oscillations of urban systems to the north (cf. Finkelstein 1988, 1995a,b; Rosen 2011b), is somewhat insulated from this system during the third millennium BCE.

Large sites are without precedent and limited to only the first half of the IBA. It is argued they are reflect to the collapse of the EB Beersheva-Arad Valley copper networks previously controlled by Arad in the middle of the third millennium BCE (Gidding 2016; Finkelstein et al. 2018). I argue that the large sites, dominated by trade goods, represent the takeover and rerouting of EB copper exchange networks by indigenous desert groups, who shift the trade pathways south into the Negev Highlands (see Dunseth et al. 2018; Finkelstein et al. 2018). It is argued that continuing prosperity in the Negev Highlands, Sinai and Transjordan is a linked

19 to the demand for Arabah copper, one which only ends with the collapse of the Egyptian Old

Kingdom (concordant with other recent interpretations from the south: Gidding 2016; Ben-

Yosef et al. 2016; Dunseth et al. 2016; Finkelstein et al. 2018).

20 Figure 1: Map of region and location of four sites presented in this research.

21 2 Background: Geology and Environment of the Negev Highlands

2.1 General Geology of the Negev Highlands

The Negev Highlands is an arid region in the south of modern Israel, made up of parallel

Cenomanian-Turonian synclines and anticlines. Elevation ranges from c. 600 m in the northeast

to c. 1000 m in the southwest. The regional bedrock is made up mainly of marine carbonates—

limestone, dolomite and chalk—and chert (Avni et al. 2006).

The modern landscape developed during the late Miocene-Pliocene and early Pleistocene, with

the uplifting of the Negev Highlands and the faulting of the Dead Sea Rift (Avni et al. 2006:

179–180). The drainage systems, wadi beds and valleys of the area are related to several phases of accumulation and fluvial erosion during the last interglacial (70–18 kya). During the late

Pleistocene and early Holocene, Sinai-Saharan aeolian silt and dust were deposited, sometimes to a thickness of several meters (Shanan 2000; Avni et al. 2006; Crouvi et al. 2008, 2009).

Reworked aeolian dust was spread and mixed with local alluvial-colluvial gravels and sediments, and re-deposited along the Negev wadis and valleys (Avni et al. 2006, 2013: 332–

334). The thickness of these sediments (a few to several meters) allows for percolation and

water retention, making the wadis and valleys rich in both flora and fauna and elevating the

agricultural potential (Evenari et al. 1982; Avni et al. 2012: 14; Avni et al. 2013: 333–334).

2.2 Modern Environmental Conditions

Average precipitation in the Negev Highlands ranges from approximately 120 mm/year in the

northeast to 80 mm/year in the southwest, the majority of which falls during the winter months of December to March (Israel Meteorological Service, IMS). Direct precipitation is generally localized (5–15 km2) and of low intensity (<15 mm/hr), with rainfall originating generally from

the Mediterranean coast (Sharon 1972; Sharon and Kutiel 1986). Rare high intensity (>30

22 mm/hr) rainstorms originate to the south from the Red Sea. Rain from both systems can cause

flash floods (Sharon and Kutiel 1986: 286–287; Evenari et al. 1982). In general, large year-to-

year variations in precipitation are common, and recent trends (over the last 40+ years) show

that both annual precipitation and number of rain-days are decreasing (Ziv et al. 2013). In

addition to direct precipitation, dew and fog add the equivalent of another 5–10 mm per 100 m

of elevation (Kidron 1999). Runoff mechanics, dependent on geomorphology of the region,

vegetation, soil cover, slope, catchment size and shape, raise the seasonal precipitation yield of

unmodified (i.e., without built terraces or other agricultural installations) wadis and valleys to

an annual equivalent of 300–500 mm, resulting in rich seasonal flora (Evenari et al. 1982;

Shanan 2000: 89–92; Avni et al. 2012). This is more than sufficient for cereal agriculture

(~150–200 mm/year) or even small-scale horticulture (~300 mm/year) (Danin 1983; Wilkinson et al. 2014).

Temperature in the Negev Highlands is dependent on elevation, with summer months averaging 25–30°C and winter 10–15°C (IMS). Measurements (recorded at the Beersheva and

Sede Boqer stations from 1970–2002) of the aridity/humidity index P/PET (annual precipitation to potential evapotranspiration ratio) of the Negev Highlands is 0.16–0.06, classifying the area as arid bordering on the hyper-arid (Kafle and Bruins 2009: 65, 74).

Natural vegetation in the Negev Highlands is characterized by Saharo-Arabian flora, with a minor Irano-Turanian component (Zohary 1962; Danin and Plitman 1987; Danin 2004).

Desert-adapted perennial woody dicotyledons dominate the landscape, including

Chenopodiaceae, Zygophyllum dumosum, and Tamarix species; major annual grasses include

Ammochloa palaestina, Asteriscus pygmaeus, Cutandia memphitica, and Stipa capensis (Hillel

and Tadmor 1962: 34–36; Danin 1983: 39–45; Danin and Plitman 1987; Vogel et al. 1986).

Most Negev annuals bloom in the spring (March–April), and are dormant by the summer

23 months (June–August). The majority of both annual and perennial are edible to

ruminants, and as such, the naturally-irrigated valleys provide the good forage for herds

(Seligman et al. 1962; Avni 2005).

2.3 Paleoenvironment

In order to evaluate ancient settlement in the Negev, a short review of the data is necessary to

define the ancient environment. Local paleoclimatic studies of the Negev are rare due to the

lack of preserved paleoenvironmental proxies for the Holocene, an issue common in arid environments. Some earlier work in the area was conducted utilizing isotopic studies (13C and

�18O) of land snail shells (Goodfriend 1988, 1990, 1991) which proved to be a problematic

proxy due to specific aspects of biomineral formation in snail shells. More recently through

palynological studies of degraded animal dung in rockshelters (Babenko et al. 2007). Although

these studies are somewhat hampered by low chronological resolution, they all show a general

trend of an ameliorated climate in comparison to today during the fourth and third millennia

BCE (EB–IBA), as well as decreasing precipitation at the end of the third millennium BCE.

Higher-resolution regional proxies including the reconstructed Dead Sea levels and

sedimentary record have been studied extensively in the past two decades (Enzel et al. 2003;

Bar-Matthews et al. 2003; Migowski et al. 2006; Litt et al. 2012; Langgut et al. 2014, 2015;

Kagan et al. 2015). In general they show similar trends, with ameliorated climate throughout the fourth and third millennia BCE. Extreme short-term climate oscillations are seemingly evident in Soreq Cave speleothems in the Judean foothills at the end of the EB around 2600

BCE (Laugomer et al. forthcoming) and during the IBA brief desiccation events are reported around ~2350 BCE (Langgut et al. 2015; Kagan et al. 2015) and ~2200 BCE (Frumkin 2009).

At the end of the IBA (~2000 BCE), drier climates prevail into the Middle Bronze Age

(Frumkin 2009; Langgut et al. 2015; Kagan et al. 2015).

24 In general, given the ameliorated paleoclimate at the beginning of the IBA there does not

appear to be any environmental conditions that might have precluded Negev seasonal

agriculture during the majority of the IBA (see full discussion in Dunseth 2013; Dunseth et al.

2016 [below]).

3 Settlement history of the Negev Highlands

3.1 Results of archaeological surveys in the Negev Highlands

The history of the arid Negev Highlands is characterized by sharp settlement oscillations (Fig.

2; Finkelstein 1995a; Rosen 2011b). Several periods, including the Early Bronze (EB),

Intermediate Bronze Age (IBA), a phase in the Iron Age and Byzantine–Early Islamic periods,

show strong evidence of human activity, while others conspicuously lack evidence for built

remains. Although the absence of built remains does not indicate the absence of human activity

in the area (Rosen 1992; Rosen et al. 2005), settlement was greatly diminished during ‘empty’

periods and likely limited to sparse nomadic pastoralism.

Hundreds of sites dating to the EB and IBA in the Negev Highlands, and Sinai have

been recorded in extensive archaeological surveys beginning early in the 20th century (e.g.,

Woolley and Lawrence 1914; Glueck 1942, 1953, 1955, 1956, 1957, 1958a,b, 1960, 1965;

Aharoni et al. 1960; Rothenberg 1970; Baron 1978) and increasingly systematically since the

1960s. More data is available due to the Negev Emergency Survey (1979–1989) (summaries

in Cohen 1999; also Rosen 2011b; Saidel and Haiman 2014), and long-term surveys in Jordan

(e.g., MacDonald et al. 1983; MacDonald 1988, 1992, 2015; Hauptmann 2007; Levy et al.

2014) and Sinai (Bar-Yosef and Phillips 1977; Oren 1973, 1989; Oren and Yekutieli 1990;

Eddy and Wendorf 1999; Beit Arieh 2003).

25 1600 1513

1400

1200

1000

800 672

600 511 516 388 400 288 277

200 116 74 66 13 12 24 2 1 2 5 0 35 0 LB IAI PN MB IAII IBA EBI PPN Chalc Paleo EBII-III Persian Epipaleo Nabatean E. IslamicE. Hellenistic LateRoman Middle Ages Middle L. Byz/E. Islamic

Figure 2: Histogram of sites in the Negev Highlands by period according to survey. Survey maps included: 160, 163, 164, 166, 167, 168, 169, 170, 173, 174, 177, 178, 193, 194, 195, 196, 198, 199, 200, 201, 203, 204, 224, 225. All data collected from Archeological Survey of Israel website: https://survey.antiquities.org.il/.

26

Comprehensive data from the Negev Emergency Survey in Israel and those in Jordan are

available publicly at https://survey.antiquities.org.il/ and https://daahl.ucsd.edu/DAAHL/

respectively. Conventionally, the results of these surveys showed peaks during the EB II and

IBA based on pottery typologies, with a lacuna of activity between them during the EB III

(Haiman 1989; Finkelstein 1995a; Rosen 2011b). However, there were hints of activity during

the ‘empty’ periods based on emerging radiocarbon data, among other considerations (Sebanne et al. 1993; Avner et al. 1994; Avner and Carmi 2001).

Because of the extensive surveys, the broad settlement distribution of Negev IBA sites is generally well understood. It is clear that the central Negev Highlands specifically show the greatest number and large archaeological sites during the IBA, with the regions to the north and south showing less evidence for activity (Rosen 2011b). Notably, the Beersheva–Arad

Valley shows little to no evidence for activity during the IBA, (e.g., Govrin 2002, 2016), although work in the area is incompletely published.

3.2 Origin and nature of IBA Negev settlement

IBA sites in the Negev specifically have been variously defined according to social, economic,

technological and cultural definitions (e.g., Dever 1980, 1992, 1995, 2014; Cohen 1992, 1999;

Finkelstein 1995a; Haiman 1996). In general, these have been thoroughly reviewed most recently by Palumbo (2001, 2008), and Gidding (2016).

From a desert perspective, the most important settlement models in brief are discussed in the subsequent sections.

3.2.1 Kochavi (1967, 2009)

Origin: Based on what was published at the time, Kochavi (1967) suggested a demographic

migration of northern Kurgan groups from the Caucasus, leading to the destruction of EB sites

27

across the ancient Near East, and resulting in the foreignness of IBA culture in the southern

Levant. However, he did not interpret the Negev population as a foreign element, as EB cultural

traditions continue in features such as rectilinear domestic structures and red-slipped and

burnished pottery (Kochavi 2009).

Nature of settlement: based on his own excavations at Har Yeruham, some regional survey and

analysis of published sites, Kochavi noted two types of Negev settlement. The first was represented by Har Yeruham Stratum II, which suggested a small defensive community subsisting mainly on pastoral nomadism, seasonal agriculture, and limited industry—small- scale pottery production as well as copper smelting/trade. The later Stratum I showed a

settlement based only on herding, similar to the smaller sites throughout the Negev Highlands.

3.2.2 Dever (1970, 1971, 1973, 1980, 1985, 1992, 1995, 2014):

Origin: Dever’s models developed over time. Originally, he moderately supported a version of

the ‘Amorite Hypothesis’ emphasizing a northern pastoralist ‘invasion’ through Jabal Bishri in

central Syria (Dever 1970, 1971, 1973). Eventually, he abandoned this in support of a

‘dimorphic pastoralist model’ resulting from the abandonment of the urban EB III populations

(Dever 1980), greatly influenced by Liverani (1970, 1973) and Rowton (1977), among others.

In this view, all (or nearly all) formerly urban inhabitants became (agro)pastoral nomads, spreading throughout Cisjordan and into the Negev during winter months. This general model was later tempered by the excavation of IBA urban and rural sites (e.g., Richard 1980; Falconer and Magness-Gardiner 1984; Richard and Boraas 1988; Palumbo 1990; Esse 1991), which he conceded show a more varied adaptation to de-urbanization. However, his views for the origin of Negev settlement specifically did not substantially change.

Nature of settlement: based on his excavations at Jebel Qa ̔aqir (London 1985; Dever 2014)

and Be’er Resisim (Cohen and Dever 1978, 1979, 1981; Dever 2014), Dever suggested that

28

Negev settlements were mainly transhumant nomadic communities that followed winter

pasture through the Judean Hills, Transjordan and the Negev. Subsistence was based primarily

on sheep/goat herding, seasonal agriculture and perhaps trade. Dever did not particularly

differentiate between the types or sizes of sites in the Negev.

3.2.3 Cohen (1983, 1986, 1992, 1999):

Origin: Cohen’s model of IBA settlement has been described as ‘idiosyncratic’ by Dever

(2014: 221, n.2). In his view, Negev IBA settlement was related to the biblical Exodus and the

movement of pastoralists from (Cohen 1983). Although he suggested that the migration

of pastoral nomads was limited to the boundaries of the Negev system, he paradoxically also

believed that this same group spread IBA culture and tradition to the rest of the sites in modern

Israel and Jordan (Cohen 1992). For many reasons which stem from archaeology and biblical

exegesis, this model was never accepted.

Nature of settlement: Based on the extensive results of the Negev Emergency Survey and his

own excavations at dozens of sites in the Negev, Cohen identified a number of different sites,

including not only central and small sites, but also specifically temporary encampments and

cave dwellings. In this model large central sites (sites larger than 0.3 hectares) were permanent

settlements, with inhabitants mainly focused on herding, seasonal agriculture and trading of

copper and desert goods (e.g., Red Sea shells, bitumen, etc.). The smaller sites (< 0.3 hectares)

were merely grazing settlements, established by a subset of the permanent populations at the central settlements. According to this model, migration occurred only within the boundaries of the Negev, rather than long-range connections to areas further north and to the west (Cohen

1992).

29

3.2.4 Finkelstein (1989, 1991a, 1995a,b)

Origin: In general, Finkelstein, characterizes the human history of Negev as a series of settlement oscillations that reflect sociopolitical changes in the surrounding regions

(Finkelstein 1988, 1989, 1995a,b; Fantalkin and Finkelstein 2006; Finkelstein et al. 2018).

Specific to the third millennium BCE, Finkelstein emphasized the deterioration of the EB III urban system as the collapse of the specialization system (i.e., separate agricultural, pastoral and industrial systems), which forced a ‘multimorphic’ rather than Dever’s ‘dimorphic’ response. Some of this resulted in the (semi-)sedentarization of pastoral nomads, which then become visible in the archaeological record, especially in the arid marginal regions (Finkelstein

1995a: 97–100). These groups were also supported by trade, specifically copper, as in earlier

(e.g., EB II at Arad, Finkelstein 1991) and later periods (e.g., IA, Finkelstein 1988, 1995b).

Nature of settlement: Finkelstein’s model generally followed those of Dever and Kochavi, suggesting that desert inhabitants subsisted on animal husbandry and seasonal agriculture

(below). In addition, he suggested that in the absence of a larger urban power, in the IBA Negev populations might also be controlling the mining and trade of copper (Finkelstein 1995a: 100–

101).

3.2.5 Haiman (1996)

Origin: Based especially on the contrast between the distribution of EB and IBA sites and distinct differences in their material assemblages in the Negev—in contrast to a clear continuation of EB III culture in the IBA in Jordan—Haiman identified the IBA settlement as a new population from south Jordan that replaced the existing desert society (Haiman 1996:

16).

Nature of settlement: based on a survey of data newly available at the time, Haiman (1996) identified two complementary desert elements: central and small temporary sites. Based on

30

differences in site distribution, architectural layouts, and material culture, Haiman suggested the two main type sites represented different subsistence economies:

Large central sites were settlements (> 2 hectares) assumed to be permanent sites. Central sites

were divided into three subgroups, named after type sites: the Har Yeruham group

(characterized by attached rectangular structures, forming a sort of ‘wall’ around the site); the

Nizzana group (groups of large rooms and courtyards, reminiscent of EB desert sites); and the

more common Ein Ziq group (Haiman 1996: 3–5). The ‘Ein Ziq group was characterized by

large groups of single-room small circular structures, many built near perennial water sources

(more below). The lack of open courtyards or animal pens led Haiman to suggest that pastoral

nomadism was not conducted at central sites. Likewise, the few sickle blades published to date

provided little evidence that farming was practiced. Substantial copper evidence, including

ingots and finished products, as well as numerous pounders and grinding stones, led Haiman

to suggest these sites were established and settled by copper specialists (producers/processors)

and traders. The distribution of central sites along an east-west vector from copper mines in

Faynan towards Sinai, led him to reconstruct First Intermediate Egypt as the main market for

copper. Later provenance studies of copper ingots at the Negev sites (Segal et al. 1996–1997)

and Khirbet Hamra Ifdan (Hauptmann et al. 2015) showed that the IBA ingots originated from

copper ores in Faynan, confirming this connection.

Small temporary sites, characterized by a few rooms encircling larger built open courtyards,

Haiman argued were inhabited by pastoral nomadic groups engaged mainly in animal

husbandry and some opportunistic farming (1996:5–10). Haiman suggested these sites supported the permanent central sites with access to meat and animal products. Haiman’s model generally became the most accepted for Negev IBA settlement processes.

31

3.2.6 Rosen (2011a,b, 2016, among others)

Origin: Rosen views Negev settlement over the longue durée as both a series of fluorescence

and decline (e.g., Rosen 2011a) and a long-term continuity (e.g., Rosen 2011b). Rosen updated

the ‘Timnian’ desert pastoral complex of Kozloff (1974) and Rothenberg and Glass (1992) that included long-lasting lithic, architectural, cultic and material culture traditions based on sheep and goat herding. Importantly, Rosen sees clear shared cultural continuity of a southern indigenous desert society throughout the (south) Negev, south Sinai and south Jordan/northern

Arabia that can be traced back at least to the 6th millennium BCE. His ‘terminal Timnian’ is

equivalent to the IBA, and is the last phase of this cultural complex in the Negev.

Nature of settlement: As described above, the Timnian complex is based on sheep and goat

herding for subsistence. In general, Rosen suggests farming was at most a minor addition for

the desert pastoralists (Rosen and Vardi 2014: 336–337). Importantly, Rosen sees lithic cultural

affinities of central sites not only to each other (e.g., Ein Ziq, Be’er Resisim, Vardi 2005, Vardi

et al. 2007), but also to smaller sites from earlier periods—indeed, there is no drastic difference

in the ad hoc lithic assemblages of (earlier EB I–II) small or IBA large sites (Rosen and Vardi

2014: 337–338). Rosen emphasizes copper exploitation as an important aspect of the Timnian

complex from at least the Chalcolithic onward, although to him this is a supporting system to

pastoral nomadism rather than a central aspect of the subsistence economy.

3.2.7 Gidding (Gidding 2016; also Ben-Yosef et al. 2016)

Origin: Based on excavations at Khirbet Hamra Ifdan (by himself and others) and a regional

analysis of Faynan and the Negev Highlands, Gidding showed a long continuity of copper

exploitation throughout the fourth and third millennia BCE. He somewhat follows Rosen’s

autochthonous desert model, although with a focus on copper production as a major industry

and driver of settlement and subsistence. Importantly, Gidding challenges the break in Negev

32

occupation assumed by others (above). Instead he reconstructs a parabolic trend of copper

activity, increasing during the EB I–II, peaking during the EB III–early IBA, and then decreasing significantly after the middle of the IBA (c. 2200 BCE). He divides this into alternating phases of dispersed (EB I–II and second half of the IBA) and centralized (EB III– early IBA) copper production, utilizing Faynan as a case study.

Nature of settlement: At Khirbet Hamra Ifdan, Gidding reconstructs a highly specialized

copper industry, one that is only partially involved in pastoralism (cf. Muniz 2007). In the

Negev, Gidding reconstructs (what he terms the desert ‘mega-sites’ of Ein Ziq and Be’er

Resisim) as settlements of itinerant traders moving copper from the production centers in

Faynan towards Egypt, a view partially based on the research presented in this dissertation.

3.3 Types of archaeological sites in this study

For this study, I rely on a simplified version of the site typology of Haiman (1996), dividing

Negev archaeological sites into two main classes: large central and small (Haiman’s

‘temporary’) sites. Generally, both site types have been assumed to be populated by the same

desert communities and/or social groups (see above).

Central sites cover between 0.3–2 hectares and in general, contain large (c. 100–200) numbers

of stone-built circular structures. Notably, these sites lack built courtyards or open structures

traditionally identified as “animal pens” (Haiman 1996: 3). There is no obvious internal plan

or organizational division to this type of settlement (Cohen 1992; Haiman 1996 contra Dever

1985), although small clusters of structures have been interpreted as possibly representing kin

groups based on ethnographic analogies (Dever 1985; Dunseth et al. 2018). IBA sites

belonging to this type are limited to four clear examples: Be’er Resisim (Cohen and Dever

1978, 1979, 1981; Cohen 1999; Dever 2014), Be’er Hayil (Baumgarten 1993; Cohen 1999:

130) and the two sites investigated in this dissertation, Mashabe Sade (Cohen 1999: 117–130),

33

and Ein Ziq (Cohen 1999: 137–188). All four sites were excavated to some extent by R. Cohen or his associates in the 1980s. Central sites are unique in the settlement history of the Negev and southern regions in general. Sites with markedly different layouts such as Har Yeruham

(Kochavi 1967), Horvat Avnon and Nahal Nizzana (Cohen 1999) were not investigated in this study.

Small sites are generally characterized by their limited size (up to 0.2 hectares) and are composed of a few stone-built units, often encircling or connected to large courtyards (Haiman

1996). These types of sites make the majority of archaeological sites in the Negev Highlands during the IBA. This plan type has long-standing parallels throughout the history of the region, including during the preceding EB II, (e.g., Ramat Matred 9, Cohen 1999: 61), Iron Age II

(Ramat Matred, Cohen and Cohen-Amin 2004: 58-62), and in pre-modern Bedouin encampments (Musil 1908: 131, 1928; Saidel and Erickson-Gini 2014). Numerous surveyed and excavated IBA small sites have been published, including various sites in the central and western Negev (Cohen 1999; Saidel and Haiman 2014) and Sinai (Clamer and Sass 1977; Oren and Yekutieli 1990). Some particularly important to this study include Rekhes Nafha 396

(Saidel 2002), Rogem Be’erotayim (Saidel et al. 2006), and the Camel Site (Rosen 2011a).

For this dissertation I looked to answer the following core questions about subsistence:

1) What direct subsistence strategies supported central and small sites?

2) Were all sites engaged in livestock herding and animal husbandry?

3) Is there evidence for dry farming during the IBA in the Negev Highlands?

4) To what extent was copper involved in supporting Negev settlement?

5) Is there evidence for metallurgical activities – production or processing – at large Negev

Highland sites (cf. Haiman 1996; Yekutieli et al. 2005)?

6) What evidence is there for trade?

34

3.4 Chronology of the Intermediate Bronze Age

3.4.1 Traditional chronology of the Intermediate Bronze Age

Until very recently, a consensus dated the IBA to 2300–2000 BCE, corresponding closely to

the fall of the Old Kingdom and the First Intermediate Period in Egypt (Stern 2008: 2126; see

Dever 1980; Finkelstein 1989; Haiman 1996; Palumbo 2001; D’Andrea 2012, among many

others). Attempts were made to subdivide this period through assumed chronological division of ceramic families (Dever 1973, 1980) and more recently, developments in ceramic

technology (D’Andrea 2012). However, these studies were hampered by assumptions of

unproven settlement gaps (see Finkelstein 1989: 133), misidentified ceramic relationships (see

petrographic results by Goren 1996) and poorly or unpublished ceramic assemblages.

Complicating the construction of ceramic sequences, very few sites experienced occupation

from the late EB III to the IBA. Notable exceptions were limited (arguably) to larger sites in

modern Jordan, such as the ‘urban’ Khirbet Iskander (Richard et al. 2010), Bab edh-dhra (Rast

and Schaub 2003) and the copper production center at Hamra Ifdan (Adams 2000; Levy et al.

2002). All three of these sites showed that the chronological aspect of the ceramic IBA families

model was untenable, and that the differences showed only regional variation rather than

temporal division. Published sites with multiple IBA phases were even more rare. In the Negev

Highlands, only Har Yeruham has more than one discernable stratigraphic IBA phase (Kochavi

1967; Cohen 1999: 111–115), although there was discussion of ‘horizontal stratigraphy’ at large sites like Ein Ziq and Be’er Resisim.

3.4.2 Radiocarbon investigations at IBA sites in the Negev and surrounding regions

Radiocarbon investigations at IBA sites in the southern deserts of the Negev, Sinai and Jordan were and remain relatively rare. The majority of determinations come from the two central sites of Ein Ziq and Be’er Resisim, as well as the large EB/IBA site of Har Dimon (Table 1). These

35

investigations generally indicated dates earlier than the accepted chronological schemes. More dates were published from the copper production sites in the Arabah, including the main center at Hamra Ifdan, as well a smaller site at Ein Yahav, and two copper sites attributed to the IBA in Timna (Table 1). Radiocarbon dates from poorly defined or unpublished contexts also exist for a number of sites in the Uvda Valley (Avner et al 1994; Avner and Carmi 2001; Avner

2006). Although, these and other Negev and Sinai dates were attributed to other periods according to the traditional chronological frameworks, they importantly show the absence of any gap in Negev activity over the late fourth and third millennia (e.g., Avner and Carmi 2001;

Avner 2006; also discussion in Gidding 2016).

With the exception of Mushabi 103 in central Sinai, smaller sites attributed to the IBA were not radiocarbon dated, often due to lack of datable remains (Table 1). However, a few small sites dating more broadly to the third millennium BCE were investigated notably the Camel

Site, Har Horesh, and a few sites in South Sinai (Table 1).

In general, the absolute chronology of small sites—especially in relation to where exactly they fit during the IBA—is poorly understood.

Recently, a comprehensive reevaluation of published radiocarbon determinations from only secure archaeological Early Bronze and IBA contexts convincingly redated the IBA framework to 2500–1950 BCE (Regev et al. 2012: 557, 559–560, Fig. 12). This places the beginning of the Negev settlement peak within the timeframe of the powerful Egyptian Old Kingdom (Late

4th–6th Dynasties) (Malek 2000; Bronk-Ramsey et al. 2010), as well as the strong urban system in Northern Syria. This time-shift has significant implications regarding socioeconomic relationships between the Negev Highlands, the Southern and Northern Levant, and Egypt, and forces a reevaluation of the models above.

36 Table 1: Radiocarbon determinations from Early and Intermediate Bronze Age sites in the Negev and surrounding regions.

Excavator Calibrated Date Calibrated Date Site Lab Number ID Age ± Context Source Attribution BCE (1 σ) BCE (2 σ) Negev Highlands Small Sites Camel Site Rta-3083 charcoal 4345 65 EBII L.041 Square: J29c upper 3082 ( 4.9%) 3068 3327 ( 7.5%) 3218 Rosen 2011a: deposit (immediately outside 3026 (63.3%) 2896 3176 ( 0.8%) 3160 Table 4.1 of the hearth dated by Rta- 3121 (87.1%) 2875 2043) Rta-2043 charcoal 4115 50 EBII L.041 Square: M29b lower 2858 (17.7%) 2810 2876 (93.7%) 2570 deposit; hearth (hearth, from 2752 (10.3%) 2722 2514 ( 1.7%) 2501 lower organic layer), 30cm 2702 (31.1%) 2617 below surface 2610 ( 9.0%) 2582

Rta-3082 charcoal 3235 55 (later intrusion) L. 032 Square I31c upper 1606 (10.3%) 1584 1633 (95.4%) 1410 deposit 1558 ( 1.4%) 1554 1545 (56.4%) 1440 Har Horesh (Site GrA-28787 ostrich egg 4570 40 EBII L. 80+87; floor ('leveling fill') 3487 ( 5.5%) 3472 3496 ( 9.2%) 3460 Saidel and Haiman 23) shell makeup above bedrock in 3372 (30.9%) 3330 3376 (40.6%) 3264 2014: 30, Fig. 2.66 room surrounding central 3216 (16.6%) 3182 3241 (45.5%) 3103 courtyard; sealed by 0.05 m 3158 (15.2%) 3124 ash accumulation Nahal Boqer 66 RTD-8673 animal bone 4473 35 EBIb-II L. 16/NB/2; grey dung and 3330 (45.4%) 3214 3341 (86.0%) 3081 Dunseth et al. (n.d.) microcharcoal-rich sediment 3186 ( 9.9%) 3156 3069 ( 9.4%) 3026 2017: Table 3 directly on bedrock floor, 3126 (12.8%) 3092 sealed by Phase II architecture RTD-8675 animal bone 4414 36 EBIb-II L. 16/NB/2; grey dung and 3096 (46.6%) 3006 3323 (12.7%) 3234 (B-sized microcharcoal-rich sediment 2988 (21.6%) 2932 3172 ( 0.9%) 3162 medium directly on bedrock floor, 3117 (81.8%) 2916 mammal) sealed by Phase II architecture RTD-8672 short-lived 4346 25 EBIb-II L. 16/NB/2; grey dung and 3010 (28.3%) 2978 3022 (95.4%) 2902 twig (Salsola microcharcoal-rich sediment 2960 ( 6.4%) 2951 tetranda) directly on bedrock floor, 2942 (33.5%) 2908 sealed by Phase II architecture RTD-8671 short-lived 4331 25 EBIb-II L. 16/NB/2; collapse mixed 3009 (19.7%) 2984 3016 (95.4%) 2896 twig with grey dung and charcoal- 2936 (48.5%) 2900 (Zygophyllum rich sediment above bedrock sp.) floor;

37 RTD-8676 animal bone 4077 35 EBIII-IBA L. 16/NB/8; grey dung and 2836 ( 9.4%) 2816 2859 (15.5%) 2809 (B-sized ash-rich sediment deposited 2668 (53.4%) 2570 2752 ( 4.8%) 2722 medium on bedrock floor 2515 ( 5.4%) 2501 2701 (63.5%) 2558 mammal) 2536 (11.5%) 2491

RTD-8674 animal bone 3966 33 EBIII-IBA L. 16/NB/8; grey dung and 2566 (32.4%) 2524 2574 (85.0%) 2432 (B-sized ash-rich sediment deposited 2497 (35.8%) 2462 2424 ( 4.2%) 2401 medium on bedrock floor 2381 ( 6.1%) 2348 mammal) RTD-8670 short-lived 3952 36 EBIII-IBA L. 16/NB/8; grey dung and 2564 (17.1%) 2532 2571 (23.7%) 2512 twig ash-rich sediment deposited 2495 (33.3%) 2451 2504 (71.4%) 2339 (Zygophyllum on bedrock floor 2420 ( 6.4%) 2405 2313 ( 0.3%) 2310 sp.) 2378 (11.4%) 2350

Negev Highlands Large Central Sites

Be'er Resisim RT-2346 ostrich egg 4085 70 IBA Construction 8c; Area A; 2858 (13.6%) 2810 2872 (95.4%) 2484 Carmi and Segal shell L.8064, B.268 2750 ( 7.0%) 2722 1992: 126; Segal 2700 (41.2%) 2566 1999: 338 2522 ( 6.3%) 2498

RT-2347 ostrich egg 4050 50 IBA Construction 42, Area B; 2832 ( 3.8%) 2820 2858 (10.1%) 2809 shell L42010; B.11 2632 (64.4%) 2488 2752 ( 3.6%) 2722 2701 (81.7%) 2469

RT-2468 ostrich egg 3930 40 IBA Construction 10; Area B; 2477 (46.6%) 2393 2565 ( 6.6%) 2532 shell L.10015; B14 2386 (21.6%) 2346 2496 (88.8%) 2294 Ein Ziq RT-885A charcoal 3960 90 IBA Area B; L. 53; B. 282 2578 (63.5%) 2333 2858 ( 3.3%) 2810 Segal and Carmi 2325 ( 4.7%) 2300 2750 ( 1.3%) 2722 2004: 145 2700 (90.8%) 2200 RT-885B1 charcoal 3880 60 IBA Area B; L. 79; B. 422 2462 (68.2%) 2292 2558 ( 1.5%) 2536 Segal 1999: 338 2491 (92.5%) 2196 2168 ( 1.4%) 2148 RT-885B charcoal 3850 50 IBA Area B; L. 79; B. 422 2453 (11.0%) 2419 2468 (93.5%) 2198 2406 (10.3%) 2376 2166 ( 1.9%) 2150 2350 (30.6%) 2274 2255 (16.2%) 2209

RTD-7678 short-lived 3857 39 IBA Area J; L. 14/J/16; Refuse pit 2454 (14.0%) 2418 2462 (78.6%) 2267 Dunseth et al. twig (Anabasis outside Structures 14/J/19 and 2406 (13.2%) 2376 2260 (16.8%) 2206 2017: Table 1 sp.) 14/J/21 2350 (34.6%) 2282 2248 ( 6.4%) 2232

RTD-7679 short-lived 3846 39 IBA Area J; L. 14/J/16; Refuse pit 2432 ( 3.2%) 2423 2460 (95.4%) 2202 twig (Retama outside Structures 14/J/19 and 2402 ( 8.3%) 2380 raetam) 14/J/21 2348 (36.5%) 2274 2256 (20.3%) 2208

38 RTD-7680 short-lived 3853 37 IBA Area J; L. 14/J/21; PT2; 2452 (11.2%) 2420 2460 (95.4%) 2206 twig Anabasis Beaten-earth floor of 2405 (11.0%) 2378 sp. Structure 14/J/21 2350 (34.8%) 2278 2250 ( 8.3%) 2230 2220 ( 3.0%) 2212

RTD-7681 short-lived 3834 37 IBA Area A; L. 14/A/4; Mixed 2344 (68.2%) 2204 2458 (94.1%) 2198 twig (Retama collapse material within 2161 ( 1.3%) 2152 raetam) Structure 14/A/4 RTD-7682 short-lived 3861 40 IBA Area A; L. 14/A/2; hearth 2455 (15.2%) 2418 2464 (80.3%) 2268 twig (Retama within Structure 14/A/2 2408 (14.6%) 2374 2260 (15.1%) 2206 raetam) 2368 (38.4%) 2286 RTD-8310 short-lived 3866 26 IBA Area F; L. 15/F/TP3; 2452 (15.5%) 2420 2462 (90.4%) 2280 twig (n.d.) Secondary ash deposit outside 2406 (15.5%) 2377 2250 ( 4.1%) 2231 of Cohen’s Structure 73 2350 (37.2%) 2290 2218 ( 0.9%) 2213 RTD-8311 short-lived 3867 25 IBA Area K; L. 15/K/8; Stone- 2453 (16.0%) 2419 2462 (91.5%) 2282 twig (n.d.) lined hearth, sealed by 2406 (15.8%) 2376 2249 ( 3.4%) 2232 collapse of Structure 15/K/5 2350 (36.3%) 2291 2218 ( 0.5%) 2214 RT-2514 charcoal 3700 45 IBA Area A; L. 13; B 463; Floor 2191 ( 4.3%) 2180 2266 ( 0.4%) 2261 Carmi and Segal (Quercus sp.) of Cohen’s Room 13 2142 (63.9%) 2030 2206 (95.0%) 1951 1992: 126 Har Dimon RT-1556 charcoal 4660 55 IBA Eastern complex; Area A, 3516 (59.9%) 3398 3632 (12.0%) 3561 Segal and Carmi Room 2, Basket 1093. 3384 ( 8.3%) 3368 3536 (83.4%) 3349 1996: 94 charcoal RT-1557 charcoal 3915 50 IBA Western complex: Area A, 2437 ( 5.6%) 2420 2466 (92.3%) 2196 Room 17, Basket 1004 2404 ( 9.3%) 2378 2170 ( 3.1%) 2147 2350 (33.0%) 2270 2260 (20.4%) 2206

RT-1558 charcoal 3845 50 IBA Western complex: Area A, 2472 (65.9%) 2338 2566 ( 5.5%) 2523 Wall 1, Basket 1140 2316 ( 2.3%) 2310 2497 (86.8%) 2278 2250 ( 2.3%) 2229 2220 ( 0.8%) 2211

Southern Coast Tell es-Sakan Beta-163590 charcoal 4280 80 EBIII Area C; Str. C-4, Layer 3; L. 3022 (53.0%) 2861 3264 ( 1.0%) 3241 Regev et al. 2012: 401; Middle EBIII; charcoal 2808 (12.4%) 2756 3104 (94.4%) 2620 Table 1 2719 ( 2.8%) 2705 Beta-163591 charcoal 4140 70 EBIII Area C; Str. C-4, Sq. AD 44; 2871 (20.4%) 2801 2892 (92.8%) 2566 L. 3021; Middle EBIII; 2780 (47.8%) 2626 2522 ( 2.6%) 2498 charcoal Beta-163587 charcoal 4090 40 EBIII Area C; Str. C-4, Layer 2; Sq. 2850 (14.6%) 2813 2866 (19.3%) 2804 AE 42; L. 3071; Middle 2740 ( 2.7%) 2730 2776 (69.2%) 2562 EBIII; charcoal 2693 ( 1.9%) 2686 2534 ( 6.9%) 2492 2680 (49.0%) 2573

39 Beta-163589 charcoal 4020 40 EBIII Area C; Str. C-4; AD-AE 44; 2576 (68.2%) 2484 2833 ( 1.7%) 2819 L. 3038; Middle EBIII; 2660 ( 0.8%) 2650 charcoal 2634 (92.9%) 2465

Beta-163588 charcoal 3860 90 EBIII Area C; Str. C-4; Sq. AE 45; 2464 (54.2%) 2268 2571 ( 4.6%) 2512 L. 3019; floor a; Middle 2260 (14.0%) 2206 2504 (87.1%) 2118 EBIII; charcoal 2096 ( 3.8%) 2040 Dead Sea Bab edh-Dhra SI-2877 wood 7235 215 EBIa or EBIII Area F3, (Field F3) L. 13, 6362 (10.4%) 6280 6502 (95.4%) 5709 Weinstein 1984; fragments in (?) occupational debris (?); 6274 (52.4%) 5971 2003 soil + ash charcoal 5954 ( 5.5%) 5912 SI-2502 charcoal, 6615 145 EBIa or EBIII Area F3, L. 13, occupational 5706 ( 3.5%) 5684 5834 ( 0.2%) 5826 Weinstein 2003 charred wood, (?) debris (?); charcoal 5676 (62.2%) 5467 5811 (95.2%) 5304 burnt bone 5402 ( 2.5%) 5388

SI-3310B 6415 110 EBIa/b Tomb A 100, Ch E 5485 (66.8%) 5296 5614 ( 1.5%) 5587 Weinstein 1984 5242 ( 1.4%) 5234 5568 (90.9%) 5206 5163 ( 1.8%) 5119 5108 ( 1.2%) 5079 SI-2869 charcoal 5090 85 EBII-IBA Field X.1, L. 28 EB II–IVA 3971 (68.2%) 3789 4049 (94.5%) 3692 Weinstein 1984 3682 ( 0.9%) 3664 SI-2876 charcoal 5080 90 EBIII Field XII.2. L. 13, EB II–III; 3970 (68.2%) 3776 4045 (95.4%) 3661 Rast and Schaub in 2003 attributed to EB III; 2003; Weinstein charcoal 1984 SI-4134 charcoal 5070 85 EBIII Field XIV.4, L. 9, 3962 (68.2%) 3781 4040 ( 2.4%) 4014 Weinstein 1984 occupational surface EB III 4002 (91.4%) 3692 3684 ( 1.6%) 3662 SI-4135 charcoal 5030 75 EBIII Field XII.5, L. 24, beam on 3943 (34.4%) 3853 3966 (95.4%) 3662 floor EB III; charcoal 3848 (30.7%) 3762 3724 ( 3.1%) 3715 SI-2871 charcoal + soil 5000 65 EBIb (?) Area F3, L. 9, pit 3936 (22.9%) 3872 3948 (95.4%) 3662 Weinstein 2003 3810 (45.3%) 3703 SI-3310A 4630 90 EBIa/b Tomb A 100, Ch E 3626 ( 5.6%) 3596 3635 (76.0%) 3262 3526 (55.6%) 3334 3249 (19.4%) 3099 3212 ( 3.9%) 3190 3153 ( 3.2%) 3134

Beta-134011 charcoal 4600 70 EBIb Field XII.4; Str IV; L40; ash 3516 (31.9%) 3397 3626 ( 2.5%) 3596 Weinstein 2003 and charcoal in sand and 3385 (18.6%) 3327 3526 (92.9%) 3096 gravel abutting wall 35, 3218 ( 9.3%) 3176 below L.i 36 and 41; charcoal 3160 ( 8.4%) 3121

40 Bab edh-Dhra Beta-134014 charcoal 4510 60 EBII-III Field XIV.3; L226; mudbrick 3346 (12.6%) 3308 3486 ( 0.6%) 3474 Weinstein 1984 debris in slope collapse area; 3301 ( 6.1%) 3282 3372 (94.8%) 3016 EB II–III; charcoal 3277 ( 3.9%) 3265 3240 (45.6%) 3105

Beta-134013 charcoal 4480 40 EBIb Area J; Str IV; L22; village 3331 (45.3%) 3214 3348 (87.3%) 3082 Weinstein 2003 house, soil layers between 3186 (10.9%) 3156 3068 ( 8.1%) 3026 walls L.i 11 and 12; charcoal 3127 (11.9%) 3095 SI-2501 4420 80 EBII-IBA Charnel house A 55, doorway 3320 ( 9.8%) 3272 3340 (95.4%) 2908 Weinstein 1984 EB II–IVA 3266 ( 7.3%) 3236 3169 ( 1.0%) 3164 3112 (50.1%) 2921

Beta-134012 charcoal 4380 60 EBIb Field XII.9; Str IV; L3; ash 3089 (15.3%) 3046 3328 (12.8%) 3218 layer with bricks and bone, 3036 (52.9%) 2913 3178 ( 1.7%) 3158 below L.i 1 and 8, associated 3122 (80.9%) 2892 with Wall 25; charcoal M-2037* 4350 180 EBII or III Charnel house A 51, fl EB II– 3341 (63.8%) 2866 3516 ( 4.2%) 3397 Weinstein 1984; III 2804 ( 4.4%) 2762 3385 (90.0%) 2564 2003 2532 ( 1.2%) 2494 SI-2870 charcoal 4320 85 IBA (early) Field X.3, L. 29, late EB III; 3093 (68.2%) 2876 3332 ( 9.0%) 3214 Weinstein 1984 in 2003 attributed to EB IV; 3188 ( 1.8%) 3154 charcoal 3130 (73.9%) 2840 2814 (10.6%) 2678

SI-2874 4320 65 EBII-IBA Charnel house A 55, NE 3022 (68.2%) 2886 3317 ( 1.4%) 3273 Weinstein 1984 corner EB II–IVA 3266 ( 1.7%) 3236 3168 ( 0.1%) 3164 3111 (87.9%) 2858 2810 ( 3.5%) 2753 2721 ( 0.8%) 2702

SI-2503 powdery 4245 80 EBIII Field XII.2, L. 7, EB II–III; in 2926 (29.3%) 2838 3084 ( 1.1%) 3066 Weinstein 1984 wood, ash 2003 attributed to EB III 2815 (38.9%) 2674 3028 (94.3%) 2581 SI-2868 wooden beam 4205 85 EBIII Field XIII.1, L. 9 dest. debris, 2901 (20.8%) 2836 3011 (94.7%) 2570 Weinstein 1984 EB IB–III; in 2003 attributed 2816 (47.4%) 2668 2514 ( 0.7%) 2502 to EB III M-2036* cloth 4160 180 EBII or III Charnel house A8, entryway, 3005 ( 1.3%) 2990 3336 ( 4.3%) 3210 Rast and Schaub burnt cloth on floor 2930 (66.9%) 2470 3192 ( 1.2%) 3151 2003; Weinstein 3138 (89.2%) 2280 1984 2250 ( 0.5%) 2231 2218 ( 0.1%) 2213

Beta-134009 charcoal 4050 50 EBII (late) Field XVII.1; Str IIIA2; 2832 ( 3.8%) 2820 2858 (10.1%) 2809 L149; domestic occupation, 2632 (64.4%) 2488 2752 ( 3.6%) 2722 2701 (81.7%) 2469

41 surface under L.us 122; charcoal

Bab edh-Dhra Beta-134010 charcoal 4020 70 EBII (late) Field XVII.I; Str IIIA2; L143; 2834 ( 3.3%) 2818 2865 ( 8.1%) 2804 occupational surface; late EB 2662 ( 2.6%) 2648 2761 (87.3%) 2342 II 2636 (62.3%) 2464 SI-2499 4015 75 EBII-IBA Charnel house A 55, fl, 2836 ( 3.8%) 2816 2865 ( 7.9%) 2804 Weinstein 1984 opposite doorway EB II–IVA 2668 (64.4%) 2460 2761 (86.4%) 2334 2324 ( 1.1%) 2304

SI-2468 charcoal 3850 60 L. 49 2454 (11.7%) 2418 2472 (90.0%) 2189 Weinstein 1984 2407 (10.7%) 2376 2182 ( 5.4%) 2141 2350 (29.3%) 2274 2256 (16.5%) 2208

SI-2872 charcoal 3805 60 IBA Field X.3, L. 49, brick fall; in 2342 (68.2%) 2141 2462 (90.2%) 2128 Rast and Schaub 2003 attributed to EB IV 2088 ( 5.2%) 2046 2003; Weinstein 1984 Beta-134016 charcoal 3800 60 IBA Field XVI.4; Str I; L7; 2340 (68.2%) 2139 2460 (88.9%) 2123 Rast and Schaub mudbrick debris with ash 2092 ( 6.5%) 2042 2003 pockets; charcoal P-2573* olive pits 3770 60 IBA Olive pits from ash pit, Ph 3, 2290 (58.8%) 2131 2455 ( 2.5%) 2418 Weinstein 1984 Area X; sample from earliest 2086 ( 9.4%) 2051 2406 ( 2.5%) 2376 habitation Level E of walled 2350 (90.4%) 2024 town Area 10; Field X, Ph 3; olive pits Beta-134017 charcoal 3690 60 IBA Field XVI.I; Str I; L12; gray 2194 ( 6.2%) 2174 2278 ( 2.5%) 2250 bricky soil, below L.us 7; 2145 (55.7%) 2012 2229 ( 0.6%) 2220 charcoal 1998 ( 6.3%) 1978 2211 (92.3%) 1916 SI-2497 3680 60 Charnel house A55, NW 2188 ( 1.0%) 2184 2273 ( 1.2%) 2258 corner, left of doorway; 2141 (67.2%) 1974 2208 (94.2%) 1898 charcoal SI-2875 charcoal 3595 70 IBA Field X.3, L. 60, late EB III; 2116 ( 4.1%) 2098 2140 (95.4%) 1751 Weinstein 1984 in 2003 attributed to EB IV 2038 (61.2%) 1878 1838 ( 2.0%) 1828 1790 ( 0.8%) 1786

SI-4137 short-lived 4310 70 EBIII SE 3/1, L. 9 W of WL 4; 3080 ( 2.6%) 3070 3317 ( 1.4%) 3273 collected in flotation 3025 (65.6%) 2878 3266 ( 1.7%) 3236 3168 ( 0.1%) 3164 3112 (82.5%) 2848 2814 ( 9.7%) 2678

Numeira P-3454 short-lived 4180 60 EBIII SE 8/1, L. 12, roof fall and 2882 (15.3%) 2840 2900 (92.3%) 2617 Weinstein 1984 occupational debris; short- 2814 (52.9%) 2678 2610 ( 3.1%) 2582 lived

42 Numeira SI-4138 4130 70 EBIII NE 3/1, L. 15, destr. debris 2866 (19.9%) 2804 2888 (91.9%) 2565 2776 ( 2.0%) 2769 2525 ( 3.5%) 2496 2764 (46.4%) 2620

P-3367 short-lived 4090 70 EBIII NE 4/4, L. 16, destr. debris; 2858 (14.5%) 2810 2873 (95.4%) 2488 short-lived 2752 ( 7.8%) 2722 2701 (41.8%) 2568 2516 ( 4.2%) 2500

SI-4136 short-lived 4085 55 EBIII Se 3/4, L. 7, base of town 2852 (13.4%) 2812 2871 (18.3%) 2801 wall, destr. debris; short-lived 2744 ( 4.8%) 2726 2780 (77.1%) 2486 2696 (45.7%) 2569 2516 ( 4.3%) 2500

Khirbet Iskander Iskander-? wood (Olea) 3930 60 IBA Area C; Phase 2B; L. 2030, 2548 ( 2.6%) 2539 2576 (91.7%) 2276 Holdorf 2010 Room 221; debris layer with 2490 (61.3%) 2336 2253 ( 3.7%) 2210 pits, patches of burnt 2322 ( 4.3%) 2308 mudbrick, abundance of sherds; ash pocket AA-50178 charcoal 3975 43 IBA Area C; Phase 2 or 3A; L. 2570 (38.4%) 2513 2617 ( 0.6%) 2610 2043, Room 232; compact 2503 (29.8%) 2462 2581 (94.8%) 2343 layer with clumps of plaster, some mudbrick, and charred timber; mixed with destruction debris (debris = Phase 2) Sinai Central Sinai Mushabi 103 RT-447B charcoal 3800 330 IBA Structure 1, 2 or 3 (?) 2836 ( 1.2%) 2816 3320 ( 0.6%) 3272 Sass and Clamer 2671 (63.1%) 1870 3266 ( 0.4%) 3236 1977: 264, Table 1846 ( 3.9%) 1776 3113 (94.4%) 1433 67 Sinai S-1 (Rahaya DRI-3268 charcoal 4470 62 EB Compound Feature-1/2: 3334 (36.1%) 3212 3356 (89.4%) 3007 Eddy and Wendorf Site) Room 1, Stratum 3 (20-30cm) 3190 (10.2%) 3153 2987 ( 6.0%) 2932 1999: Fig. 9.13- 3134 (13.6%) 3086 9.15 3062 ( 8.3%) 3030 DRI-3269 charcoal 5518 121 Compound Feature 1/2: Main 4496 (68.2%) 4240 4652 ( 0.4%) 4642 Enclosure, Lower 4616 (95.0%) 4048 Component: F-3 hearth, Stratum V (80-85 cm) DRI-3270 charcoal 4350 73 EB Compound, Feat. 4: 3088 (10.5%) 3054 3336 (11.5%) 3210 Southernmost Satellite Room: 3032 (57.7%) 2894 3192 ( 2.6%) 3151 F-4 room, Stratum II (20-30 3138 (81.3%) 2872 cm) DRI-3272 charcoal 4639 91 EB Compound, Feat. 4: 3628 ( 8.3%) 3586 3636 (78.8%) 3264 Southernmost Satellite Room: 3530 (56.6%) 3338 3242 (16.6%) 3102 F-8 hearth, Stratum II (20- 3208 ( 2.2%) 3194 26cm) 3148 ( 1.1%) 3141

43 Sinai S-1 (cont'd) DRI-3273 charcoal 3290 78 Compound, Feat. 4: Main 1662 (65.9%) 1496 1752 (95.4%) 1412 Enclosure, "truncated pit 1472 ( 2.3%) 1464 feature" F-6 hearth (5-11cm) Sinai S-10 ('Kite Gd-11317 charcoal 4390 110 EBII-IBA Main room, Unit A/6 (15-20 3322 (16.0%) 3234 3370 (93.0%) 2860 Eddy and Wendorf Site') (Timnian II) cm) 3171 ( 1.5%) 3162 2808 ( 2.0%) 2756 1999: Fig. 9.13- 3116 (50.8%) 2902 2720 ( 0.5%) 2704 9.15 Gd-7948 charcoal 4530 60 EBII-IBA Main room, Unit B/15 (15cm) 3358 (16.3%) 3311 3494 ( 2.3%) 3467 (Timnian II) 3295 ( 2.7%) 3286 3374 (88.6%) 3079 3274 ( 2.9%) 3265 3071 ( 4.5%) 3024 3238 (46.2%) 3106 Gd-7953 charcoal 4420 80 EBII-IBA Main room, hearth, U-B/6 3320 ( 9.8%) 3272 3340 (95.4%) 2908 (Timnian II) (20-30 cm) 3266 ( 7.3%) 3236 3169 ( 1.0%) 3164 3112 (50.1%) 2921 South Sinai Gebel Gunna (25) SMU-659 charcoal 4056 72 EBII L. 11; rounded room with 2840 ( 7.0%) 2814 2876 (95.4%) 2462 Bar-Yosef et al numerous artifacts (lithics), 2676 (61.2%) 2476 1986: 132; 147- sherds, grinding stones, and 149 charcoal Gebel Gunna SMU-659 charcoal 4374 64 EBII Charcoal, from large hearth 3089 (14.4%) 3046 3329 (12.7%) 3216 Bar-Yosef et al (100) directly outside structure 3036 (53.8%) 2908 3182 ( 2.0%) 3158 1986: 132; 147- L.105 3124 (80.7%) 2888 149

Sheikh Muhsein HV-5296 charcoal 4710 50 EBII 3627 (17.1%) 3589 3634 (29.3%) 3552 Beit-Arieh 1977: 3528 (14.7%) 3498 3541 (21.7%) 3484 199 3454 (36.4%) 3377 3475 (44.4%) 3370 Sheikh 'Awad RT-1806 charcoal 4325 55 EBII L. 73/90, Basket 73 0.5m 3011 (68.2%) 2894 3263 ( 0.7%) 3247 Segal and Carmi (Acacia below surface; charcoal from 3100 (94.7%) 2872 1996: 94; raddiana) Acacia raddiana (*noted as Lipschitz 2003: L.37 in Segal and Carmi 263-264 1996:104; and in Lipschitz 2003) Samar (East) Pta-3627 charcoal 3940 60 Neolithic-IBA Charcoal; 'Hearth' - habitation 2562 ( 9.3%) 2534 2580 (92.5%) 2276 Holzer et al. 2010 unit; holemouth jars 2493 (58.9%) 2343 2254 ( 2.9%) 2210 (Neolithic-IBA) RT-2715 charcoal 3775 40 Neolithic-IBA Charcoal; 'Hearth' - habitation 2281 (17.2%) 2249 2338 ( 1.4%) 2322 unit; holemouth jars 2232 (51.0%) 2139 2310 (84.1%) 2118 (Neolithic-IBA) 2098 ( 9.9%) 2039 RT-2716 charcoal 4080 25 Neolithic-IBA Charcoal; Top of living floor- 2833 ( 9.7%) 2818 2852 (14.8%) 2812 habitation unit sealing kite 2660 ( 6.2%) 2649 2744 ( 2.0%) 2726 below 2635 (52.3%) 2574 2696 (72.8%) 2566 2522 ( 5.8%) 2497 Wadi Arabah Wadi Fidan 4 HD-16327 charcoal 4718 25 EBI Area D, L 4-9 [habitation] 3626 (21.0%) 3598 3632 (32.0%) 3562 Adams and Genz 3526 (13.4%) 3506 3536 (20.7%) 3496 1995: 19 3427 (33.8%) 3381 3458 (42.7%) 3376 HD-16380 charcoal 4702 37 EBI Area D, L. 4-14 [habitation] 3622 ( 9.3%) 3604 3632 (21.3%) 3562 3524 (15.1%) 3498 3536 (21.7%) 3485 3454 (43.8%) 3378 3474 (52.4%) 3371

44 Wadi Fidan 4 HD-13776 charcoal 4684 50 EBI Area A, L 50 [habitation] 3618 ( 3.0%) 3611 3631 (16.7%) 3564 3521 (15.7%) 3490 3536 (78.7%) 3364 3470 (49.5%) 3373 HD-16379 charcoal 4576 45 EBI Area A, L 5 [habitation] 3493 (10.3%) 3468 3500 (16.1%) 3434 3374 (28.7%) 3328 3380 (38.3%) 3264 3216 (15.4%) 3179 3242 (40.9%) 3102 3158 (13.9%) 3122 HD-16378 charcoal 4424 51 EBI Charcoal, Area A, L.22 3308 ( 1.8%) 3300 3334 (25.5%) 3212 [habitation] 3282 ( 1.5%) 3276 3191 ( 5.6%) 3152 3264 ( 7.2%) 3240 3136 (64.3%) 2916 3105 (57.7%) 2928 Barqa al-Hetiye HD13975 charcoal 4376 57 EBII-III House 1, 27/91 L 13 3088 (12.2%) 3056 3326 ( 9.8%) 3231 Fritz 1994 [habitation] 3031 (56.0%) 2911 3224 ( 0.3%) 3220 3174 ( 1.0%) 3160 3119 (84.4%) 2891 HD13976 charcoal 4267 43 EBII-III House 1, 8/90, L. 3 2922 (62.5%) 2871 3013 (77.9%) 2858 Fritz 1994 [habitation] 2800 ( 5.7%) 2780 2810 (14.6%) 2752 2722 ( 2.9%) 2701 Feinan 9 HD-10577 4140 109 EBII-III smelting site 2876 (64.7%) 2618 3011 ( 1.6%) 2949 Hauptmann 1989 2607 ( 1.8%) 2599 2944 (93.8%) 2458 2594 ( 1.7%) 2586 HD-10993 3981 50 EBII-III furnaces nrs. 7 and 8 2576 (68.2%) 2460 2828 ( 0.2%) 2824 Hauptmann 2007: 2626 (94.4%) 2338 88, Table 5.1 2322 ( 0.8%) 2309 HD-10994 3973 85 EBII-III furnace nr. 25 2618 ( 1.7%) 2608 2858 ( 3.9%) 2810 Hauptmann 2007: 2598 ( 0.5%) 2594 2751 ( 1.6%) 2722 88, Table 5.1 2584 (66.0%) 2340 2700 (86.5%) 2268 2260 ( 3.4%) 2206 HD-10584 3812 77 EBII-III furnace nr. 24 2432 ( 2.0%) 2423 2470 (87.2%) 2111 Hauptmann 2007: 2402 ( 4.9%) 2380 2104 ( 8.2%) 2036 88, Table 5.1 2348 (61.3%) 2140 Wadi Ghuwair 4 HD-10573 charcoal 4059 55 EBII-III Copper smelting site; 2835 ( 6.1%) 2816 2864 (12.9%) 2806 Avner and Carmi charcoal 2667 (43.5%) 2548 2760 (82.5%) 2470 2001; Hauptman 2540 (18.5%) 2489 2007:89, Table 5.1 Ras en-Naqab 1 HD-10574 charcoal 3971 67 EBII-III flat slag heap, copper smelt; 2577 (53.9%) 2431 2836 ( 1.5%) 2815 Hauptmann 2007: charcoal; -.3m 2424 ( 5.8%) 2402 2672 (93.0%) 2282 88, Table 5.1 2381 ( 8.5%) 2348 2249 ( 0.8%) 2232 2218 ( 0.1%) 2214 Wadi Ghuwair 3 HD-16529 charcoal 3919 26 EBII-III Copper smelt; charcoal; slag 2470 (28.1%) 2432 2476 (91.5%) 2334 Hauptmann 2007: heap, -0.5m 2424 (16.0%) 2402 2324 ( 3.9%) 2304 89, Table 5.1 2381 (24.0%) 2348 Feinan 16 HD-10579 charcoal 3923 61 EBII-III slag heap, -0.3m 2484 (60.6%) 2333 2572 (90.2%) 2274 Hauptmann 2007: 2325 ( 7.6%) 2300 2256 ( 5.2%) 2208 88, Table 5.1 Khirbet Hamra HD-16533 charcoal 4044 40 EBIII Trench 1, L.114 2620 (38.2%) 2550 2848 ( 6.1%) 2812 Adams 2000 Ifdan 2538 (30.0%) 2490 2692 ( 0.2%) 2690 2679 (89.1%) 2470 Beta-143811 charcoal 4020 70 L.1236 2834 ( 3.3%) 2818 2865 ( 8.1%) 2804 Levy et al. 2002 2662 ( 2.6%) 2648 2761 (87.3%) 2342 2636 (62.3%) 2464

45 Khirbet Hamra Beta-143810 charcoal 3970 40 L.1010-14758 2569 (36.5%) 2516 2578 (95.4%) 2346 Levy et al. 2002 Ifdan 2500 (31.7%) 2460 Beta-143813 charcoal 3960 50 L.1602 2569 (25.8%) 2516 2580 (95.4%) 2295 Levy et al. 2002 2500 (27.5%) 2450 2420 ( 5.2%) 2405 2378 ( 9.6%) 2350 AA-68210 charcoal 3949 52 2565 (14.0%) 2526 2579 (95.4%) 2289 Ben-Yosef et al. 2496 (40.9%) 2400 2016: Table 3 2382 (13.3%) 2348 AA-68206 charcoal 3934 37 2482 (47.9%) 2398 2565 ( 7.5%) 2532 Ben-Yosef et al. 2384 (20.3%) 2346 2496 (87.9%) 2297 2016: Table 3 AA-68207 charcoal 3932 42 2481 (68.2%) 2346 2566 ( 8.9%) 2524 Ben-Yosef et al. 2497 (86.5%) 2294 2016: Table 3 HD-16534 charcoal 3914 45 IBA Trench 2, L.209 2471 (68.2%) 2342 2564 ( 3.5%) 2532 Adams 2000 2495 (89.9%) 2280 2250 ( 1.6%) 2232 2218 ( 0.3%) 2214 AA-68212 charcoal 3850 47 2451 (10.6%) 2420 2466 (94.5%) 2198 Ben-Yosef et al. 2405 (10.1%) 2378 2160 ( 0.9%) 2153 2016: Table 3 2350 (31.9%) 2276 2254 (15.6%) 2210 Beta-143812 charcoal 3650 60 L.1010-14208 2132 (19.5%) 2082 2200 (95.4%) 1884 Levy et al. 2002 2059 (48.7%) 1942 Ein Yahav RTT-4683 charcoal 3615 40 IBA L. 107; small semi-circular 2028 (68.2%) 1926 2131 ( 8.1%) 2086 Yekutieli et al. pit filled with slag, ash and 2050 (87.3%) 1884 2005 sediment Timna

Timna 30 Ham-215 charcoal 4020 100 Copper smelt; charcoal 2856 ( 7.5%) 2812 2876 (95.4%) 2292 Avner and Carmi 2747 ( 3.2%) 2724 2001 2698 (56.2%) 2456 2417 ( 1.2%) 2408 Timna S28 BONN-2363 charcoal 4000 90 EBII (?) Site 212, mine S28; EB II (?); 2834 ( 2.7%) 2818 2871 ( 7.7%) 2801 Weinstein 1984 charcoal 2662 ( 2.5%) 2646 2780 (87.1%) 2284 2637 (53.6%) 2430 2247 ( 0.6%) 2234 2424 ( 3.9%) 2401 2382 ( 5.5%) 2348 BONN-2362 charcoal 3890 70 EBII (?) Site 212, mine S28; EB II (?); 2470 (66.7%) 2286 2568 ( 5.1%) 2517 Weinstein 1984 charcoal 2246 ( 0.7%) 2244 2500 (88.3%) 2196 2238 ( 0.7%) 2236 2172 ( 2.0%) 2146

46 For this dissertation I looked at the following questions:

1) How do central and small sites fit within the new chronology of the IBA? Is there a

continual use of the two types of sites throughout the IBA? Or are there distinct periods of

use and/or reuse?

2) Is there a chronological division between central and small sites? If so, does this represent

changes in subsistence practices in the Negev Highlands? How does this correlate with data

from Egypt?

3) Are the fire features within structures and those in open areas contemporary? If not, do

these reflect multiple phases at the settlement? Additionally, do these dates provide

evidence for the archaeologically ‘invisible’ nomadic pastoral activity phenomena in other

periods (cf. Rosen et al. 2005)?

4) Are sites built and occupied at one time? Or is there ‘horizontal stratigraphy’, where sites

expand over time? 4 Methodology

4.1 Geo-ethnoarchaeology and reconstructing ancient subsistence practices

The approach of this research stems from the development of geo-ethnoarchaeology, which combines features of archaeological formation theory, ethnography and geosciences to

‘simultaneously study cultural and non-cultural formation processes’ (Shahack-Gross 2017:

38). Particularly, this work bases its interpretations on decades of ethno- and geoethno- archaeological work that has shown the importance of animal dung to archaeological reconstructions. The analysis of the organic and inorganic constituents of dung has been used to inform various aspects of ancient animal and human activities, including chemical biomarkers for species identification (e.g., Shillito et al. 2011; Prost et al. 2017), reconstruction of fuel sources and use (e.g., Miller 1984; Sillar 2000; Portillo et al. 2012; Gur-Arieh et al.

47 2013, 2014), construction material (Gur-Arieh et al. 2018), and—most importantly for this

study—archaeologies of space, activity and subsistence economies (e.g., Brochier et al. 1992;

Reddy 1999; Shahack-Gross et al. 2003, 2014; Valamotti and Charles 2005; Portillo et al. 2014;

Polo-Diaz et al. 2016; Dunseth et al. 2016, 2018, 2019; Smith et al. 2018; Baetan et al. 2018).

This work specifically focuses on the durable inorganic remains of animal dung—namely,

calcitic dung spherulites and opaline phytoliths—which have been found to be effective

indicators of herding, grazing and foddering strategies (Shahack-Gross 2011 and references therein). When identified together, ancient animal diet can be reconstructed. In turn, ethnoarchaeological work has shown that human modes of subsistence can be inferred from domestic animal dung as proxy, as animals raised by ‘pure’ pastoralists produce dung showing foddering with wild plants while agropastoralist herds show evidence of wild plants and

agricultural byproducts (Shahack-Gross et al. 2003; Valamotti and Charles 2005; Albert et al.

2008; Portillo et al. 2014; Elliot et al. 2015; Polo-Diaz et al. 2016).

4.1.1 Dung spherulites

Dung spherulites are microscopic (5–25 μm) dumbbell to spheroid calcitic microremains made

of radially-oriented acicular crystallites (Canti 1997; Canti and Brochier 2017). They are

known to form in the digestive system of a variety of animal species, most abundantly in

ruminants (e.g., Canti 1997, 1998, 1999; Goren 1999; Korstanje 2005; Shahack-Gross and

Finkelstein 2008; Portillo et al. 2014), but both their formation mechanism and exact

composition are still unclear. It has been shown empirically that they form in the middle of the

small intestine in sheep (Canti 1999: 252–253). There has been some suggestion that dung

2+ − spherulites precipitate directly from free calcium (Ca ) and bicarbonate (HCO3 ) ions in the

increasingly alkaline lower intestine (Canti 1999), or as bacterially-mediated

monohydrocalcite (CaCO3•H2O, Shahack-Gross 2011: 208; cf. Rodriguez-Navarro et al. 2007;

48 Zhang et al. 2017). In archaeological deposits, dung spherulites are composed of calcite

(Dunseth and Shahack-Gross 2018).

As calcitic microremains that possess large surface areas relative to volume, they are prone to dissolution in acidic or neutral conditions (Canti 1997; Gur-Arieh et al. 2014). In archaeological contexts they are mainly found at cave sites, rock shelters and sites in arid or semi-arid environments (e.g., Brochier et al. 1992; Matthews et al. 1997; Karkanas 2006;

Portillo et al. 2014; Polo-Diaz et al. 2016; Dunseth et al. 2016, 2018). In the scope of

microarchaeological investigations in the Negev, they are generally well-preserved in the

alkaline calcareous sediments common to the region (Shahack-Gross and Finkelstein 2008;

Shahack-Gross et al., 2014; Dunseth et al. 2016, 2018).

4.1.2 Phytoliths

Phytoliths are silica biominerals (SiO2•nH2O) that infill and take the shape of cells, and thus their morphologies have taxonomic value for identifying archaeological plant material (cf.

Piperno 2006; Ball et al. 2016). Decades of study on modern plants has shown that different species—and within species, specific plant organs—produce different morphotypes and quantities of phytoliths (e.g., Twiss et al. 1969; Piperno 2006; Rapp and Mulholland 1992;

Albert and Weiner 2001; Tsartsidou et al. 2007). In general, annual monocotyledonous plants

(e.g., grasses) produce significantly more phytoliths than herbaceous or woody dicotyledonous plants (by orders of magnitude, Albert and Weiner 2001) and are characterized by different morphologies. These aspects allow us to utilize archaeological different phytolith assemblages and to reconstruct their botanical origin(s).

Specifically this study focused on phytolith concentrations and dendritic morphotypes, which form in the inflorescence of many monocotyledonous grasses. Dendritic phytoliths are especially abundant in cultivated cereal species (e.g., wheat and barley), and can be used as

49

proxy for foddering with agricultural byproducts. Research on modern plants showed that wild grasses contain less than 7–8% dendritic morphotypes, while cultivated cereal species show

more than 8% (Albert et al. 2008). Geo-ethnoarchaeological data of pre-modern Bedouin sites

in the Negev showed that animal dung from free-grazing herds in desert areas is typified by

very low phytolith concentrations and less than 1% presence of dendritic morphotypes, as

animals primarily feed on phytolith-poor local shrub vegetation (Shahack-Gross and

Finkelstein 2008). Animal dung related to agro-pastoral activities in the Negev is characterized

by higher phytolith concentrations associated with at least 3–4% of dendritic morphotypes, as

animals feed on both phytolith-poor local shrub vegetation and the straw and chaff byproducts

of domestic cereals (Shahack-Gross et al. 2014).

Using this approach on archaeological sites, previous work dealt with longstanding debate over

Iron Age activity in the Negev, specifically in regard to the Negev ‘fortresses’, their perceived destruction and their proximity to agricultural terraces (e.g., Cohen 1979; Haiman 1994). This research showed that Iron IIA livestock herds at two sites in the Negev Highlands were free- grazing, with no evidence for foddering with agricultural byproducts (Shahack-Gross and

Finkelstein 2008; Shahack-Gross et al. 2014). In contrast, later Byzantine-Early Islamic sites consistently showed evidence for herds foddered on both wild vegetation and supplemented with cereal byproducts (Shahack-Gross et al. 2014; Dunseth et al. 2019), which corroborated the textual record (Kraemer 1958; Mayerson 1962). Coupled with radiocarbon determinations

from the studied sites and OSL ages from agricultural terraces by other researchers, this showed

that (other than possibly one exception) agriculture on the large scale in the Negev did not

predate the Roman period (Avni et al. 2012, 2013, 2019; Shahack-Gross and Finkelstein 2008,

2015, 2017).

A recent methodological study led by the author recently compared three archaeobotanical proxies (phytoliths, pollen and seeds) in modern and archaeological dung pellets from the

50

Negev Highlands (Dunseth et al. 2019). This research indicated that while phytoliths have low

taxonomic resolution, as an inorganic substance they have the highest potential of preservation.

In addition, of the three archaeobotanical methods, only phytoliths consistently provided direct

evidence for foddering with agricultural byproducts.

4.1.3 Ash pseudomorphs

Ash pseudomorphs are calcitic pseudomorphs after biogenic calcium oxalate monohydrate

(whewellite, CaC2O4•H2O) or dihydrate (weddellite, CaC2O4•2H2O) crystals that form in the

wood and bark of many dicotyledonous plants, as well as in a few monocotyledonous species

(Arnot and Pautard 1970; Wattez and Courty 1987; Brochier and Thinon 2003; Shahack-Gross

and Ayalon 2013). Thermal decomposition of these calcium oxalate crystals into calcite occur

at temperatures between approximately 400–700 °C (Kociba and Gallagher 1996), usually during the burning of organic matter. Common forms include rhombs (frequently twinned), druses or needle-like raphides (Canti 2003).

Ash pseudomorphs tend to preserve better than dung spherulites, due to their smaller surface area to volume ratios (Gur-Arieh et al. 2014). Like dung spherulites, they are more likely to preserve in sheltered cave sites (e.g., Schiegl et al. 1996; Karkanas et al. 2007), although they have been identified in abundance in fire features at open-air Mediterranean tell sites (Gur-

Arieh et al. 2014), buried open-air Middle Paleolithic sites (e.g. Friesem et al. 2014b) and at arid open-air sites in the Negev (Shahack-Gross and Finkelstein 2008; Shahack-Gross et al.

2014; Dunseth and Shahack-Gross 2018; Dunseth et al. 2016, 2018, 2019). The presence of such ash pseudomorphs are direct indicators of archaeological ash, although other analyses

(such as FTIR and/or micromorphology) are needed to identify in situ burning or secondary deposition. Importantly for this study, they facilitate the differentiation of sediments composed of ash, dung or mixtures of the two, so that associated archaeobotanical assemblages (in this

51

dissertation, phytoliths) can be analyzed and interpreted accordingly (see also Dunseth et al.

2019).

5 Materials and methods

5.1 Description of sites, their microenvironments and previous excavations

For this study two central and two small sites from different microenvironments transecting

the Negev Highlands were selected to control for inter-regional variation, including decreasing

precipitation from north to south (and west to east, in the case of Ein Ziq), variable proximity

to perennial water sources, and differing geomorphological landscapes (more information

below in the respective articles). Except for Nahal Nizzana 332, all sites were previously

excavated in traditional methods.

5.1.1 Mashabe Sade

The central site of Mashabe Sade (NIG: 179800 543100, 428 m a.s.l.), the largest IBA site in

the Negev Highlands, is located on a c. 250 m long limestone ridge of the Turonian Shivta

Formation (Starinsky et al. 2010). Steep slopes and narrow wadis rich in flora border the site

to the north, east and west. Annual precipitation in the area is 100–120 mm/year (Israel

Meteorological Service, IMS). The site is approximately 5 km from the nearest modern water sources in the area (Bor Mashabbim [Arabic: Bir ‘Asluj] and Bor Sade). Ancient agricultural terraces are limited in the region, although there are some in the wadis northeast of the site.

Mashabe Sade is located on a narrow ridge that provides a commanding view of the surrounding hillscape and the wadis to the north and west. This might suggest some defensive considerations to its location.

The site was surveyed and excavated by Rudolf Cohen (Cohen 1999: 117–130; Fig. 3 below) in the 1980s. Approximately 25 structures of the 200+ structures were excavated, revealing a

52

single period of IBA occupation based on ceramic typology. One multiroom rectangular structure (L. 31, 34, 35) was exposed, unique in the Negev Highlands. Excavations of another collapsed structure (L. 24, Cohen 1999: Fig. 74) informed how the typical IBA Negev buildings were constructed, with central drum pillars supporting thin limestone roofing slabs (Cohen

1992: Fig. 7).

A small collection of Family S ceramics (Dever 1973, 1980), including storage vessels and amphoriskoi, was found throughout the excavated structures, as well as grinding stones in many rooms. Notably, numerous copper artifacts were found during excavations, including a

fragment of a copper ingot, scrap, as well as finished copper awls and a dagger. No

macrobotanical remains were noted in the final report and only a few ovicaprine faunal remains

were collected during the excavations (Hakker-Orion 1999).

53 0 20 N

Figure 3: Plan of Mashabe Sade, redrawn after Cohen 1999: Fig. 71. Surveyed structures indicated as plain ovals; excavated structures detailed. Scale is in meters.

54

5.1.2 Ein Ziq

The central site of Ein Ziq (NIG: 186170 523900, 325 m a.s.l.); is located on the lower terrace

of a Pleistocene alluvial fan (terrace conglomerate) of Nahal Ziq (Avni and Weiler 2013).

Annual precipitation is approximately 80–100 mm/year (IMS). Two natural springs, Ein Ziq

and Ein Shaviv, are in the immediate vicinity, c. 1 km southwest of the site. The site is located

on an ancient east-west trading route known from later periods between Transjordan and Sinai

(Haiman 1996).

Ein Ziq (Figs. 4AB) is the second largest (c. 2 ha) IBA site in the Negev Highlands, spreading

over a pair of alluvial terraces above the dry wadi bed of Nahal Ziq. The site was extensively

excavated (~100 rooms/structures out of over 200 visible on the surface) in the early 1980s by

Cohen (1999: 137–188). Structures at the site (c. 2–4 m in diameter) are all stone-built,

preserved in some areas up to 1 m and dug c. 15–20 cm into the alluvial surface. Ceramics

found in the excavation showed one main period of activity, during the IBA, with some reuse

of structures as Nabatean tombs (e.g., Cohen 1999: 142), and a lime kiln radiocarbon-dated to

the Early Islamic period (Segal and Carmi 1996: 95).

Faunal remains recovered in Cohen's excavations, though limited (n = 242 number of

individual specimens), showed a preference for meat-bearing parts of younger ovicaprines (c.

88%), as well as a smaller component (c. 10%) of wild desert animals including birds, hares

and gazelle (Hakker-Orion, 1999: 329–330). Macroscopic botanical remains from selected loci

were reported including charred and uncharred wood species, all local to the immediate

environment (Baruch 1999: 7*–11*). There was no evidence for economic plants (i.e., cultivated seeds, etc.). Phytolith analysis of five sediment samples, each yielding only around

30 phytoliths, was inconclusive (Rosen 1999).

55 A

B

New Excavations Surveyed (Cohen 1999) Excavated (Cohen 1999)

0 20 m

Figure 4: A) Aerial photo of Ein Ziq, taken in 2015 after new excavations. B) Ein Ziq site plan, with new areas excavated and sampled in red. Redrawn after Cohen 1999: Fig. 88.

56

The analysis of lithic material from Ein Ziq was limited by older collection methodologies

(Vardi 2005; Vardi et al., 2007: 103). Nevertheless, Vardi et al. (2007: 112–114) concluded that: 1) the lithic assemblage suggests ad hoc domestic activity, including production and utilization of a variety of materials (organic, leather, bone, shell, etc.), and 2) the five ‘sickle blades’ found at the site are insufficient evidence for agricultural activity. Large quantities of hammerstones and grinding stones (photographed though unpublished) led Haiman (1996) to suggest large-scale copper processing activities , although Vardi et al. (2007: 113) questioned this identification due to the lack of other copper production items (e.g., crucibles, molds, etc.).

5.1.3 Nahal Boqer 66

Nahal Boqer 66 is a small site (0.2 ha) c. 4 km north of modern Sede Boqer (NIG: 179900

0535400, 521 m a.s.l.). Annual rainfall in the area is 80–100 mm/year (IMS). The site is located in a small saddle between two low Turonian limestone ridges of the Nezer Formation (Avni and Weiler 2013). It is composed of two stone-built complexes typical to the Negev Highlands

(see Finkelstein 1995a: 37–49; Haiman 1996), each with oval or rectilinear rooms (c. 3–5m in diameter) attached to large courtyards (8–14 m in diameter). Most walls are preserved only to

1–2 courses (c. 50–60 cm) and are made of limestone blocks from the immediate vicinity.

The site was excavated in the 1970s by R. Cohen (1985: 40–41, 1999: 60–62). Lithic evidence in the immediate vicinity suggests earlier Pre-Pottery Neolithic B (PPNB) activity (Noy and

Cohen 1974), while the structures were dated by ceramic typology to the EB II and IBA (Cohen

1985: 40; Cohen 1999: 61–62). The distribution of pottery (and to a lesser extent, minor architectural changes) identified in the earlier excavation showed two distinct occupations, with the excavated portions of the Northern Complex dominated by IBA sherds, and the

Southern Complex representing a mix of EB and IBA sherds (Cohen, 1999: Fig. 41). No faunal or botanical remains were reported from the earlier excavations.

57 A

B

Southern Complex Northern Complex

0 4 8 12 16 m

Figure 5: A) Aerial photo of Nahal Boqer 66 taken in 2018 (courtesy Omer Ze’evi). B) Nahal Boqer 66 site plan redrawn after Cohen 1999: Fig. 41. New excavation areas shown in red.

58 5.1.4 Nahal Nizzana 332

Nahal Nizzana 332 (NIG: 171900 500140, 855 m a.s.l.) is comprised of a series of stone-built

complexes and single structures built over c. 5 hectares on a Cenomanian dolomitic-marl spur bordered by branches of Nahal Nizzana to the north and west (Zilberman and Avni 2004).

Annual precipitation in the area is c. 70–80 mm (IMS). Most of the structures are preserved only to a single course and built directly on bedrock.

Figure 6: Georectified orthophoto of Nahal Nizzana 332.1, produced using Agisoft Photoscan 1.42.

59

No previous excavations were conducted at the site. Surface survey of the area identified

diagnostic pottery dated to the Early Bronze, Intermediate Bronze (majority of sherds) and Iron

Age IIA (Haiman 1991: 127–128). No pottery was published, and independent complexes were

not associated with specific periods. Immediately to the east is Bor Hemet, a massive (c. 18 x

16 m) stone-lined reservoir, one of many cisterns in the area generally attributed to the Iron

Age (e.g., Evenari et al. 1958; Haiman 1994; although see new data presented in Junge et al.

2018).

5.2 Analysis

A battery of microarchaeological analyses including Fourier transform infrared spectroscopy,

X-ray florescence, optical microscopy (analysis of dung spherulite, phytolith, and ash pseudomorph morphologies and quantifications) were utilized in this research to characterize archaeological sediments and reconstruct subsistence strategies, activity areas, and site formation following informed by geoarchaeological and geo-ethnoarchaeological studies of the last few decades (see Section 4.1, above). These analyses were carried out by the author at the Kimmel Center for Archaeological Science, Weizmann Institute of Science (2014–2015) and at the Laboratory of Sedimentary Archaeology, University of Haifa (2015–2018).

5.2.1 Fourier transform infrared (FTIR) spectroscopy

FTIR analysis of all samples reported in this dissertation were conducted to determine the mineralogical composition of sediments and to evaluate if sediments had been exposed to heat

(cf. Berna et al. 2007; Forget et al. 2015; experimental data). Samples were prepared following the conventional potassium bromide (KBr) method: approximately 0.3 mg of sediment were powdered and homogenized using an agate mortar and pestle and mixed with c. 0.4 mg of KBr

(Weiner et al. 1993). The sample was then pressed into a pellet using Specac 2-ton hand press.

Each pellet was analyzed between 4000 and 250/400 cm-1 at a 4 cm-1 resolution using either a

60

Thermo Scientific Nicolet 380 (at the Weizmann Institute of Science) or a Nicolet iS5 spectrometer (at the University of Haifa) with Omnic 9.3 software. Spectra were compared to an extensive reference library (courtesy of the Kimmel Center for Archaeological Science,

Weizmann Institute of Science), published and experimental data (Berna et al. 2007; Regev et al. 2010; Weiner 2010; Friesem et al. 2014a; Forget et al. 2015; Dunseth and Shahack-Gross

2018).

5.2.2 X-Ray fluorescence (XRF)

Copper processing activities were evaluated in archaeological contexts (e.g., ash, building collapse, floors) and control sediments at the central sites of Mashabe Sade (n = 5) and Ein Ziq

(n = 63). Elemental composition was determined using a Spectro-XEPOS Energy Dispersive

X-Ray Fluorescence unit (ED-XRF) at the Kimmel Center for Archaeological Science,

Weizmann Institute of Science (courtesy S. Weiner). Samples were manually ground and

homogenized to a particle size of < 425 μm through sieving, following analytical procedures

for powder samples outlined by Eliyahu-Behar et al. (2012: 258). A known analytical soil

standard (GSS-1; Xie et al. 1985, 1989) was run with each tray of 11 samples to calculate daily

measurement error for each element.

5.2.3 Microremains: phytoliths, dung spherulites and ash pseudomorphs

Phytoliths were extracted from the ashed dung pellet samples following the rapid extraction

method of Katz et al. (2010). Phytoliths were quantified by systematically counting specimens

across 16 random fields of view at 200× using a Nikon Eclipse 50i POL petrographic microscope. All individual phytoliths in anatomical connection (i.e., ‘multicells’) were counted to inhibit bias; percentage of anatomically connected phytoliths is reported and used as a proxy for preservation (Cabanes et al. 2010). Phytolith morphologies were identified following standard literature (Twiss et al. 1969; Rapp and Mullholland 1992; Albert and Weiner 2001;

61

Piperno 2006) and a reference collection of Negev plants at the Laboratory for Sedimentary

Archaeology (University of Haifa). Morphologies were determined at 400× under plane-

polarized light (PPL) by identifying at least 200 phytoliths with consistent morphologies

(Albert and Weiner 2001), and greater than 250 if possible (Zurro 2018). All morphotype

descriptions follow the International Code for Phytolith Nomenclature (ICPN) when possible

(Madella et al. 2005).

Calcitic dung spherulites and ash pseudomorphs were extracted and dispersed using the sodium

polytungstate (SPT) procedure outlined in Gur-Arieh et al. (2013) and discussed in many of

the articles below. All calcitic microremains were counted at 400× in 16 random fields of view

using the same microscope stated above, utilizing both PPL and cross-polarized light (XPL).

All microremain concentrations are reported in millions per 1 gram of sediment. For the Negev samples counted in duplicate (n = 24) a coefficient of variance error is variable according to

site and sample (Table 2).

It is notable that samples with very low concentrations of microremains are skewed towards

higher coefficient of variance, while samples with concentrations at orders of magnitudes

higher are as low as 6–7% (Table 2, Fig. 7). Samples with concentrations above 0.5 million per

1 g of sediment average approximately 44% for phytoliths (n = 5), 32% for dung spherulites

(n = 6) and 22% for ash pseudomorphs (n = 1). This is similar to the error of c. 20–30%

generally reported for microremain studies (Albert and Weiner 2001; Katz et al. 2010; Gur-

Arieh et al. 2013), as well as other studies reported by the author (e.g., 15–23% reported for

samples with higher microremain concentrations in Dunseth et al. 2019: 172).

62 Phytoliths Dung Spherulites Ash Pseudomorphs Site Sample n Average S.D. (±) COV Average S.D. (±) COV Average S.D. (±) COV Ein Ziq EZ-084 2 0.07 0.05 71% 0.04 0.06 150% 0.12 0.06 50% Ein Ziq EZ-105 2 0.06 0.02 33% n.d. n.d. n.d. n.d. n.d. n.d. Ein Ziq EZ-106 2 0.01 0.01 100% n.d. n.d. n.d. n.d. n.d. n.d. Ein Ziq EZ-107 2 0.02 0 - n.d. n.d. n.d. n.d. n.d. n.d. Ein Ziq EZ-115 2 0.02 0.02 100% n.d. n.d. n.d. n.d. n.d. n.d. Ein Ziq EZ-330 2 n.d. n.d. n.d. 0 0 - 0 0 - Ein Ziq EZ-343 2 0.27 0.11 41% n.d. n.d. n.d. n.d. n.d. n.d. Ein Ziq EZ-345 2 0.52 0.38 73% 0 0 - 0.69 0.15 22% Ein Ziq EZ-349 2 0.56 0.26 46% 0.1 0.14 140% 0.21 0.18 86% Ein Ziq EZ-C10 2 n.d. n.d. n.d. 0 0 - 0 0 - Ein Ziq EZ-C11 2 0.15 0.21 140% 0.02 0.03 150% 0 0 - Ein Ziq EZ-C14 2 0 0 - 0.05 0.02 40% 0 0 - Mashabe Sade (IBA) ZMS-12.41 2 0.2 0.3 150% 0.05 0.07 140% 0 0 - Mashabe Sade (Iron Age) ZMS-12.69 2 1.9 0.4 21% 1.7 1.4 82% 0 0 - Nahal Boqer 66 NB-02.02 2 3.9 1.53 39% n.d. n.d. n.d. n.d. n.d. n.d. Nahal Boqer 66 NB-03.03 2 n.d. n.d. n.d. 163 45 28% 0.3 0.4 133% Nahal Boqer 66 NB-09.02 2 n.d. n.d. n.d. 43 15 35% 0 0 - Nahal Boqer 66 NB-12.04 2 2.76 1.15 42% n.d. n.d. n.d. n.d. n.d. n.d. Nahal Nizzana 332 N11.2 2 n.d. n.d. n.d. 35 2 6% 0.06 0.03 50% Nahal Nizzana 332 N12.2 2 n.d. n.d. n.d. 14 1 7% 0 0 - Nahal Nizzana 332 N14.2 2 n.d. n.d. n.d. 43 20 47% 0.03 0.04 133% Nahal Nizzana 332 N17.2 2 n.d. n.d. n.d. 0.5 0.1 20% 0 0 - Nahal Nizzana 332 N18.2 2 n.d. n.d. n.d. 0.1 0.1 100% 0.03 0.05 167% Nahal Nizzana 332 N19.2 2 n.d. n.d. n.d. 0 0 - 0 0 -

Table 2: Coefficient of variation data for all samples counted in duplicate. Note high error for samples with low concentrations of microremains.

63 A 160% Phytoliths 140%

120%

100%

80% y = -0.113ln(x) + 0.5533 60% R = 0.2528

40%

20%

0% 0 1 2 3 4 5 6

B 160% Dung Spherulites 140%

120%

100%

80%

60%

40%

20% y = -0.133ln(x) + 0.7357 R = 0.5897 0% 0 50 100 150 200 250

C 180% 160% Ash Pseudomorphs

140%

120%

100%

80% y = -0.258ln(x) + 0.3587 60% R = 0.3254

40%

20%

0% 0 0.2 0.4 0.6 0.8 1

Figure 7: Coefficient of variation (CoV) of samples analyzed in duplicate to microremain concentration. All concentrations reported in millions per g of sediment.A) Phytolith concentrations to CoV. B) Dung spherulite concentrations to CoV. C) Ash pseudomorphs to CoV. Note trendlines leveling off around 30-40% in the phytolith and ash pseudomorph plots, approximately the same error as reported in Albert and Weiner 2001; Katz et al. 2010 and Gur-Arieh et al. 2013. The samples exceptionally rich in dung spherulites (orders of magnitude greater than the other microremains) show lower error.

64

5.2.4 Spatial analysis and statistics

For Nahal Nizzana 332, a systematic sampling strategy was adopted to explore spatial

patterning of microremains (above). Spatial analysis and patterns were analyzed using a

combination of QGIS 3.2 and ArcGIS 10.1. Spatial limits of subunits and sampling

neighborhoods were defined and drawn using the Avenza MAPublisher 9.9 plugin for Adobe

Illustrator CS6 from georectified orthophotos created using Agisoft Photoscan 1.4.2. Due to

varying sizes of excavated subunits (as limited by architecture), artifact concentrations were normalized to the exact area of sediment excavated (in final publication this will be by volume, as approximated by elevation values from the corners and center of each subunit).

Lithics are presented as counts per square meter. Due to the collection strategy, distribution of certain macroscopic artifacts is shown as points (for artifacts identified in field) and chloropleth maps.

Microremain concentrations were logarithmically normalized with an added constant of 2

using PAST 3.2 (Hammer et al. 2001). Inverse distance weighting (IDW) was used to

interpolate the microremain data over the excavated and unexcavated space (cf. Rondelli et al.

2014; Butler et al. 2018). IDW interpolation is appropriate for smaller datasets (n < 100) and

for local spatial models without prior knowledge of the spatial structure or stationarity (Butler

et al. 2018: 980 and references therein). The dataset was separated into smaller local sampling

neighborhoods as delimited by architecture. Cross-validation and the root mean square

prediction error (RMSE) were calculated in ArcGIS to establish the power coefficient and the

resulting interpolations (Liao et al. 2011; Chen and Liu 2012). Additional statistics were

calculated using IBM SPSS 24.

65

6 General description of the articles and their contribution

This dissertation includes the following four published articles:

6.1 Methodological contribution

Dunseth, Z.C. and Shahack-Gross, R. 2018. Calcitic dung spherulites and the potential for rapid identification of degraded animal dung at archaeological sites using FTIR spectroscopy.

Journal of Archaeological Science 98: 118–124.

This article deals with a new method developed to identify spherulite-rich degraded animal dung at archaeological sites using infrared spectroscopy (FTIR). This method has the potential for a wide-ranging impact on identification of animal dung during routine lab or field FTIR analysis at arid, semi-arid and cave sites throughout the world. In addition, the method suggests that the mineralogical signature of ancient animal dung is resistant to diagenetic change in arid environments, and may become an efficient proxy in the future to assess the state of preservation of deposits used for isotopic studies, paleoclimate reconstructions and radiocarbon dating.

6.2 Chronology of IBA settlement in the Negev

Dunseth, Z.C., Junge, A., Lomax, J., Boaretto, E., Fuchs, M., Finkelstein, I. and Shahack-

Gross, R. 2017. Dating archaeological sites in an arid environment: A multi-method case study in the Negev Highlands, Israel. Journal of Arid Environments 144: 156–169.

This article reports on the chronology of the Intermediate Bronze Age (IBA) sites studied in this dissertation. Data from radiocarbon, OSL and micromorphology (to define context) from the central site of Ein Ziq and the small (ephemeral) site of Nahal Boqer 66 are presented and discussed. Most important is third millennium BCE continuity at Nahal Boqer (Early Bronze

Ib – beginning of the IBA), while Ein Ziq is limited to the beginning of the IBA. Both sites are

66

abandoned after the first half of the IBA, corresponding to the collapse of the Egyptian Old

Kingdom.

6.3 Subsistence economies and trading systems of IBA Negev settlement

Dunseth, Z.C., Finkelstein, I. and Shahack-Gross, R. 2018. Intermediate Bronze Age

subsistence practices in the Negev Highlands, Israel: Macro- and microarchaeological results

from the sites of Ein Ziq and Nahal Boqer 66. Journal of Archaeological Science: Reports 19:

712–726.

This paper presents and summarizes the macro- and microarchaeological data from our excavations at Ein Ziq and Nahal Boqer 66, with reference to Mashabe Sade (below), and

discusses the wider patterns of subsistence economies, trade, and settlement during the IBA in

the Negev Highlands. Most importantly, we identify two different subsistence economies

during the period, one based on animal husbandry at small sites, and one relying only on

interregional (copper) trade, and not copper and/or food production, at central sites. This paper presents a new paradigm for IBA settlement in the Negev Highlands.

Dunseth, Z.C., Junge, A., Fuchs, M., Finkelstein, I. and Shahack-Gross, R. 2016.

Geoarchaeological Investigations in the Intermediate Bronze Age site of Mashabe Sade, the

Negev Highlands. Tel Aviv 43: 43–75.

This article is included to support and parallel the results from the central site of Ein Ziq. It

includes data from two seasons of excavation in 2013 and 2014 and builds upon the preliminary results of Dunseth (2013). It also references OSL dating at the site that showed that the site

was abandoned at the end of the IBA. The micro- and macroarchaeological data at Mashabe

Sade exhibit the same patterns as those identified at Ein Ziq and supports the conclusions

regarding subsistence economies at central Negev sites during the IBA.

67

6.4 Preliminary results from Nizzana 332

One additional chapter includes unpublished data (radiocarbon dates, results of macro- and microarchaeological analyses) from a high-resolution investigation at Nahal Nizzana 332, excavated in 2017 as a parallel to the published small site of Nahal Boqer 66. The data presented here exhibits the same pattern as Nahal Boqer 66, as well as validates the small-scale sampling methods utilized in this dissertation.

7. Articles and preliminary results from Nahal Nizzana

The published articles are included in the following pages.

68

7.1 Methodological contribution (Dunseth and Shahack-Gross 2018)

69 Journal of Archaeological Science 97 (2018) 118–124

Contents lists available at ScienceDirect

Journal of Archaeological Science

journal homepage: www.elsevier.com/locate/jas

Calcitic dung spherulites and the potential for rapid identification of degraded animal dung at archaeological sites using FTIR spectroscopy T

∗ ∗∗ Zachary C. Dunsetha,b, , Ruth Shahack-Grossb, a Institute of Archaeology, Tel Aviv University, Tel Aviv 6997801, Israel b Laboratory for Sedimentary Archaeology, Department of Maritime Civilizations, Recanati Institute of Maritime Studies, Leon H. Charney School of Marine Sciences, University of Haifa, Haifa 3498838, Israel

ARTICLE INFO ABSTRACT

Keywords: Animal dung is increasingly a valuable resource for reconstructing ancient activity in archaeology. One of the Dung spherulites most common archaeological indicators of dung in caves and arid environments are calcitic dung spherulites that Grinding curve form in the digestive system of a variety of animals. Although many aspects of their formation and taphonomy FTIR are understood, details of their mineralogy remain poorly-defined. Using the Fourier transform infrared (FTIR) grinding curve method, we report here that archaeological sediments containing large amounts of dung spherulites can be differentiated from sediments composed of other forms of geogenic and pyrogenic calcites. We propose that this attribute can be used to rapidly identify well-preserved degraded dung deposits at archae- ological sites with routine laboratory or on-site field FTIR analysis. This observation at a 5000-year-old open air site suggests that the grinding curve method also has potential to be used for assessing preservation of dung spherulites for future radiocarbon or stable isotope investigations.

1. Introduction most abundantly in ruminants (e.g., Canti, 1999; Korstanje, 2005; Shahack-Gross and Finkelstein, 2008; Portillo et al., 2014), but both Three decades of ethno- and geo-ethnoarchaeological work has their formation mechanism and exact composition are still unclear. It shown the importance of animal dung to archaeological reconstruc- has been shown experimentally that they form in the middle of the tions. The analysis of the organic and inorganic constituents of dung has small intestine in sheep (Canti, 1999: 252–253). There has been some been used to inform various aspects of ancient animal and human ac- suggestion that these precipitate directly from free calcium (Ca2+) and − tivities, including identification of species (e.g., Shillito et al., 2011; bicarbonate (HCO3 ) ions in the increasingly alkaline lower intestine Prost et al., 2017), fuel use (e.g., Miller and Smart, 1984; Sillar, 2000; (Canti, 1999), or as bacterially-mediated monohydrocalcite (Ca-

Portillo et al., 2012; Gur-Arieh et al., 2013, 2014), and archaeology of CO3*H2O, Shahack-Gross, 2011: 208; cf. Rodriguez-Navarro et al., space, human subsistence and foddering strategies (e.g., Brochier et al., 2007; Zhang et al., 2017). In practice, sediments composed of large 1992; Reddy, 1999; Shahack-Gross et al., 2003, 2014; Valamoti and amounts of dung spherulites are rich in calcite (see more below), im- Charles, 2005; Portillo et al., 2014; Polo-Diaz et al., 2016; Dunseth plying that this is their composition in archaeological deposits. et al., 2016, 2018). As calcareous microremains that possess large surface areas relative A variety of geoarchaeological methods are currently utilized to to bulk volume, they are prone to dissolution (cf. Gur-Arieh et al., identify dung at archaeological sites (see Shahack-Gross, 2011 and re- 2014). They are mainly found at cave sites, rock shelters and sites in ferences therein). One of the most established is the identification and arid or semi-arid environments (e.g., Brochier et al., 1992; Matthews quantification of inorganic calcareous dung spherulites (and phytoliths et al., 1996; Karkanas, 2006; Shahack-Gross et al., 2014; Portillo et al., in tandem) using optical microscopy (Shahack-Gross, 2011; Lancelotti 2014; Polo-Diaz et al., 2016; Dunseth et al., 2016, 2018). Because of and Madella, 2012). their susceptibility to dissolution, it has so far been difficult to separate Dung spherulites are microscopic (5–25 μm) roughly-spherical par- them from associated organic and/or mineral components by estab- ticles made of radially-oriented acicular crystallites (Canti, 1997). They lished extraction procedures (Canti, 1997, 1998; Canti and Nicosia, are known to form in the digestive system of a variety of animal species, 2018; Shahack-Gross personal observations). Therefore, the infrared

∗ Corresponding author. Institute of Archaeology, Tel Aviv University, Tel Aviv 6997801, Israel. ∗∗ Corresponding author. E-mail addresses: [email protected] (Z.C. Dunseth), [email protected] (R. Shahack-Gross). https://doi.org/10.1016/j.jas.2018.07.005 Received 27 April 2018; Received in revised form 29 June 2018; Accepted 13 July 2018 0305-4403/ © 2018 Elsevier Ltd. All rights reserved. Z.C. Dunseth, R. Shahack-Gross Journal of Archaeological Science 97 (2018) 118–124 signature of these archaeologically important calcitic micro-remains is show: one, sediments dominated by dung spherulites can be differ- relatively little explored. entiated from other calcite-containing sediments, and two, sediments Fourier Transform Infrared (FTIR) spectroscopy has been used to that plot closer to geogenic calcite as well as wood ash grinding curves study an extensive suite of archaeological materials (see recent review may be the result of post-depositional mixing between dung and geo- in Monnier, 2018). Calcite (the mineral polymorph of archaeologically- genic deposits. In both cases, heat has no effect on where dung-derived preserved dung spherulites) has three important characteristic vibra- sediments plot. Furthermore, we posit that this observation can be used −1 tional peaks in the mid-IR range: 712 cm (ν4, in-plane CO3 bend), to rapidly identify the presence of degraded dung in archaeological −1 −1 874 cm (ν2, out-of-plane CO3 bend) and 1420 cm (ν3, asymmetric sediments. CO3 stretch) (White, 1974). Several archaeologically-relevant calcites, including geogenic (spar, limestone, flowstones, and chalk) and pyro- genic specimens (wood ash and lime plaster) have recently been studied 2. Materials and methods (Chu et al., 2008; Regev et al., 2010, 2011; Poduska et al., 2011, 2012; Xu et al., 2015, 2016; Goshen et al., 2017). Although particle size, 2.1. Archaeological and control samples distribution and porosity (Duyckaerts, 1959; Lane, 1999; Surovell and Stiner, 2001) and crystal morphologies (Ruppin and Englman, 1970; Archaeological sediments were collected from Nahal Boqer 66 Koike et al., 2010) are known to affect peak height and shape in IR (WGS84: 30.9086° N, 34.7906° E, 521 m a.s.l.), an open-air site located in the Negev Highlands (∼90 mm annual precipitation; Israel spectra, studies showed that plotting the ν2 and ν4 peak heights nor- Meteorological Service) on a small saddle between two Turonian malized to the ν3 height decouples these effects and provides in- formation on local atomic disorder (Regev et al., 2010; Poduska et al., limestone ridges (Avni and Weiler, 2013). The ridges in the study area 2011). Although more recent publications have shown limitations to are mantled by calcite-containing aeolian dust deposits. Samples (c. 10 specifying the underlying causes of crystalline disorder (Xu et al., g of sediment from various loci) were collected during excavations that 2015), the method reliably shows variation among calcites of diverse were carried out in 2016 by the authors and served for microremain origins and is useful for rapidly differentiating pyrogenic from geogenic analyses to address questions related to subsistence practices (Dunseth archaeological materials during routine on-site or lab-based FTIR et al., 2018). For the current study we selected 4 unequivocal degraded spectroscopy analysis (cf. Weiner, 2010). dung samples that according to the previous study originate from li- ff During the analysis of archaeological sediments at Nahal Boqer 66, vestock enclosures and had di erent concentrations of dung spher- Negev Highlands (modern Israel), it was noted that all sediment sam- ulites, ranging between 32 and 195 million per 1 gr of sediment ples plotted in a range that corresponds to Regev et al.'s (2010) modern (Table 1 and Appendix 1). Note that the analytical error in determining wood ash and lime plaster grinding curves (Fig. 1). However, none of concentrations of dung spherulites can be around ± 30% of the calcu- these samples were, or contained, lime plaster or large amounts of lated value (Table 1; for more details see Gur-Arieh et al., 2013). All wood ash. At first approximation, it was noted that the samples that four samples studied here have very high concentrations of spherulites plotted near and above the lime plaster grinding curve included high that correspond to those recorded in modern dung of sheep/goats, concentrations of dung spherulites. In order to test whether this ob- animals typically herded in the study region (Shahack-Gross and servation can be developed into a screening method for identification of Finkelstein, 2008). Samples NB-1.2 and NB-2.6 are associated with a dung-containing archaeological sediments, we present here an appli- degraded dung deposit dated by four radiocarbon determinations – σ cation of the FTIR spectroscopy grinding curve method. In this paper we spanning c. 3300 2900 BCE (2 , Dunseth et al., 2017: Table 3). Samples NB-3.3 and NB-9.2 are from excavated loci associated with

Fig. 1. Data points from all sediments studied from the Early and Intermediate Bronze Age site of Nahal Boqer 66 (Negev Highlands, Israel) plotted against the reference trends reported in Regev et al. (2010). Note location of aeolian dust controls in this study (green triangles) near geogenic trendlines, while all archae- ological sediments (brown circles) plot in the range between chalk/ash and lime plaster trendlines. Following infrared and optical analysis, none of the archaeological samples are composed of wood ash or lime plaster, or showed indicators of being severely affected by heat (Dunseth et al., 2018: Appendix 2). (n.a.u. = normalized absorbance units). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

119 Z.C. Dunseth, R. Shahack-Gross Journal of Archaeological Science 97 (2018) 118–124

Table 1 List of samples, location, date and calcitic microremain concentrations (reported in millions per 1 g of sediment/ash). Error of microremain concentrations is reported for two samples that were prepared in duplicate. (* denotes radiocarbon-dated contexts, Dunseth et al., 2017: Table 3).

Sample Description Site Date (* = 14C age) Dung Spherulites (millions/1 g Ash Pseudomorphs (millions/1 g sediment) sediment)

Modern Dung Sample W.ZEIT-15.3 ashed modern ovicaprine pellet Wadi Zeitan, Israel December 2015 616 1.7 Degraded Dung Sediments NB-1.2 degraded dung sediment Nahal Boqer 66, 3300-2900 BCE* 75 0 Israel NB-2.6 degraded dung sediment Nahal Boqer 66, 3300-2900 BCE* 132 0 Israel NB-3.3 degraded dung sediment + minor ash Nahal Boqer 66, c. 3000-1950 BCE 164 ± 45 0.3 ± 0.4 component (?) Israel NB-9.2 degraded dung sediment Nahal Boqer 66, c. 3000-1950 BCE 43 ± 15 0 Israel Modern Ash Samples ASH-4 carob (Ceratonia siliqua) 2013 0 218 ASH-5 oak (Quercus calliprinos) 2013 0 279 Control Samples NB-C2 aeolian dust Nahal Boqer 66, – 00 Israel NB-C7 limestone from wadi east of site Nahal Boqer 66, –– – Israel NB-C8 limestone from ridge immediately south of Nahal Boqer 66, –– – site Israel pottery from the Early Bronze (c. 3000–2500 BCE) and Intermediate 1993). Samples were analyzed in mid-IR range between 4000 and − − Bronze Ages (c. 2500–1950 BCE) (Cohen, 1999:60–61). As the site is 400 cm 1 at 4 cm 1 resolution using a Thermo Scientific Nicolet iS5 located on a calcareous terrane, aeolian dust and limestone samples with Omnic 9.3 software. Grinding curves were made following the were collected from the immediate landscape beyond the limits of the procedures outlined in previous studies of calcite, with the same KBr site as geogenic controls (Table 1). pellet ground to different extents using an agate mortar and pestle As dung was a common fuel source in antiquity, we used ashed (Regev et al., 2010; Poduska et al., 2011). A minimum of five grinds of modern ovicaprine dung as a control (collected from Wadi Zeitan, the same pellet were made per sample. Baselines for height measure- Western Negev, December 2015). Fresh dung was not used as control to ments were determined following Chu et al. (2008: 907), examples of avoid interference due to organic matter in the spectra. As one of the which are shown in Fig. 2. archaeological degraded dung samples also contained calcitic pseudo- morphs indicative of wood ash, we used modern wood ash of oak 3. Results (Quercus calliprinos) and carob (Ceratonia siliqua) as further controls. All ffl modern controls were ashed at 500 °C for 4 h in a mu e furnace Fig. 2 shows representative spectra from the degraded archae- fi (Thermo Scienti c Thermolyne F6000) and allowed to cool at room ological dung, local aeolian dust, local limestone and experimental temperature for at least 48 h before FTIR analysis. modern wood and dung ashes. Calcite peaks dominate all spectra. Note ff To evaluate mixing e ects on FTIR spectra and grinding curve data, the similarity in composition between the degraded dung and aeolian we created mixtures of pre-determined weight ratios from sample NB- dust samples, containing calcite, clay and quartz. The limestone and 3.3 (the archaeological degraded dung sediment with the highest con- modern wood ash samples are composed entirely of calcite. Narrowing centration of dung spherulites) and sample NB-C2 (aeolian dust without of absorbance bands and a slight shift of the ν3 carbonate peak − dung spherulites) (Table 1). Before mixing, degraded dung sediments (+1 cm 1 over 5 grinds) occurs with successive grinding of degraded and aeolian dust were lightly homogenized with an agate mortar and dung samples (Fig. 3). μ pestle so that all particles passed through a 500 m sieve to standardize Grinding curve trendlines of the materials studied here are shown in the range of particle sizes used in these experiments. comparison to trendlines reported in Regev et al. (2010) (Fig. 4). The trendlines of all four archaeological degraded dung sediments and the 2.2. Determination of microremain concentrations modern ashed dung control are clearly separated from those of aeolian dust and limestones from the vicinity of the site of Nahal Boqer. They Calcitic dung spherulites and ash pseudomorphs were extracted have a similar trend and position as modern and archaeological lime from controls, ash and degraded dung sediment samples and a mea- plasters reported earlier (Regev et al., 2010; Regev, 2011: Fig. 11; sured aliquot was evenly dispersed on a microscope slide following the Poduska et al., 2012; Xu et al., 2015, 2016; Goshen et al., 2017). The sodium polytungstate (SPT) method outlined by Gur-Arieh et al. aeolian dust and limestone from the Negev Highlands studied here (2013). Calcitic microremains were counted systematically in 16 overlap with Regev et al.'s (2010) geogenic limestones. The trendlines random fields of view at 400 × in plane-polarized (PPL) and cross-po- of both modern wood ash samples also overlap with those of Regev larized light (XPL) using a Nikon Eclipse 50i POL petrographic micro- et al. (2010). scope. Microremain concentrations from the archaeological and control Fig. 5 shows the trendlines obtained from grinding mixtures of sediments were previously quantified and reported in millions per 1 g of known ratios of degraded archaeological dung and aeolian dust. With sediment in Dunseth et al., 2018: Appendix 2). dilution of the degraded dung component, the position of the trendlines of the mixtures shift towards the pure aeolian dust trendline, reflecting 2.3. Fourier transform infrared (FTIR) spectroscopy the changes in relative proportions of the two end members. There is no trend between archaeological spherulite concentrations FTIR analysis of all samples followed the conventional KBr pellet and position on the grinding curve plot (Fig. 4). Results of our mixing method, utilizing about 0.3 g of the sediment/ash sample (Weiner et al., experiments (Fig. 5) rule out the possibility of post-depositional mixing

120 Z.C. Dunseth, R. Shahack-Gross Journal of Archaeological Science 97 (2018) 118–124

Fig. 3. Effect of grinding on spectra of a representative degraded dung sediment (NB-3.3). Note absorbance bands resolve after the second grind and narrow −1 with subsequent grinds. The ν2 peak location shifts from 875 to 876 cm with continued grinding.

infrared spectral attributes. As seen in the trendlines of their calcite grinding curves, it is easy to differentiate sediments enriched in dung spherulites from other calcareous sediments and rocks. The corre- spondence between our geogenic controls and previously published trendlines of geogenic materials (i.e., Regev et al., 2010, 2015) in-

dicates that our observations regarding the trendlines of normalized ν2 to ν4 peak heights are not artifacts of the studied site's local environ- ment but reflect the infrared signal of calcitic dung spherulites. The trendline of spherulite-rich deposits studied here is similar to and in the range of those previously reported for modern lime plaster, known to be composed of micritic calcite (Regev et al., 2010; Poduska et al., 2012). Lime plaster was recently shown through X-Ray Dif- fractometry (XRD) to be composed of highly disordered calcite (Xu et al., 2016). We tentatively suggest the similarity in trendlines between lime plaster and degraded dung deposits may mean that dung spher- ulites are also composed of calcite that is highly disordered. This might Fig. 2. Selected FTIR spectra of the materials used in this study: archaeological be related to the biogenic formation pathway of dung spherulites from degraded dung sediments, ashed modern dung, ashed modern wood, limestone an amorphous or another disordered precursor (cf. amorphous calcium and aeolian dust from the study area. Indicative calcite absorbance bands are carbonate: Addadi et al., 2003; Politi et al., 2008; Bots et al., 2012; labeled. Thin red lines indicate position of baselines used for calculation of peak other metastable calcium carbonates: Rodriguez-Blanco et al., 2014). heights. Note similarities between degraded dung, the dung control and aeolian Interestingly, the trendlines of deposits rich in dung spherulites is dust. The degraded dung sediments show a heterogeneous composition, dominated by calcite and also including unheated clay (absorption peaks in- higher in our study than trendlines of other biogenic calcites of marine − dicating structural water between 3600 and 3700 cm 1), quartz (peaks at 1083, origin (sea urchin spines and tests: Regev et al., 2010: Fig. 2). We also − 797, 778, and 695 cm 1), and carbonated hydroxylapatite (shoulders at 605 suggest considering the contribution of the acicular crystal habit of − and 565 cm 1). (a.u. = arbitrary units). (For interpretation of the references to calcite forming dung spherulites (Canti, 1997; Canti and Nicosia, 2018), colour in this figure legend, the reader is referred to the web version of this and/or the spherulitic arrangement itself, as possible causes for the article.) trendlines observed here. While these observations may merit further research, in this article, however, we would like to focus on the ar- (i.e., dilution of dung spherulites by aeolian dust) as such an instance chaeological implications of this study. would have resulted in change of trendline position. Additionally, all The coincidence of sediments rich in dung spherulites and the area spectra reported here (including the aeolian dust controls) are domi- where they plot on the calcite grinding curve shows that this char- nated by calcite, implying that there is no relationship between calcite acteristic can be used as a tool to effectively and rapidly differentiate concentration in the samples and their position on the grinding curve geogenic calcite-containing aeolian deposits and degraded dung de- plot. We therefore suspect that this lack of correspondence between posits during routine field or laboratory FTIR analysis of sediments. dung spherulite concentrations and position on the plot reflects an This is obvious in the observations at the site studied here—Nahal original archaeological signal, possibly related to differences in space Boqer 66—(Fig. 4), and we propose that it may also be a useful in- use at the site (e.g., different animal density at certain localities in the dicator in other contexts where dung spherulites are abundant. site or different types of stabled animals). The mixing experiments reported on here show that the signal strength for dung spherulites decreases with increased mixing with aeolian dust. The resultant trendlines of mixed sediments plot in an area 4. Discussion where chalk, flowstones, ash and aeolian dust also coincide. Site-spe- cific reference curves should be produced to utilize grinding curves We show here that archaeological sediments and modern ashed effectively in the study of archaeological sediments. As other studies dung containing abundant dung spherulites exhibit characteristic

121 Z.C. Dunseth, R. Shahack-Gross Journal of Archaeological Science 97 (2018) 118–124

Fig. 4. Grinding curves of all samples studied here in comparison to curves published in Regev et al. (2010). Dung spherulite (DS) and ash pseudomorph (AP) microremain concentrations are included for reference; all values are in millions per 1 g of sediment/ash. Note that archaeological dung sediments and ashed modern dung plot together and separately from pyrogenic wood ashes and geogenic sediments and rocks. Note also that dung-rich sediments and the modern dung control plot where lime plaster trendlines have been previously reported. (n.a.u. = normalized absorbance units). have noted, final characterization of sediments—and especially those 5. Conclusion whose grinding curves plot in the area where chalk, ash and aeolian dust overlap (Fig. 5)—should be confirmed with other methods, such as We show here that the application of the calcite FTIR grinding curve optical microscopy. method can be extended in archaeological contexts from looking into Previous research has shown that FTIR can be an effective tool for pyrogenic materials to also rapidly identifying sediments with high assessing the state of preservation of disordered pyrogenic carbonate concentrations of dung spherulites. While future application of this minerals to prescreen specimens for radiocarbon dating (Regev et al., observation in other contexts and regions has yet to be carried out, it 2011 and Poduska et al., 2012 for calcite in ash and plaster; Toffolo may open up new vistas for in-field identification of dung deposits et al., 2017 for aragonite in ash). If our suggestion regarding the high which may aid reframing research questions and excavation strategies disorder of calcite forming archaeological dung spherulites can be va- while excavation is on-going. The potential of dung spherulites as an lidated, then this attribute may be used to study the effect of diagenesis indicator of preservation, a resource for radiocarbon dating, a resource on dung spherulites, which may then be utilized as another potential for environmental reconstruction using their stable carbon, oxygen and archaeological resource for radiocarbon dating. Additionally, the nitrogen isotopes, as well as their formation pathways and taphonomy, carbon, nitrogen and oxygen in well-preserved calcitic dung spherulites are several open questions that may benefit archaeological investiga- have potential to be used for stable isotope paleoclimatic reconstruction tions with further research. or investigations into ancient animal diet (cf. Shahack-Gross et al., 2008).

Fig. 5. Grinding curves of quantitative mixtures (by weight % ratios) of degraded dung and aeolian dust. Note trend from 'pure dung' towards 'pure aeolian dust' with increasing ratio of dust relative to dung. This mixing experiment indicates that the utility of the grinding curve method for the identification of dung deposits is expected to diminish with increased post-depositional mixing of dung with geogenic deposits. (n.a.u. = normalized absorbance units).

122 Z.C. Dunseth, R. Shahack-Gross Journal of Archaeological Science 97 (2018) 118–124

Acknowledgements Dan David Scholarship for Archaeology and the Natural Sciences (2017- 2018). We thank all the volunteers that participated in the excavation This work was supported by the German-Israeli Foundation for of Nahal Boqer 66 in 2013 and 2016, and Guy Bar-Oz and Yotam Scientific Research and Development (GIF; Grant no. I-1244-107.4/ Tepper for the modern dung sample from Wadi Zeitan. Special thanks to 2014) to R.S.-G. and Markus Fuchs (Justus-Liebig-University Gieβen) as Lior Regev, David Friesem and Yotam Asscher for training and guidance principle investigators and Israel Finkelstein (Tel Aviv University) as on FTIR and the grinding curve method in the early stages of Z.C.D.’s co-investigator, and internal funds available at the Laboratory for research. Thanks also to Don Butler for insightful discussion on spher- Sedimentary Archaeology, The Leon H. Charney School of Marines ulites and the methods described here. Finally, we would like to thank Science, University of Haifa. In addition, Z.C.D. was supported by the the two anonymous reviewers who helped us improve the manuscript.

Appendix 1

Comparison between dung-derived sediments of the same sample weight (10.5 mg) with different dung spherulite concentrations. A) Focus- stacked crossed-polarized light (XPL) image of a representative field of view of sample NB-3.3 with the highest dung spherulite concentrations in this study; dung spherulites can be seen well-dispersed throughout the field of view, a result of the extraction and dispersion method (Gur-Arieh et al. 2013). Note that the image only captures 50% of the optical field of view. B) Focus-stacked XPL image of a representative field of view of sample NB- 9.2, with the lowest dung spherulite concentration in this study. C) Box plots of the number of spherulites counted in each of the 16 fields of view for both samples. Note the correspondence between the visual and statistical representations of the counting method, which must be based on knowledge of the weight of sediment taken for this type of analysis.

References Chu, V., Regev, L., Weiner, S., Boaretto, E., 2008. Differentiating between anthropogenic calcite in plaster, ash and natural calcite using infrared spectroscopy: implications in archaeology. J. Archaeol. Sci. 35, 905–911. Addadi, L., Raz, S., Weiner, S., 2003. Taking advantage of disorder: amorphous calcium Cohen, R., 1999. Ancient Settlement of the Central Negev, Vol. I: the Chalcolithic Period, carbonate and its roles in biomineralization. Adv. Math. 15, 959–970. the Early Bronze Age and the Middle Bronze I. (Israel Antiquities Authority, Avni, Y., Weiler, N., 2013. Geological Map of Israel: Sede Boqer (Sheet 18-IV) (1:50,000). Jerusalem). Geological Survey of Israel, (Jerusalem). Duyckaerts, G., 1959. The infra-red analysis of solid substances: a review. Analyst 84, Bots, P., Benning, L.G., Rodriguez-Blanco, J.-D., Roncal-Herrero, T., Shaw, S., 2012. 201–214. Mechanistic insights into the crystallization of amorphous calcium carbonate (ACC). Dunseth, Z.C., Junge, A., Fuchs, M., Finkelstein, I., Shahack-Gross, R., 2016. Cryst. Growth Des. 12, 3806–3814. Geoarchaeological investigation in the intermediate Bronze Age site of Mashabe sade. Brochier, J.E., Villa, P., Giacomarra, M., 1992. Shepherds and sediments: geo-eth- Tel Aviv 43, 43–75. noarchaeology of pastoral sites. J. Anthropol. Archaeol. 11, 47–102. Dunseth, Z.C., Junge, A., Lomax, J., Boaretto, E., Finkelstein, I., Fuchs, M., Shahack-Gross, Canti, M.G., 1997. An investigation into microscopic calcareous spherulites from herbi- R., 2017. Dating archaeological sites in an arid environment: a multi-method case vore dungs. J. Archaeol. Sci. 23, 219–231. study in the Negev Highlands, Israel. J. Arid Environ. 144, 156–169. Canti, M.G., 1998. The micromorphological identification of faecal spherulites from ar- Dunseth, Z.C., Finkelstein, I., Shahack-Gross, R., 2018. Intermediate Bronze Age sub- chaeological and modern materials. J. Archaeol. Sci. 25, 435–444. sistence practices in the Negev Highlands, Israel: Macro- and microarchaeological Canti, M.G., 1999. The production and preservation of faecal spherulites: animals, en- results from the sites of Ein Ziq and Nahal Boqer 66. J. Archaeol. Sci. Rep 19, vironment and taphonomy. J. Archaeol. Sci. 26, 251–258. 712–726. Canti, M.G., Nicosia, C., 2018. Formation, morphology and interpretation of darkened Goshen, N., Yasur-Landau, A., Cline, E.H., Shahack-Gross, R., 2017. Palatial architecture faecal spherulites. J. Archaeol. Sci. 89, 32–45. under the microscope: production, maintenance and spatiotemporal changes gleaned

123 Z.C. Dunseth, R. Shahack-Gross Journal of Archaeological Science 97 (2018) 118–124

from plastered surfaces at a Canaanite palace complex, Tel Kabri, Israel. J. Archaeol. dissertation. Weizmann Institute of Science, Rehovot, Israel. Sci. Rep. 11, 189–199. Regev, L., Poduska, K.M., Addadi, L., Weiner, S., Boaretto, E., 2010. Distinguishing be- Gur-Arieh, S., Mintz, E., Boaretto, E., Shahack-Gross, R., 2013. An ethnoarchaeological tween calcites formed by different mechanisms using infrared spectrometry: ar- study of cooking installations in rural Uzbekistan: development of a new method for chaeological applications. J. Archaeol. Sci. 37, 3022–3029. identification of fuel sources. J. Archaeol. Sci. 40, 4331–4347. Regev, L., Eckmeier, E., Mintz, E., Weiner, S., Boaretto, E., 2011. Radiocarbon con- Gur-Arieh, S., Shahack-Gross, R., Maeir, A.M., Lehmann, G., Hitchcock, L.A., Boaretto, E., centrations of wood ash calcite: potential for dating. Radiocarbon 53 (1), 117–127. 2014. The taphonomy and preservation of wood and dung ashes found in archae- Regev, L., Cabanes, D., Homsher, R., Kleiman, A., Weiner, S., Finkelstein, I., Shahack- ological cooking installations: case studies from Iron Age Israel. J. Archaeol. Sci. 46, Gross, R., 2015. Geoarchaeological investigation in a domestic Iron Age quarter, Tel 25–45. Megiddo, Israel. Bull. Am. Sch. Orient. Res. 374, 135–157. Karkanas, P., 2006. Late Neolithic household activities in marginal areas: the micro- Rodriguez-Blanco, J.-D., Shaw, S., Bots, P., Roncal-Herrero, T., Benning, L.G., 2014. The morphological evidence from the Kouveleiki caves, Peloponnese, Greece. J. Archaeol. role of Mg in the crystallization of monohydrocalcite. Geochem. Cosmochim. Acta Sci. 33, 1628–1641. 127, 204–220. Koike, C., Imai, Y., Chihara, H., Suto, H., Murata, K., Tsuchiyama, A., Tachibana, S., Rodriguez-Navarro, C., Jimenez-Lopez, C., Rodriguez-Navarro, A., Gonzalez-Munoz, Ohara, S., 2010. Effects of forsterite grain shape on infrared spectra. Geophys. J. 709, M.T., Rodriguez-Gallego, M., 2007. Bacterially mediated mineralization of vaterite. 983–992. Geochem. Cosmochim. Acta 71, 1197–1213. Korstanje, M.A., 2005. Microfossils in camelid dung: Taphonomic considerations for the Ruppin, R., Englman, R., 1970. Optical phonons of small crystals. Rep. Prog. Phys. 33 (2), archaeological study of agriculture and pastoralism. In: O'Connor, T. (Ed.), Biosphere 149–196. to Lithosphere: New Studies in Vertebrate Taphonomy. Oxbow Books, Oxford, pp. Shahack-Gross, R., 2011. Herbivorous livestock dung: formation, taphonomy, methods 69–77. for identification and archaeological significance. J. Archaeol. Sci. 38, 205–218. Lancelotti, C., Madella, M., 2012. The ‘invisible’ product: developing markers for iden- Shahack-Gross, R., Finkelstein, I., 2008. Subsistence practices in an arid environment: a tifying dung in archaeological contexts. J. Archaeol. Sci. 39, 953–963. geoarchaeological investigation in an Iron Age site, the Negev Highlands, Israel. J. Lane, M.D., 1999. Midinfrared optical constants of calcite and their relationship to par- Archaeol. Sci. 35, 965–982. ticle size effects in thermal emission spectra of granular calcite. J. Geophys. Res. Shahack-Gross, R., Marshall, F., Weiner, S., 2003. Geo-ethnoarchaeology of pastoral sites: Planets 104 (E6), 14099–14108. the identification of livestock enclosures in abandoned Maasai settlements. J. Matthews, W., French, C., Lawrence, T., Cutler, D., 1996. Multiple surfaces: the micro- Archaeol. Sci. 30, 439–459. morphology. In: Hodder, I. (Ed.), On the Surface Çatalhöyük 1993–95. McDonald Shahack-Gross, R., Simons, A., Ambrose, S.H., 2008. Identification of pastoral sites using Institute for Archaeological Research/BIAA, pp. 301–342. stable nitrogen and carbon isotopes from bulk sediment samples: a case study in Miller, N., Smart, T., 1984. Intentional burning of dung as fuel: a mechanism for the modern and archaeological pastoral settlements in Kenya. J. Archaeol. Sci. 35, incorporation of charred seeds into the archaeological record. J. Ethnobiol. 4, 15–28. 983–990. Monnier, G.F., 2018. A review of infrared spectroscopy in microarchaeology: methods, Shahack-Gross, R., Boaretto, E., Cabanes, D., Katz, O., Finkelstein, I., 2014. Subsistence applications and recent trends. J. Archaeol. Sci. Rep 18, 806–823. economy in the Negev Highlands: the Iron Age and the Byzantine/early Islamic Poduska, K.M., Regev, L., Boaretto, E., Addadi, L., Weiner, S., Kronik, L., Curtarolo, S., period. Levant 46, 98–117. 2011. Decoupling local disorder and optical effects in infrared spectra: differentiating Shillito, L.-M., Bull, I.D., Matthews, W., Almond, M.J., Williams, J.M., Evershed, R.P., between calcites with different origins. Adv. Mater. 23 (4), 550–554. 2011. Biomolecular and micromorphological analysis of suspected faecal deposits at Poduska, K.M., Regev, L., Berna, F., Mintz, E., Milevski, I., Khalaily, H., Weiner, S., Neolithic Çatalhöyük, . J. Archaeol. Sci. 38, 1869–1877. Boaretto, E., 2012. Plaster characterization at the PPNB site of Yiftahel (Israel) in- Sillar, B., 2000. Dung by preference: the choice of fuel as an example of how Andean cluding the use of 14C: implications for plaster production, preservation, and dating. pottery production is embedded within wider technical, social, and economic prac- Radiocarbon 54, 887–896. tices. Archaeometry 42, 43–60. Politi, Y., Metzler, R.A., Abrecht, M., Gilbert, B., Wilt, F.H., Sagi, I., Addadi, L., Weiner, S., Surovell, T.A., Stiner, M.C., 2001. Standardizing infrared measures of bone mineral Gilbert, P.U.P.A., 2008. Transformation mechanism of amorphous calcium carbonate crystallinity: an experimental approach. J. Archaeol. Sci. 28, 633–642. into calcite in the sea urchin larval spicule. Proc. Natl. Acad. Sci. Unit. States Am. 105 Toffolo, M.B., Regev, L., Mintz, E., Poduska, K.M., Shahack-Gross, R., Berthold, C., Miller, (45), 17362–17366. C.E., Boaretto, E., 2017. Accurate radiocarbon dating of archaeological ash using Polo-Diaz, A., Eguíluz, M., Ruiz, M., Pérez, S., Mújika, J., Albert, R.M., Fernández Eraso, pyrogenic aragonite. Radiocarbon 59, 231–249. J., 2016. Management of residues and natural resources at San Cristóbal rock-shelter: Valamoti, S.M., Charles, M., 2005. Distinguishing food from fodder through the study of contribution of the characterization of chalcolithic agropastoral groups in the Iberian charred plant remains: an experimental approach to dung-derived chaff. Veg. Hist. Peninsula. Quat. Int. 414, 202–225. Archaeobotany 14, 528–533. Portillo, M., Valenzuela, S., Albert, R.M., 2012. Domestic patterns in the Numidian site of White, W.B., 1974. The carbonate minerals. In: Farmer, V.C. (Ed.), The Infrared Spectra of Althiburos (northern ): the results from a combined study of animal bones, Minerals. Mineralogical Society, London, pp. 227–284. dung and plant remains. Quat. Int. 275, 84–96. Weiner, S., Goldberg, P., Bar-Yosef, O., 1993. Bone preservation in Kebara cave, Israel Portillo, M., Kadowski, S., Nishiaki, Y., Albert, R.M., 2014. Early Neolithic household using on-site Fourier transform infrared spectrometry. J. Archaeol. Sci. 20, 613–627. behavior at Tell Seker al-Aheimar (Upper Khabur, Syria): a comparison to eth- Weiner, S., 2010. Microarchaeology: Beyond the Visible Archaeological Record. noarchaeological study of phytoliths and dung spherulites. J. Archaeol. Sci. 42, Cambridge University Press, Cambridge, UK. 107–118. Xu, B., Toffolo, M.B., Regev, L., Boaretto, E., Poduska, K.M., 2015. Structural differences Prost, K., Birk, J.J., Lehndorff, E., Gerlach, R., Amelung, W., 2017. Steroid biomarkers in archaeologically relevant calcite. Anal. Methods 7, 9304–9309. revisited – improved source identification of faecal remains in archaeological soil Xu, B., Toffolo, M.B., Boaretto, E., Poduska, K.M., 2016. Assessing local and long-range material. PLoS One 12 (1), e0164882. https://doi.org/10.1371/journal.pone. structural disorder in aggregate-free lime binders. Ind. Eng. Chem. Res. 55, 0164882. 8334–8340. Reddy, S.N., 1999. Fueling the hearths in India: the role of dung in plaeoethnobotanical Zhang, C., Lv, J., Li, F., Li, X., 2017. Nucleation and growth of Mg-calcite spherulites interpretation. Paleorient 24, 61–70. induced by the bacterium Curvibacter lanceolatus strain HJ-1. Microsc. Microanal. 23, Regev, L., 2011. Exploiting Disorder in Calcite Crystals for Direct Radiocarbon Dating of 1189–1196. Plaster and Ash and Reconstructing Aspects of Past Human Behavior. PhD.

124 7.2 Chronology of IBA settlement in the Negev (Dunseth et al. 2017)

77 Journal of Arid Environments 144 (2017) 156e169

Contents lists available at ScienceDirect

Journal of Arid Environments

journal homepage: www.elsevier.com/locate/jaridenv

Dating archaeological sites in an arid environment: A multi-method case study in the Negev Highlands, Israel

* Zachary C. Dunseth a, d, , 1, Andrea Junge b, 1, Johanna Lomax b, Elisabetta Boaretto c, ** *** Israel Finkelstein a, , 2, Markus Fuchs b, 2, Ruth Shahack-Gross d, , 2 a Institute of Archaeology, Tel Aviv University, Tel Aviv 69978, Israel b Department of Geography, Justus-Liebig-University Giessen, 35390 Giessen, Germany c The Kimmel Center for Archaeological Science, Weizmann Institute of Science, Rehovot 76100, Israel d Laboratory for Sedimentary Archaeology, Department of Maritime Civilizations, University of Haifa, Haifa 3498838, Israel article info abstract

Article history: Archaeological surveys of the Negev Highlands show that the settlement history of this arid environment Received 29 November 2016 oscillated widely over time. This observation is almost entirely based on scant sherd assemblages from Received in revised form surveys, with only a few chronometric ages from one or two archaeological features at a given site. The 26 April 2017 reasons for the scarcity of chronometric ages include insufficient attention to radiocarbon dating in past Accepted 9 May 2017 research, low amounts of datable organic material for radiocarbon dating and issues related to low rate of Available online 20 May 2017 site accumulation, and incomplete preservation of activity remains. In order to overcome these problems, we present here the results of a detailed chronometric radiocarbon and optically stimulated lumines- Keywords: Negev Highlands cence (OSL) dating study exploring the development of Negev archaeological sites in the third millen- fi Early bronze age nium BCE. The study included micromorphological analyses to aid identi cation of sedimentological and Intermediate bronze age post-depositional processes at the studied sites. At Nahal Boqer 66, one of many small Negev third Radiocarbon millennium BCE sites, seven radiocarbon ages were determined from archaeological contexts that sug- Optically stimulated luminescence gest repeated discontinuous activity throughout the Early Bronze (EB) and early part of the Intermediate Arid environments Bronze Age (IBA) (c. 3300e2350 BCE). At Ein Ziq e one of a few large sites in the region e seven samples Micromorphology were dated; they show a very short period of activity in the beginning of the IBA (c. 2450e2200 BCE). OSL age determinations at this site provided evidence for the rapidity of site burial by sediment accumula- tion. Also, OSL ages from secure depositional contexts e verified via micromorphology e are in agree- ment with those obtained by radiocarbon dating. Taken together, the results provide new systematic evidence for the timing of EBeIBA activity in the arid Negev Highlands. © 2017 Elsevier Ltd. All rights reserved.

1. Introduction preferred but a single method cannot always address all aspects of human habitation and abandonment. In archaeology, the most Central to the investigation of archaeological settlements is the common method is radiocarbon dating, which targets organic determination of their place in time. Chronometric dating is materials deposited in conjunction with human activity. While radiocarbon dating is limited to the last 50,000 years, optically stimulated luminescence (OSL) dating covers even longer periods * Corresponding author. Institute of Archaeology, Tel Aviv University, Tel Aviv of time, serving as an important dating method in prehistory (e.g. at 69978, Israel. arid sites, Feathers et al., 2006; Olley et al., 2006; Holzer et al., 2010; ** Corresponding author. Armitage et al., 2011; Chazan et al., 2013; Porat et al., 2013; *** Corresponding author. Davidovich et al., 2014; Fattahi, 2015). Another main difference E-mail addresses: [email protected] (Z.C. Dunseth), andrea.junge@geogr. uni-giessen.de (A. Junge), [email protected] (J. Lomax), between radiocarbon and OSL dating is the dated event, with [email protected] (E. Boaretto), fi[email protected] radiocarbon dating the time of death of organisms associated with (I. Finkelstein), [email protected] (M. Fuchs), [email protected]. human activity (e.g., fire activities, food remains, burials) while OSL ac.il (R. Shahack-Gross). dates sedimentation events, therefore determining the time before, 1 Equal contribution. 2 Project directors, equal contribution. concurrent with, or after human activity (Boaretto, 2007; Junge http://dx.doi.org/10.1016/j.jaridenv.2017.05.006 0140-1963/© 2017 Elsevier Ltd. All rights reserved. Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169 157 et al., 2016). Complications for both dating methods include in ceramic typologies (e.g. survey results in Cohen, 1985), predomi- radiocarbon the possibility of a time gap between the death of the nantly based on sherd collection rather than large assemblages of dated organism and its deposition (e.g., ‘old wood effect’, Schiffer, vessels. However, several site reports include a few radiocarbon 1986; Olsen et al., 2013), and in OSL issues of mixing or incom- ages (e.g. Har Dimon, Nahieli and Tahal, 1993; Be'er Resisim, Ein Ziq plete bleaching are well-known (e.g. Porat et al., 2006; Araujo et al., and others, Segal, 1999; The Camel Site, Rosen, 2011), while other 2008; Feathers et al., 2010). sites were radiocarbon-dated, but are incompletely published (e.g. High-resolution dating of human activity within arid landscapes sites in the Uvda Valley, Avner and Carmi, 2001). More rarely, OSL is notoriously difficult. The main problem stems from the short- investigations of terraces in the Negev (e.g. Avni et al., 2013), and term nature of activity in harsh environments, resulting in sites occasionally at prehistoric open air sites (e.g. Holzer et al., 2010) that generally represent seasonal occupation rather than long-term have also been carried out. sedentism. Even when sites in arid environments show evidence Nevertheless, systematic site-specific chronometric dating in the for activities spanning more than one period, habitation gaps are Negev Highlands only began in the 2000s. A research project that characteristic. Therefore, while dating of multi-period sites in explores the settlement oscillations in the Negev Highlands, led by sedentary areas can rely on well-developed stratigraphic and two of the authors (I.F. and R.S-G.), demonstrated that the Iron Age ceramic sequences to constrain systematic chronometric schemes, wave of settlement was confined to c. 100 years spanning the late dating short-term sites e particularly those located in arid envi- 10th to late 9th centuries BCE (Shahack-Gross and Finkelstein, ronments e is more challenging. 2015). Though just slightly later than the traditional dating, this Apart from the absence of clear stratigraphic sequences, habi- investigation led to a novel historical reconstruction. It used tation phases of sites in arid environments are difficult to date radiocarbon age determination of short-lived charred materials because of thin occupation layers, reflecting the relatively short obtained from securely defined activity contexts such as hearths, duration of habitation (Banning and Kohler-Rollefson,€ 1992; Saidel sealed by c. 1 m thick deposits of collapsed constructional debris and Erickson-Gini, 2014). A further issue is post-depositional (stones and mud-based building materials) and covered by aeolian mixing at these sites, either by later human activity or due to deposits (Boaretto et al., 2010; Shahack-Gross et al., 2014). scavengers and faunal and floral turbation (e.g. in general, Wood The Negev Highlands Iron Age sites, despite having only one thin and Johnson, 1978; for radiocarbon, Boaretto, 2007; for OSL, occupation horizon and being rather short-lived, supplied well- Bateman et al., 2007; David et al., 2007). Post-depositional mixing preserved datable organic material for radiocarbon dating from and shifting of light/small artifacts (i.e., removal of top parts of clearly sealed depositional contexts. Yet, the second phase of the anthropogenically deposited materials) may also be caused by high Negev Highlands project, which explores the IBA settlement wave intensity precipitation or strong winds. Lastly, post-depositional in the region, encountered difficulties with dating. This is because processes may reduce the preservation of datable organic mate- the IBA sites generally consist of much shallower (thinly covered) rials (charcoal, bone) due to prolonged surface exposure (e.g. occupation deposits, possibly featuring different modes of activity. Behrensmeyer, 1978). It is therefore highly important to deploy all In our experience these sites rarely include charred organic mate- possible tools to successfully date archaeological sites in arid rials (Dunseth et al., 2016), a condition that underscores the environments. importance of applying OSL dating to archaeological contexts. An Recently we have conducted microarchaeological investigations initial study at the site of Mashabe Sade (Junge et al., 2016) has which included a dating component at Bronze and Iron Age sites in shown that despite lower precision relative to radiocarbon dating, the Negev Highlands, Israel e a region which belongs to the arid OSL age determinations can be valuable in desert sites, especially belt in the Southern Levant. Average temperature in the Negev when questions which do not require high-precision dating are at ranges from 10 C in the winter months (NovembereMarch) to stake or radiocarbon dating is not possible due to the lack of organic 25 C during the summer. Direct precipitation is highly variable material. Moreover, OSL dating not only provides ages of certain year to year (Evenari et al., 1971; Shanan, 2000: 88), averaging layers, but can also establish a chronostratigraphy of a sequence to between 80 and 150 mm, mostly concentrated during the winter allow the investigation of the sedimentation history. In addition, months (Goldreich, 2003). Local flora is dominated by Saharo- Junge et al. (2016) used micromorphology in tandem with OSL Arabian and (to a lesser extent) Irano-Turanian species such as dating which helped verifying that the dated sediments accumu- Zygophyllum dumosum, Retama raetam, and Artemisia herba alba lated following site abandonment. Micromorphology also contrib- (Danin and Plitmann, 1987). uted information on the modes of sediment deposition, showing Paleoclimatic studies suggest the Negev Highlands have been that post-abandonment, the circular stone structures typical of the arid throughout the last 5000 years (Babenko et al., 2007; for the central IBA sites act as traps for both wind-borne material and Southern Levant generally see Kagan et al., 2015; Langgut et al., water-lain sediments. Lastly, micromorphology showed that bio- 2015; Langgut et al., 2016), and relatively similar to current cli- turbation had a negligible effect on the sediment profile, thus matic conditions. Archaeological surveys (e.g. Cohen, 1981, 1985; rendering a rather precise dating sequence possible. Haiman, 1991, 1993; Avni, 1992; Rosen, 1994; Baumgarten, 2004) Here we present a chronological study that explored two attest to several waves of settlement during the past few millennia. archaeological sites in the region that exhibit IBA pottery (Fig. 1), Based on ceramic typologies they have traditionally been dated to employing OSL dating, radiocarbon dating and micromorphological the Early Bronze (EB) II (c. 3000-2900 BCE, Regev et al., 2012), In- investigations in tandem, in order to obtain a comprehensive pic- termediate Bronze (IBA, c. 2500-1950 BCE, Regev et al., 2012; also ture that allows chronometric dating with higher interpretational known as Early Bronze Age IV), Iron Age IIA (c. 940-780 BCE, certainty. One site e Ein Ziq e is large (often termed 'central'; Finkelstein, 2014), and Byzantine/Early Islamic (c. 300e900 CE, Haiman, 1996; Dunseth et al., 2016). According to ceramic evidence Magness, 2003) cultural-historical complexes. Other eras, such as it was inhabited during the IBA and revisited much later during the the Middle and Late Bronze Ages and medieval periods are hardly Nabatean period (Hellenistic-Roman era, c. 3rd century BCE e 2nd represented archaeologically (for detailed discussions, see Rosen, century CE; Cohen, 1999:137e188). The second site e Nahal Boqer 1987, 2011; Haiman, 1989; Avner and Carmi, 2001; Finkelstein, 66 e is much smaller and characterized by large courtyards, which 1995 and references therein). Sharp settlement oscillations thus are presumed to have served as animal pens (Cohen, 1999:60e61). characterize the history of the Negev Highlands. According to conventional pottery typology it was inhabited in two Most sites in the Negev Highlands have been dated only using periods e the Early Bronze Age II (radiocarbon dated to 3000e2900 158 Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169

Fig. 1. Satellite image showing the study area and location of sites mentioned in the text. Stars: sites studied here; yellow circles: sites mentioned in the text; small red circles indicate modern settlements. Adapted from Google Earth. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

BCE; Regev et al., 2012), and the IBA (radiocarbon dated to photographs of the six contexts studied. Samples for radiocarbon 2500e1950 BCE, the end date includes archaeological-historical dating were collected either by hand-picking during excavation, or considerations; Regev et al., 2012; Bietak, 2002). The aim of the following dry sieving of sediments from well-defined archaeolog- study presented here is to provide new ages for these two sites, and ical contexts using 1 mm mesh. discuss the methodological issues related to their dating. Samples for OSL dating were collected from two sections also selected for radiocarbon dating (Table 2; Fig. 3E, F). Three samples were taken from the north section of Square M21 (Fig. 3E) located 2. The sites and materials collected for analysis in an open area between the remains of two oval stone structures (Fig. 2, Area J). Within the section a greyish ash-rich layer (L. 14/J/ 2.1. Ein Ziq 16) indicates the occupation horizon. One sample was taken from the sterile sediment beneath the occupation layer (L. 14/J/20) (GI- Ein Ziq, the largest IBA site in the region (Cohen, 1999), is located 119), the second within the greyish layer (L. 14/J/16) (GI-120) and in a more desolate environment than the Negev Highlands proper, the third one in the aeolian accumulation above it (L. 14/J/10) (GI- c. 10 km southeast of the modern settlement of Sede Boqer (New 121). Israel Grid [NIG]: 186170/523870, 325 m a.s.l.). The site covers an OSL samples GI-122 and GI-123 were taken from the western area of 2 ha, with c. 200 oval structures spread over two Pleistocene section of Structure 14/A/4 (Fig. 3F), located within a partially alluvial terraces composed of limestone pebbles/boulders and excavated stone structure (Cohen's L.14, Cohen, 1999:139e141, aeolian dust. It is bordered by a dry wadi bed to the south. Sur- Fig. 86). Both samples were taken from the fine aeolian material rounding the area are outcrops of Paleocene marl (Taqiya Forma- trapped between the collapsed wall stones. Intact sedimentary tion) and two natural springs (Ein Ziq and Ein Shaviv) are located c. profiles were extracted next to the localities where samples for 1 km to the southwest of the archaeological site (see Avni and radiocarbon and OSL dating have been collected, and used for Weiler, 2013). micromorphological analyses (Fig. 3). Excavations in the early 1980s by R. Cohen yielded ceramics and copper ingots exclusively from the IBA, with the exception of a few intrusive Nabatean tombs (Cohen, 1999:137e188). Radiocarbon 2.2. Nahal Boqer 66 ages were produced from certain structures, with results falling mainly in the early IBA (according to the current dating scheme for Nahal Boqer 66 is a small site c. 4 km north of Sede Boqer (NIG: the EB and IBA e Regev et al., 2012) and one in the Islamic period 179900/535400, 521 m a.s.l.). Covering 0.2 ha, the site is situated on (see RT samples in Table 1). In 2014e2015, our team conducted two a small saddle between two low Turonian limestone ridges of the seasons of small-scale excavation at the site. Charcoal was present Nezer Formation (Avni and Weiler, 2013). The site is bordered on all in several clearly defined depositional contexts (e.g. floors, hearths, sides by narrow wadis. It is composed of two complexes, each of refuse pits) and from vertical sections (i.e. excavation baulks). small rooms arranged around a central courtyard (Fig. 4). Based on Seven samples were collected for radiocarbon dating. Table 1 lists small-scale excavations conducted in the 1970s and on conven- the new samples (annotated RTD), their archaeological context, and tional understanding of lithic and pottery typologies, Nahal Boqer botanical identification when determined. Fig. 2 indicates the 66 features evidence for human activity during three periods e Pre- location of samples reported in this study. Fig. 3 provides field pottery Neolithic B (PPNB, c. 7000 BCE, Gopher, 2012; Noy and Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169 159

Table 1 Results of radiocarbon analyses from Ein Ziq. All new samples (annotated RTD) are small thin charred branches of vegetal origin. Botanical identification conducted on 5 of the 7 samples indicates they are of local desert vegetation. Samples annotated RT, at the bottom of the table, are from previous reports.

Lab # Locus, basket, Height Archaeological Botanical ID C 14C age ± 1s Calibrated range Calibrated range Publication and field (m a.s.l.) context % (year BP) ±1s (BCE) ±2s (BCE) number

RTD-7678 L. 14/J/16 323.65 Refuse pit outside Anabasis sp. 68 3857 ± 39 2455 (14.0%) 2420 BCE 2465 (78.3%) 2270 BCE This study PT1/LB16 Structures 14/J/19 2410 (13.2%) 2375 BCE 2260 (17.1%) 2205 BCE and 14/J/21; assoc. 2350 (33.8%) 2280 BCE with OSL GI-120 2250 (6.4%) 2230 BCE and 2220 (0.9%) 2215 BCE micromorphology blocks EZB-1e3 RTD-7679 L. 14/J/16 323.65 Refuse pit outside Retama 64 3846 ± 39 2435 (3.5%) 2425 BCE 2460 (95.4%) 2200 BCE This study PT1/LB16 Structures 14/J/19 raetam 2400 (8.6%) 2380 BCE and 14/J/21; assoc. 2350 (35.9%) 2275 BCE with OSL GI-120 2260 (20.3%) 2210 BCE and micromorphology blocks EZB-1e3 RTD-7680 L. 14/J/21 323.75 Beaten-earth floor Anabasis sp. 66 3853 ± 37 2455 (11.6%) 2420 BCE 2460 (95.4%) 2205 BCE This study PT2/LB160 of Structure 14/J/21 2405 (11.2%) 2380 BCE 2350 (33.6%) 2280 BCE 2250 (8.4%) 2230 BCE 2220 (3.3%) 2210 BCE RTD-7681 L. 14/A/4 324.73 Mixed collapse Retama 62 3834 ± 37 2395 (1.9%) 2385 BCE 2460 (93.0%) 2200 BCE This study PT1/LB93 material within raetam 2345 (66.3%) 2205 BCE 2165 (2.4%) 2150 BCE Structure 14/A/4, assoc. with OSL GI- 122e123 and micromorphology block EZB-4 RTD-7682 L. 14/A/2 324.63 Simple hearth Retama 71 3861 ± 40 2455 (15.2%) 2420 BCE 2465 (80.3%) 2270 BCE This study PT1/LB66 within Structure raetam 2410 (14.6%) 2375 BCE 2260 (15.1%) 2205 BCE 14/A/2 2370 (38.4%) 2287 BCE RTD-8310 L. 15/F/TP3 323.10 Ash deposit outside n.d. 55 3866 ± 26 2455 (16.2%) 2420 BCE 2465 (90.7%) 2280 BCE This study PT1/R1 of Cohen's 2405 (15.9%) 2380 BCE 2250 (3.9%) 2230 BCE Structure 73 2350 (36.1%) 2290 BCE 2220 (0.8%) 2215 BCE RTD-8311 L. 15/K/8 323.38 Stone-lined hearth, n.d. 58 3867 ± 25 2455 (16.4%) 2420 BCE 2465 (91.0%) 2280 BCE This study PT1/R1 sealed by collapse 2405 (15.9%) 2380 BCE 2250 (3.7%) 2230 BCE of Structure 15/K/5 2350 (35.9%) 2290 BCE 2220 (0.7%) 2215 BCE RT-2514 Area A, L.13; e Floor of Cohen's Quercus sp. e 3700 ± 45 2190 (4.3%) 2180 BCE 2265 (0.4%) 2260 Carmi and Segal 1992: B.463 Room 13 2140 (63.9%) 2030 BCE BCE2205 (95.0%) 1950 126; Segal 1999: 338; BCE Avner and Carmi 2001 RT-885A Area B, L.53, e Cohen's Room 53 n.d. e 3960 ± 90 2580 (63.5%) 2335 2860 (3.3%) 2810 Segal and Carmi 2004: B.282 BCE2325 (4.7%) 2300 BCE2750 (1.3%) 2720 145 BCE BCE2700 (90.8%) 2200 BCE RT-885B Area C, L.79, e Cohen's Room 79 n.d. e 3850 ± 50 2453 (11.0%) 2420 2470 (93.5%) 2200 Segal 1999: 338 B.422 BCE2405 (10.3%) 2375 BCE2165 (1.9%) 2150 BCE2350 (30.6%) 2275 BCE BCE2255 (16.2%) 2210 BCE RT-885B1 Area C, L.79, e Cohen's Room 79 n.d. e 3880 ± 60 2460 (68.2%) 2290 BCE 2560 (1.5%) 2535 Segal 1999: 338 B.422 BCE2490 (92.5%) 2195 BCE2170 (1.4%) 2150 BCE RT-2213 Area H, L.21, e Lime pit Acacia sp. e 1080 ± 45 900 (17.8%) 920 CE950 780 (0.9%) 790 CE830 Segal and Carmi 1996: B. 176 (50.4%) 1015 CE (0.2%) 835 CE870 95 (94.3%) 1030 CE

Cohen, 1974), EBII and IBA (Cohen, 1985:42e46; Cohen 1999: 8; Figs. 4 and 5) where several bone fragments were collected by 60e61). According to sherds found within the structures, the ma- hand picking and charcoal fragments were collected following dry jority of the architectural remains were dated to the EBII, with some sieving of sediments using 1 mm mesh. Micromorphological reuse and construction activities occurring during the IBA. analysis was not conducted at the site. However, micro- Our team conducted one short test sampling at the site in 2013 archaeological analysis of the contexts that contained the charcoal and a small-scale excavation in 2016. Optimal archaeological con- and bones sampled for radiocarbon dating indicate that they are texts for dating (e.g. hearths) were not uncovered, charred remains dominated by high concentrations of phytoliths and dung spher- were rare, sediment cover in most loci was less than 30 cm, and the ulites. The latter association indicates presence of degraded live- sediment profile was poorly stratifiedd i.e., a clear distinction be- stock dung, which further implies that the samples were collected tween the possible periods of occupation at the site was not from contexts with evidence for herding activity (Dunseth et al., in observed. Fig. 4 shows the site plan. Depositional contexts selected preparation). Preliminary analysis indicated that collagen was well- for radiocarbon dating relate to two floors (Loci 16/NB/2 and 16/NB/ preserved in some of the bones. Following these pre-screening 160 Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169

Fig. 2. Plan of Ein Ziq (adapted from Cohen, 1999: Fig. 88). Red squares indicate newly excavated or sampled localities. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

analyses, seven samples were radiocarbon dated: 4 bones that Sample irradiation was performed with a 90Y/90Sr b-source (1.9 showed 1e4% (by weight) presence of well-preserved collagen and GBq), resulting in a dose rate for coarse-grain quartz on stainless 3 charcoal fragments (Table 3). steel cups of ca. 0.126 ± 0.003 Gy/s (20.03.2016). For the De deter- mination a single-aliquot regenerative-dose (SAR) protocol after Murray and Wintle (2000) was applied. 3. Methods The OSL signals were measured for 50 s at elevated tempera- tures (125 C) after a preheat at 180 C (10 s) and a cut-heat at 3.1. Radiocarbon 160 C for the natural and regenerated signals. The preheat and cut- heat temperatures were chosen after preheat-plateau test mea- Samples for radiocarbon dating were pre-treated according to surements and combined preheat-dose recovery tests (Wintle and material (i.e. bone, charcoal) following the protocols detailed in Murray, 2006). For D determination, the integral of the first 0.5 s of Yizhaq et al. (2005) and Boaretto et al. (2009). All samples provided e the quartz shine-down curves was used, after subtracting a back- a sufficient amount of carbon for 14C AMS measurement at the ground of 40e50 s from the signal. Aliquots with a 4 mm diameter Dangoor Research Accelerator Mass Spectrometer (D-REAMS) at (c. 500 grains; Fuchs and Wagner, 2003) were used for D deter- the Weizmann Institute of Science. All radiocarbon ages were e mination of the coarse-grain fraction. The calculation of the D calibrated using OxCal 4.2.4 (Bronk Ramsey, 2013) and the cali- e values was conducted by establishing growth curves with single bration curve IntCal13 (Reimer et al., 2013). exponential saturation functions. After passing the rejection criteria of 10% for the recycling ratio, the test dose error and the 3.2. OSL recuperation value, the mean De of the aliquots was calculated using the central age model after Galbraith et al. (1999) and the OSL sampling and sample preparation were performed resulting OSL ages were calculated using the parameters listed in following the protocols detailed in Junge et al. (2016), where the Table 2. More detailed information on the test and measurement samples were wet-sieved, treated with HCl and H2O2, followed by a procedures are given in Junge et al. (2016). density separation and a treatment with HF and HCl to extract the The dose rate (Ḋ) for OSL age calculation was determined by coarse grain (90e200 mm) quartz fraction. The OSL measurements thick source a-counting (U and Th) and neutron activation analysis to determine De were executed on a Lexsyg reader (Lomax et al., (K). Cosmic-ray dose rates were calculated according to Prescott 2014). For stimulation, the reader was operated with green LEDs and Hutton (1994). The water content of the samples was set to 7 (525 ± 25 nm), while the detection window was restricted to ± 5%. This value and its error represent the possible water content 350e400 nm (Hamamatsu H7360 PMT; filter: 5 mm Semrock range, based on the porosity of sandy silt and the climate condi- HC377/50 and 3 mm Schott BG3; Huntley et al., 1991; Lomax et al., tions. The used water content values were checked by measuring 2015). Stimulation with IR laser diodes (850 ± 3 nm) and detection the in situ water contents of the samples, showing conformity at 395e430 nm (filter: 3 mm Schott BG39 and 3.5 mm Semrock within the given errors. HC414/46) was used to check for the purity of the quartz extracts. Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169 161

Fig. 3. Field photographs showing the contexts sampled for radiocarbon dating (stars), OSL dating (dashed orange circles), and micromorphological analyses (dashed blue rect- angles). Dotted grey line indicates approximate ancient activity surface. (A) Stone-lined hearth (L. 15/K/8) sealed by the collapse of Structure 15/K/5. (B) Hearth (L. 15/J/14) in the eastern section of Structure 14/J/21; radiocarbon sample from floor immediately north (L. 14/J/21). (C) Hearth in east section of L. 14/A/2. (D) Ash deposit in L. 15/F/TP3. (E) North section of Square M21, including ash deposit (L. 14/J/16; between two dashed lines); sterile sediment below ash (L. 14/J/20), overlying aeolian sediments above ash (L. 14/J/10). (F) Fill and/or collapse above floor in the western section of L. 14/A/4. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Table 2 OSL data.

Sample ID Locus Sample Height U Th K Cosmic dose rate Ḋ Overdispersion De OSL age [ka BP] (depth below surface) [ppm] [ppm] [%] [Gy/ka] [Gy/ka] [%] [Gy]

GI-119 L. 14/J/20 323.63 m (41 cm) 2.41 ± 0.13 2.34 ± 0.40 0.39 ± 0.02 0.20 ± 0.03 1.24 ± 0.08 12.9 13.40 ± 0.53 10.9 ± 0.8 GI-120 L. 14/J/16 323.77 m (27 cm) 2.46 ± 0.12 2.22 ± 0.39 0.43 ± 0.02 0.21 ± 0.03 1.28 ± 0.09 11.4 5.39 ± 0.20 4.2 ± 0.3 GI-121 L. 14/J/10 323.84 m (20 cm) 2.36 ± 0.14 2.24 ± 0.46 0.37 ± 0.02 0.21 ± 0.03 1.20 ± 0.08 10.2 4.51 ± 0.16 3.8 ± 0.3 GI-122 L. 14/A/4 324.68 m (27 cm) 2.23 ± 0.12 1.44 ± 0.38 0.48 ± 0.02 0.21 ± 0.03 1.22 ± 0.08 9.4 6.40 ± 0.23 5.2 ± 0.4 GI-123 L. 14/A/4 324.80 m (15 cm) 2.52 ± 0.14 2.13 ± 0.46 0.45 ± 0.02 0.21 ± 0.03 1.31 ± 0.09 14.1 6.62 ± 0.28 5.1 ± 0.4

3.3. Micromorphology 4. Results

Micromorphological blocks were consolidated using polyester 4.1. Ein Ziq resin according to conventional procedures (Courty et al., 1989; Stoops, 2003). After curing, the blocks were sliced and thin sec- Charcoal samples from Ein Ziq prepared for the radiocarbon tions were prepared by Arizona Quality Thin Sections, Tucson, AZ. determination show a good state of preservation as the carbon Thin sections were observed using a Nikon Eclipse 50i POL petro- content (C%) in the pre-treated samples was between 55 and 71%. graphic microscope and described according to conventional ter- Radiocarbon dating from all 7 samples range primarily between minology (found in Courty et al., 1989; Stoops, 2003). c. 2450e2200 BCE (Fig. 6a). This range correlates with the first half of the IBA in Southern Levantine chronology (Regev et al., 2012). 162 Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169

Fig. 4. Plan of the site of Nahal Boqer 66 (adapted from Cohen, 1999: Fig. 41). Red rectangles show excavated areas noting the two loci sampled for radiocarbon dating (red stars). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 5. Field photographs of the two loci sampled at Nahal Boqer 66 for radiocarbon dating (red stars). (A) Locus 16/NB/2. (B) Locus 16/NB/8. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Despite being collected from loci spread across this relatively large Fig. 8 presents the dating results associated with their archaeo- site (Areas A, F, J and K, on both the upper and lower terraces) and logical context, and is reported in BCE. representing different archaeological contexts (collapse, floors, Micromorphological analysis shows that non-archaeological refuse pits, and hearths) the results are very similar. The de- sediments in the study area are dominated by calcitic-clay terminations also reflect good agreement with ages from R. Cohen's groundmass with quartz silt, characteristic of the Negev aeolian excavations in the 1980s (RT-885A, RT-885B, RT-885B1). Only one deposits (Crouvi et al., 2009). This groundmass includes limestone age from the previous excavation (RT-2415) lies outside of this very fragments, reflecting the alluvial nature of the study area. Bones, narrow settlement window. charcoal and ash are present in the archaeological sediments, For the OSL age determination, the samples from both localities sometimes associated with geogenic rocks and sediments. These within the site show a typical decay behavior under green stimu- remains are indicative of human activity in the pit context in Square lation. Also, the OSL signals appear sufficiently bright to allow a M21 e dated using both radiocarbon and OSL. Bones are either precise determination of equivalent doses (Fig. 7). The SAR protocol unburnt or burnt and may appear in vertical orientations (Fig. 9a). with a preheat/cut-heat temperature combination of 180/160 C Ash and charcoal are abundant and associated with aeolian com- proved to be suitable by the application of the dose recovery test, ponents (Fig. 9b). There is no evidence for heat at the bottom part of which allowed the replication of the given dose within 10% errors. the pit, meaning that no evidence for in situ combustion is present, In addition, the suitability of the sample material was shown by the which together with the vertical orientation of certain artifacts distinct 110 C TL-peaks and the purity of quartz demonstrated by indicates that the pit served as a dump for activity refuse. the absence of an IRSL signal. The De values of the samples show a The section from L. 14/A/4 is another context where OSL and unimodal and not significantly skewed distribution. The relative radiocarbon samples were collected in tandem. Micromorphology standard deviation of the samples ranged from 10 to 15%, and over- shows that this context is mixed. It includes a patch of human ac- dispersion ranged between 9% and 15%, indicating relatively nar- tivity remains and aeolian components (Fig. 9c, upper right) and a row De distributions. diagonal feature that is composed of fine aeolian groundmass and The analysis of the samples from the north section of Square coarse marl fragments, most of the latter in vertical orientations M21 yielded OSL ages of 10.9 ± 0.8 ka (GI-119), 4.2 ± 0.3 ka (GI-120) (Fig. 9d). The marl originates from the Paleocene Taqiya Formation and 3.8 ± 0.3 ka (GI-121). The calculated ages from the section of outcropping in the wadi below the site. It therefore cannot be of Structure 14/A/4 are 5.2 ± 0.4 ka (GI-122) and 5.1 ± 0.4 ka (GI-123). alluvial origin and must have been brought up to the site by Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169 163

Table 3 Results of radiocarbon analyses conducted on samples from Nahal Boqer 66. The botanical identification of the three thin charred branch samples is given; all are from local desert vegetation. Bone samples are too small to be identified to species.

Lab # Locus and Height Archaeological context ID C 14C age ± 1s year BP Calibrated range Calibrated range basket # (m a.s.l.) % ±1s (BCE) ±2s (BCE)

RTD-8671 16/NB/2 518.86e518.74 Phase I, collapse associated Zygophylum sp. 61 4331 ± 25 3010 (19.7%) 2985 BCE 3015 (95.4%) 2900 BCE PT2/R1 with grey dung and charcoal 2935 (48.5%) 2900 BCE rich sediment above bedrock floor, below Phase II architecture RTD-8672 16/NB/2 518.67 Phase I, grey dung- and Salsola tetranda 63 4346 ± 25 3010 (28.3%) 2980 BCE 3020 (95.4%) 2900 BCE PT3/R1 microcharcoal-rich sediment 2960 (6.4%) 2950 BCE directly on bedrock floor, below 2945 (33.5%) 2910 BCE Phase II architecture RTD-8675 16/NB/2 518.86e518.74 Phase I, grey dung- and B-sized medium 29 4414 ± 36 3095 (46.6%) 3010 BCE 3325 (13.5%) 3235 BCE PT3/LB3 microcharcoal-rich sediment mammal (femur, 2990 (21.6%) 2930 BCE 3175 (1.1%) 3160 BCE (Bone 3) directly on bedrock floor, below gnaw marks) 3120 (80.8%) 2920 BCE Phase II architecture RTD-8673 16/NB/2 518.86e518.74 Phase I, grey dung- and n.d. 34 4473 ± 35 3330 (45.3%) 3215 BCE 3340 (86.0%) 3080 BCE PT3/LB3 microcharcoal-rich sediment 3185 (10.1%) 3155 BCE 3070 (9.4%) 3025 BCE (Bone 4) directly on bedrock floor, below 3130 (12.8%) 3090 BCE Phase II architecture RTD-8670 NB16/8 518.59e518.49 Grey dung and ash-rich Zygophylum sp. 31a 3952 ± 36 2565 (17.1%) 2530 BCE 2570 (23.7%) 2515 BCE PT1/R1 sediment deposited on floor 2495 (33.3%) 2450 BCE 2505 (71.4%) 2340 BCE (Phase II?) 2420 (6.4%) 2405 BCE 2315 (0.3%) 2310 BCE 2380 (11.4%) 2350 BCE RTD-8674 16/NB/8 518.59e518.49 Grey dung and ash-rich B-sized medium 34 3966 ± 33 2565 (33.1%) 2525 BCE 2575 (85.2%) 2430 BCE PT1/LB1 sediment deposited on floor mammal 2500 (35.1%) 2460 BCE 2425 (4.1%) 2400 BCE (Bone 3) (Phase II?) (long bone) 2380 (6.0%) 2350 BCE RTD-8676 16/NB/8 518.59e518.49 Grey dung and ash-rich B-sized medium 38 4077 ± 35 2835 (9.5%) 2815 BCE 2860 (15.7%) 2810 BCE PT1/LB1 sediment deposited on floor mammal 2670 (53.6%) 2570 BCE 2750 (4.9%) 2720 BCE (Bone 2) (Phase II?) (thoracic vertebra) 2515 (5.1%) 2500 BCE 2700 (63.5%) 2560 BCE 2535 (11.3%) 2490 BCE

a Very small sample, 0.2 g initial weight. humans, most plausibly serving as construction material. 5. Discussion Charcoal for radiocarbon dating was sampled from various other contexts. Micromorphology analyses show that hearth contexts are This study provides new radiocarbon and OSL ages from two IBA especially well-preserved, including clear heated sediment sub- sites in the Negev Highlands, associated with sedimentological strates which are overlain by ash and charcoal (Fig. 9e); the lami- insights from micromorphology. nated microstructure of ash is intact (Fig. 9f).

5.1. Ein Ziq 4.2. Nahal Boqer 66 Ein Ziq is unique in its southeastern location relative to most Samples for radiocarbon dating included both charcoal and sites of this period in the Negev Highlands. The environment is animal bones. After pre-treatment the carbon percentage was slightly more arid, but the site is close to two springs. Due to the within the range for well-preserved charred botanical and faunal aridity of the area and accumulation of aeolian sediments, remains (Table 3). Only sample RTD-8670 had a low carbon per- anthropogenic material is relatively well-preserved, as attested by centage (31%). In this case the pre-treated sample weighed only the micromorphological observations and high carbon content in 0.7 mg, which provided 0.2 mg carbon after graphitization for the the dated organic materials. Unlike other Negev sites (e.g. Mashabe AMS measurement. Our understanding is that part of the material Sade, Dunseth et al., 2016), Ein Ziq also shows good preservation in after pre-treatment was not charcoal, but also included insoluble open spaces between stone structures. This environment thus (silicate) minerals. This sample also provided the youngest age in provides a prime case-study to verify the effectiveness of OSL in the Nahal Boqer series. We cannot exclude that there may have dating both sediment deposition associated with human activity been some modern contamination. (occupation layers) and the sediment accumulation after the In contrast to the situation at Ein Ziq, radiocarbon dating from abandonment of sites in the Negev. the 7 samples from Nahal Boqer 66 span a large range of ages The section studied in Square M21 shows the suitability of OSL (Fig. 6b). The two loci dated (16/NB/2 and 16/NB/8) are located in to date activity remains outside ancient buildings in the Negev. The different rooms of the Southern Complex (Cohen's L. 16 and 12 lowermost OSL sample, dating aeolian accumulation before human respectively). L. 16/NB/2 was sealed by later architecture, and thus activity, yielded an age of 10.9 ± 0.8 ka (GI-119). Within un- the radiocarbon samples are expected to belong to an early phase at certainties, there is an agreement between OSL (GI-120) and the site. According to the distribution of the radiocarbon ages, it radiocarbon ages from the anthropogenic feature L. 14/J/16 (Fig. 6c). appears that activity in L. 16/NB/2 took place between 3340 and Field and micromorphological observations of this ash-rich deposit 2900 cal BCE (2s) which corresponds to the archaeological periods suggest it is not related to in situ burning, but rather to dumping of EBIb and EBII. Radiocarbon ages from L. 16/NB/8 indicate activity materials following cleaning activities. The anthropogenic deposits between 2860 and 2310 cal BCE (2s), i.e., through the EBIII and are associated with aeolian components. The OSL ages can be early IBA. In terms of archaeological periods, this range spans from interpreted as showing complete bleaching of quartz grains related the Early Bronze Age Ib through the first half of the IBA. to possible cleaning activities (e.g., sweeping of sediments from 164 Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169

Fig. 6. Radiocarbon results. (a) Ein Ziq; new ages are in purple, previously published ages in dark grey (cf. Segal, 1999). (b) Nahal Boqer 66. (c) Comparison between radiocarbon and OSL results from Ein Ziq. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.) Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169 165

Fig. 7. (A) Luminescence properties of the sediments at both sites represented by the samples GI-120 (L. 14/J/16) and GI-122 (L. 14/A/4). Shown are OSL shine-down curves of the natural signal, TL curves of the first test dose cut-heat, and the growth curves. The graphs are created by the use of the function plot RLum() in R by using the luminescence package (R Luminescence Developer Team, 2015). (B) Kernel density estimation plot for each sample. For every sample, the accepted aliquots are shown in ascending order; with n: number of aliquots, RSD: relative standard deviation, De: equivalent dose. The graphs are created by the use of the of the function plot_KDE() in R by using the luminescence package (R Luminescence Developer Team, 2015).

floors and/or hearth/s within buildings or outside of them), as well Unlike the situation above, the ages obtained from L. 14/A/4 as contemporaneous aeolian input. Thus in this case it is evident show an offset by over a millennium between the radiocarbon and that aeolian sediment formation and deposition of human artifacts OSL results (Fig. 6c), with the latter being older than expected. occurred at the same time. Micromorphological observations show that the dated sediments The chronological overlap between GI-120 (4.2 ± 0.3), the are from two sources: Paleocene marl of the Taqiya Formation and anthropogenic ash deposit, and GI-121 (3.8 ± 0.3), aeolian sedi- aeolian sediment. Two interpretational options are thus possible: ments immediately above the ash deposit (Fig. 6c), suggest rapid (a) incorporation of un- or incompletely bleached quartz from the burial of the anthropogenic sediments. Paleocene marl; (b) incorporation of older aeolian and/or fluvial 166 Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169

Fig. 8. OSL and radiocarbon ages from Ein Ziq in context; all are calibrated to BCE. (A) Square M21, ash deposit in open space between stone structures. (B) Structure 14/A/4, fill and/ or collapse above floor.

Fig. 9. Micromorphological observations at dated loci from Ein Ziq. (a) Scan of a thin section (EZB-2) from the pit deposit, Square M21. Note the general reddish yellow fine grained groundmass, dotted by charcoal (black) and pottery (red) fragments, as well as coarser limestone pebbles (white). Note the vertical burnt bone fragment (arrow). (b) Higher magnification image of the pit deposit, showing charcoal, ash crystals (arrows) surrounded by fine rounded aeolian components (Plane Polarized Light, PPL). (c) Scan of a thin section from Structure 14/A/4 (EZB-4). Note the fine-grained sediment at the upper right that includes black and red fragments, representing anthropogenic deposit, and the diagonal feature that is composed of collapsed Paleocene marl fragments. (d) Higher magnification showing the association of anthropogenic fine-grained heated deposits con- taining charcoal (arrows) with marl fragments (PPL). (e). Scan of a thin section (EZB-5) of a hearth within Structure 14/J/21. Note the heated floor substrate (arrow) covered by ash and charcoal. (f) Higher magnification showing the heated floor substrate (bottom) overlain by undisturbed laminated ash containing charcoal (arrow) (PPL). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.) material from pre-habitation underlying layers. However, micro- developments during site use, such as the blocking of doorways at morphological observations do not support any of these possibil- Be'er Resisim (Dever, 1985: 20), or repurposing of structures as ities, as no quartz grains are seen within the Paleocene marl and no trash dumps at Be'er Resisim (see Building 7 in Dever, 2014: 169) upwards bioturbation was identified. We currently do not know and Mashabe Sade (Dunseth et al., 2016). The combination of a how to explain these older-than-expected OSL ages in this context. small plateau in the radiocarbon calibration curve and a lack of Ein Ziq is now the best-dated IBA site in the Negev Highlands, superimposed layers at the site, hamper attempts to check if this with 11 radiocarbon ages including those reported here and by timeframe can be narrowed any further. The pottery of Ein Ziq in Segal (1999:338e339). It is clear from the radiocarbon ages (all particular (Cohen, 1999:165e180), and the IBA in the Negev in from short-lived twigs of local Saharo-Arabian and Irano-Turanian general, hardly allows accurate observations of phases within the woody plants (Table 1), that activity at the site was limited almost period. Still, the majority of loci dated in this study and in Cohen's exclusively to the 25the23rd centuries BCE (Fig. 6a), that is the early excavation include red-slipped and burnished sherds (though very phase of the IBA. This relatively narrow range of ages from different few diagnostic), proposed as an early form of the IBA (Dever, 1980; depositional contexts across the entire site makes it unlikely that D'Andrea, 2012). we missed sampling later IBA occupation at the site. Indeed, unlike A single radiocarbon age (RT-2514; not produced by us) seems to other central IBA sites, there is little evidence for architectural represent a later visit/activity at the site. Incidentally, RT-2514 Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169 167

(Quercus sp.) is the only identified species not native to the Negev in terms of understanding the dated sedimentary units, and dis- Highlands, suggesting this material was brought from elsewhere tinguishing between disturbed and undisturbed contexts. and is unrelated to the construction or main period of habitation at The ages from Nahal Boqer and Ein Ziq together follow an the site. emerging pattern from the south of a developing settlement system The OSL ages from Square M21 are in agreement with the during the 3rd millennium BCE. Occupation at Ein Ziq is limited to radiocarbon ages within their respective errors (4.2 ± 0.3 ka, c. the early phase of the IBA (c. 25the23rd centuries BCE). Human 2500e1900 BCE), though due to the relatively large uncertainty activity at Nahal Boqer 66 e likely discontinuous e spans from the they span the entire IBA. The OSL ages cover a boarder time interval EBIb through the same early phase of the IBA, but no later. The due to the errors associated with this method, but this is also relationship between these two types of sites is still unclear. The caused by the physical size of the samples taken, which possibly new ages presented here provide a basis for future discussion about cover sediment accumulations representing several deposition the economy and history of the Negev Highlands. events. Overall, the combination of the three methods addresses the Funding interaction of depositional processes and materials collected for dating. This in turn provides a more secure dating of shallow sites. This work was supported by the German-Israeli Foundation for At Ein Ziq it was difficult to determine whether sediments accu- Scientific Research and Development (GIF; Grant No. I-1244-107.4/ mulated during or shortly after human activity, or were mixed. 2014 to R.S-G. and M.F. as principal investigators and I.F. as co- Thus, processes of sediment deposition and mixing have been investigator). addressed with the aid of micromorphology, affecting both radio- carbon and OSL interpretations. This approach can be applied to Acknowledgements sites elsewhere, in both arid and other regions. We thank all volunteers that helped in the excavation of the two 5.2. Nahal Boqer 66 sites. Special thanks go to Abra Spiciarich for identification of the faunal remains, Valentina Caracuta, Mark Cavanagh and Mordechay The complexity of dating small Negev sites is clearly seen in the Benzaquen for identification of the botanical samples for radio- case of Nahal Boqer 66. Unlike Ein Ziq, structures in Nahal Boqer are carbon dating, Eugenia Mintz and Lior Regev from the D-REAMS built directly on bedrock. There is no evidence of built roofs, shown AMS facility at the Weizmann Institute, and Manfred Fischer from to act as good dust traps at other sites (e.g. Mashabe Sade, Junge the Bayreuth University for U and Th measurements. We would also et al., 2016) and sediment accumulation is rather shallow and like to thank Associate Editor Liora Kolska Horwitz, the editor-in- poorly stratified. Because of the latter (see Avni et al., 2013), OSL chief and three anonymous reviewers for their helpful sugges- dating was not conducted. tions that improved the final manuscript. The radiocarbon ages for the site of Nahal Boqer 66 indicate (discontinuous) phases of activity throughout the EB and IBA Appendix A. Supplementary data (Fig. 6b). Cohen previously suggested that the site was occupied during the EBIIeIII and IBA (Cohen, 1985:42e46; though later this Supplementary data related to this article can be found at http:// was limited to the EBII and IBA, Cohen, 1999:60e61), based on dx.doi.org/10.1016/j.jaridenv.2017.05.006. pottery typology and evidence for architectural phasing. We found very few diagnostic sherds at the site; in L. 16/NB/2 there was a References single rim of a holemouth cooking pot (EBIIeIII, Sebanne et al., 1993) and not a single sherd was found in L. 16/NB/8. We doubt if Araujo, A.G.M., Feathers, J.K., Arroyo-Kalin, M., Tizuka, M.M., 2008. Lapa das boleiras the pottery found by Cohen allows distinguishing between phases rockshelter: stratigraphy and formation processes at a paleoamerican site in Central Brazil. J. Archaeol. Sci. 35, 3186e3202. within the Early Bronze Age. Radiocarbon ages from the two phases Armitage, S.J., Jasim, S.A., Marks, A.E., Parker, A.G., Usik, V.I., Uerpmann, H.-P., 2011. at the Camel Site (also dated by the excavator to the EBII and IBA), The southern route “out of Africa”: evidence for an early expansion of modern e are very similar to the ages at Nahal Boqer 66 (RTA-3083: humans into Arabia. Science 331, 453 456. e s e s Avner, U., Carmi, I., 2001. Settlement patterns in the Southern Levant deserts during 3082 2897 BCE 1 ; RTA-2043: 2856 2582 BCE 1 ; Rosen, 2011: the 6the3rd Millennia BC: a revision based on 14C dating. Radiocarbon 43, 59e65, Table 4.1). 1203e1216. The question whether central and small IBA sites existed Avni, G., 1992. Archaeological Survey of Israel e Map of Har Saggi Northeast (225). Israel Antiquities Authority, Jerusalem. simultaneously has been brought up in previous studies (e.g. Avni, Y., Weiler, N., 2013. Geological Map of Israel 1:50,000 Sede Boqer (Sheet 18- Finkelstein, 1989: 134; Dever, 2014: 227). The data presented here IV). The Geological Survey of Israel, Jerusalem. shows some chronological overlap between the two sites, but due Avni, G., Porat, N., Avni, Y., 2013. ByzantineeEarly Islamic agricultural systems in the Negev Highlands: stages of development as interpreted through OSL dating. to the very few other dated smaller sites (e.g. only the Camel Site, J. Field Archaeol. 38, 332e346. and possibly Yotvata 6 South; Segal and Carmi, 1996) this cannot be Babenko, A.N., Kiseleva, N.K., Plakht, I., Rosen, S., Savinetskii, A.B., Khasanov, B.F., completely addressed at this time. 2007. Reconstruction of the Holocene vegetation in the central Negev desert, Israel, on the basis of palynological data on the Atzmaut Zoogenic deposit. Russ. J. Ecol. 38, 417e426. 6. Conclusion Banning, E.B., Kohler-Rollefson,€ I., 1992. Ethnographic lesson for the prehistoric past: camp locations and material remains near Beidha, Southern Jordan. In: The combination of radiocarbon and OSL age determination Bar-Yosef, O., Khazanov, A. (Eds.), Pastoralism in the Levant: Archaeological Ma- terials in Anthropological Perspectives (Monographs in World Archaeology 10). highlights the advantages and limitations of dating archaeological Prehistory Press, Chicago, pp. 181e204. sites in an arid landscape. At both sites radiocarbon ages, with their Bateman, M.D., Boulter, C.H., Carr, A.S., Frederick, C.D., Peter, D., Wilder, M., 2007. relatively higher precision, allowed us to suggest periods of activity Detecting post-depositional sediment disturbance in sandy deposits using op- tical luminescence. Quat. Geochronol. 2, 57e64. and explore methodological issues of vertical and horizontal stra- Baumgarten, Y., 2004. Archaeological Survey of Israel e Map of Shivta (166). Israel tigraphy. At Ein Ziq OSL dating provided information on the rapidity Antiquities Authority, Jerusalem. of site burial. This provides another facet by which to explain the Behrensmeyer, A.K., 1978. Taphonomic and ecologic information from bone e fi weathering. Paleobiology 4 (2), 150 162. good preservation identi ed in this site and to understand site Bietak, M., 2002. Relative and absolute chronology of the Middle Bronze Age: formation processes. Micromorphological analyses added strength comments on the present state of research. In: Bietak, M. (Ed.), The Middle 168 Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169

Bronze Age in the Levant (Proceedings of an International Conference on MB IIA revolution. In: Gopher, A. (Ed.), Village Communities of the Pottery Neolithic Ceramic Material, Vienna 24the26th of January 2001). Osterreichischen Aka- Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of demie der Wissenschaften, Wein, pp. 29e42. Nahal Zehora, Vol. 3, Monograph Series of the Institute of Archaeology of Tel Boaretto, E., 2007. Determining the chronology of an archaeological site using Aviv University 29, Tel Aviv, pp. 1525e1579. radiocarbon: minimizing uncertainty. Israel J. Earth Sci. 56, 207e216. Haiman, M., 1989. Preliminary report of the western Negev Highlands emergency Boaretto, E., Wu, X., Yuan, J., Bar-Yosef, O., Chu, V., Pan, Y., Liu, K., Cohen, D., Jiao, T., survey. Israel Explor. J. 39, 173e191. Li, S., Gu, H., Goldberg, P., Weiner, S., 2009. Radiocarbon dating of charcoal and Haiman, M., 1991. Archaeological Survey of Israel e Map of Mizpe Ramon South- bone collagen associated with early pottery at Yuchanyan Cave, Hunan Prov- west (200). Israel Antiquities Authority, Jerusalem. ince, China. Proc. Natl. Acad. Sci. 106, 9595e9600. Haiman, M., 1993. Archaeological Survey of Israel e Map of Har Hamran Southeast Boaretto, E., Finkelstein, I., Shahack-Gross, R., 2010. Radiocarbon results from the (199). Israel Antiquities Authority, Jerusalem. Iron IIA site of Atar Haroa in the Negev Highlands and their archaeological and Haiman, M., 1996. Early Bronze Age IV settlement pattern of the Negev and Sinai historical implications. Radiocarbon 52, 1e12. deserts: view from small marginal temporary sites. Bull. Am. Sch. Orient. Res. Bronk Ramsey, C., 2013. OxCal 4.2.4. https://c14.arch.ox.ac.uk/oxcal/OxCal.html. 303, 1e32. Carmi, I., Segal, D., 1992. Rehovot radiocarbon measurements IV. Radiocarbon 34, Holzer, A., Avner, U., Porat, N., Horwitz, L.K., 2010. Desert kites in the Negev desert 115e132. and northeast Sinai: their function, chronology and ecology. J. Arid Environ. 74, Chazan, M., Porat, N., Sumner, T.A., Horwitz, L.K., 2013. The use of OSL dating in 806e817. unstructured sands: the archaeology and chronology of the Hutton sands at Huntley, D.J., Godfrey-Smith, D.I., Haskell, E.H., 1991. Light-induced emission spectra Canteen Kopje (Northern Cape Province, South Africa). Archaeol. Anthropol. Sci. from some quartz and feldspars. International J. Radiat. Applications and Ins- 5, 351e363. trum. Part D. Nucl. Tracks Radiat. Meas. 18, 127e131. Cohen, R., 1981. Archaeological Survey of Israel e Map of Sede Boqer-East (168). Junge, A., Lomax, J., Shahack-Gross, R., Dunseth, Z.C., Finkelstein, I., Fuchs, M., 2016. Israel Antiquities Authority, Jerusalem. OSL age determination of archaeological structures using trapped aeolian Cohen, R., 1985. Archaeological Survey of Israel e Map of Sede Boqer-West (167). sediments: a case study from the Negev Highlands, Israel. Geoarchaeology 31, Israel Antiquities Authority, Jerusalem. 550e563. Cohen, R., 1999. Ancient Settlement of the Central Negev, Vol. 1: the Chalcolithic Kagan, E.J., Langgut, D., Boaretto, E., Neumann, F.H., Stein, M., 2015. Dead Sea levels Period, the Early Bronze Age and the Middle Bronze Age I (Israel Antiquities during the Bronze and Iron Ages. Radiocarbon 57, 237e252. Authority, Reports, No. 6). Israel Antiquities Authority, Jerusalem (Hebrew with Langgut, D., Finkelstein, I., Litt, T., Neumann, F.H., Stein, M., 2015. Vegetation and English Summary). climate changes during the Bronze and Iron Ages (~3600e600 BCE) in the Courty, M.A., Goldberg, P., Macphail, R., 1989. Soils and Micromorphology in Southern Levant based on palynological records. Radiocarbon 57, 217e235. Archaeology. Cambridge University Press, Cambridge. Langgut, D., Adams, M.J., Finkelstein, I., 2016. Climate, settlement patterns and olive Crouvi, O., Amit, R., Porat, N., Gillespie, A.R., McDonald, E.V., Enzel, Y., 2009. Sig- horticulture in the Southern Levant during the Early Bronze and Intermediate nificance of primary hilltop loess in reconstructing dust chronology, accretion Bronze Ages (3600e1950 BC). Levant 48, 117e134. rates, and sources: an example from the Negev desert, Israel. J. Geophys. Res. Lomax, J., Kreutzer, S., Fuchs, M., 2014. Performance tests using the Lexsyg lumi- Earth Sci. 114, 1e16. nescence reader. Geochronometria 41, 327e333. Danin, A., Plitmann, U., 1987. Revision of the plant geographical territories of Israel Lomax, J., Mittelstraß, D., Kreutzer, S., Fuchs, M., 2015. OSL, TL and IRSL emission and Sinai. Plant Syst. Evol. 156, 43e53. spectra of sedimentary quartz and feldspar samples. Radiat. Meas. 81, 251e256. David, B., Roberts, R.G., Magee, J., Mialanes, J., Turney, C., Bird, M., White, C., R Luminescence Developer Team, 2015. Luminescence: Comprehensive Lumines- Fifield, L.K., Tibby, J., 2007. Sediment mixing at Nonda Rock: investigations of cence Dating Data Analysis. R Package Version 0.4.5. http://CRAN.R-project.org/ stratigraphic integrity at an early archaeological site in northern Australia and package¼Luminescence. implications for the human colonisation of the continent. J. Quat. Sci. 22, Magness, J., 2003. The Archaeology of Early Islamic Settlement in Palestine. 449e479. Eisenbrauns, Winona Lake. Davidovich, U., Goldsmith, Y., Porat, R., Porat, N., 2014. Dating and interpreting Murray, A.S., Wintle, A.G., 2000. Luminescence dating of quartz using an improved desert structures: the enclosures of the Judean desert, Southern Levant, re- single-aliquot regenerative-dose protocol. Radiat. Meas. 32, 57e73. evaluated. Archaeometry 56, 878e897. Nahieli, D., Tahal, G., 1993. Har Dimon. Hadashot Arkheologiyot 99, 89e90 Dever, W.G., 1980. New vistas on the EB IV (“MB I”) horizon in Syria-Palestine. Bull. (Hebrew). Am. Sch. Orient. Res. 237, 35e64. Noy, T., Cohen, R., 1974. Nahal Boqer: an early Pre-Pottery Neolithic B site. Mitekufat Dever, W.G., 1985. Village planning at Be’er Resisim and socio-economic structure in Haeven 13, 15e25 (Hebrew). Early Bronze Age IV Palestine. Eretz Isr. 18, 18*e28*. Olley, J.M., Roberts, R.G., Yoshida, H., Bowler, J.M., 2006. Single-grain optical dating > Dever, W.G., 2014. Excavations at the Early Bronze IV Sites of Jebel Qa aqir and of grave-infill associated with human burials at Lake Mungo, Australia. Quat. < Be er Resisim (Studies in the Archaeology and History of the Levant 6). Sci. Rev. 25, 2469e2474. Eisenbrauns, Winona Lake. Olsen, J., Heinemeier, J., Hornstrup, K.M., Bennike, P., Thrane, H., 2013. Old wood' Dunseth, Z.C., Junge, A., Fuchs, M., Finkelstein, I., Shahack-Gross, R., 2016. Geo- effect in radiocarbon dating of prehistoric cremated remains? J. Archaeol. Sci. archaeological investigation in the Intermediate Bronze Age site of Mashabe 40, 30e34. Sade, the Negev Highlands. Tel Aviv 43, 43e75. Porat, N., Rosen, S.A., Boaretto, E., Avni, Y., 2006. Dating the Ramat Saharonim late Dunseth, Z.C., Finkelstein, I., Shahack-Gross, R. Subsistence practices during the Neolithic desert cult site. J. Archaeol. Sci. 33, 1341e1355. Early and Intermediate Bronze Ages in the Negev Highlands: the micro- Porat, N., Avner, U., Holzer, A., Shemtov, R., Horwitz, L.K., 2013. Fourth-millennium- archaeological perspective, In preparation. BC “leopard traps” from the Negev desert (Israel). Antiquity 87, 714e727. D'Andrea, M., 2012. The Early Bronze IV period in south-central Transjordan: Prescott, J.R., Hutton, J.T., 1994. Cosmic ray contributions to dose rates for lumi- reconsidering chronology through ceramic technology. Levant 44, 17e50. nescence and ESR dating: large depths and long-term time variations. Radiat. Evenari, M., Shanan, L., Tadmor, N.H., 1971. The Negev: Challenge of a Desert. Meas. 23, 497e500. Harvard University Press, Cambridge, Massachusetts. Regev, J., Miroschedji, de P., Greenberg, R., Braun, E., Greenhut, Z., Boaretto, E., 2012. Fattahi, M., 2015. OSL dating of the Miam Qanat (KARIZ) system in NE . Chronology of the Early Bronze Age in the Southern Levant: new analysis for a J. Archaeol. Sci. 59, 54e63. high chronology. Radiocarbon 54, 525e566. Feathers, J.K., Holliday, V.T., Meltzer, D.J., 2006. Optically stimulated luminescence Reimer, P.J., Bard, E., Bayliss, A., Beck, J.W., Blackwell, P.G., Bronk Ramsey, C., dating of southern high plains archaeological sites. J. Archaeol. Sci. 33, Buck, C.E., Chang, H., Edwards, R.L., Friedrich, M., Grootes, P.M., Guilderson, T.P., 1651e1665. Haflidason, H., Hajdas, I., Hatte, C., Heaton, T.J., Hoffmann, D.L., Hogg, A.G., Feathers, J., Kipnis, R., Pilo, L., Arroyo-Kalin, M., Coblentz, D., 2010. How old is Luzia? Hughen, K.A., Kaiser, K.F., Kromer, B., Manning, S.W., Niu, M., Reimer, R.W., Luminescence dating and stratigraphic integrity at Lapa Vermelha, Lagoa Santa, Richards, D.A., Scott, E.M., Southon, J.R., Staff, R.A., Turney, C.S.M., van der Brazil. Geoarchaeology 25, 395e436. Plicht, J., 2013. IntCal13 and Marine13 radiocarbon age calibration curves Finkelstein, I., 1989. Further observations on the socio-demographic structure of the 0e50,000 years cal BP. Radiocarbon 55, 1869e1887. Intermediate Bronze Age. Levant 21, 129e140. Rosen, S.A., 1987. Demographic trends in the Negev Highlands: preliminary results Finkelstein, I., 1995. Living on the Fringe: the Archaeology and History of the Negev, from the emergency survey. Bull. Am. Sch. Orient. Res. 266, 45e58. Sinai and Neighboring Regions in the Bronze and Iron Ages (Monographs in Rosen, S.A., 1994. Archaeological Survey of Israel e Map of Maktesh Ramon (204). Mediterranean Archaeology 6). Sheffield Academic Press, Sheffield. Publications of the Israel Antiquities Authority, Jerusalem. Finkelstein, I., 2014. The southern steppe of the Levant ca. 1050e750 BCE: a Rosen, S.A., 2011. An Investigation into Early Desert Pastoralism: Excavations at the framework for a territorial history. Palest. Explor. Q. 146, 89e104. Camel Site. Negev (Cotsen Institute of Archaeology Monograph 69). Cotsen Fuchs, M., Wagner, G.A., 2003. Recognition of insufficient bleaching by small ali- Institute of Archaeology Press, Los Angeles. quots of quartz for reconstructing soil erosion in Greece. Quat. Sci. Rev. 22, Saidel, B.A., Erickson-Gini, T., 2014. A note on the excavation of an Ottoman and 1161e1167. British Mandate period Bedouin campground at Nahal Be’erotayim West in the Galbraith, R.F., Roberts, R.G., Laslett, G.M., Yoshide, H., Olley, J.M., 1999. Optical Negev desert, Israel. Arabian Archaeol. Epigr. 25, 138e145. dating of single and multiple grains of quartz from Jinmium Rock Shelter, Schiffer, M.B., 1986. Radiocarbon dating and the “old wood” problem: the case of northern Australia: Part I, experimental design and statistical models. the Hohokam chronology. J. Archaeol. Sci. 13, 13e30. Archaeometry 41, 339e364. Sebanne, M., Ilan, O., Avner, U., Ilan, D., 1993. The dating of Early Bronze Age set- Goldreich, Y., 2003. The Climate of Israel: Observation, Research, Application. tlements in the Negev and Sinai. Tel Aviv 20, 41e54. Springer, New York. Segal, D., 1999. Carbon-14 dating from early and Middle Bronze Age sites in Israel Gopher, A., 2012. The Pottery Neolithic in the Southern Levant e a second Neolithic and the region. In: Cohen, R. (Ed.), Ancient Settlement of the Central Negev, Vol. Z.C. Dunseth et al. / Journal of Arid Environments 144 (2017) 156e169 169

1: the Chalcolithic Period, the Early Bronze Age and the Middle Bronze Age I Berkowicz, S.M. (Eds.), The Hydrology-geomorphology Interface: Rainfall, (IAA Reports, No. 6). Israel Antiquities Authority, Jerusalem, pp. 336e339 Floods, Sedimentation, Land Use. IAHS Press, Jerusalem, pp. 75e106. (Hebrew). Stoops, G., 2003. Guidelines for Analysis and Description of Soil and Regolith Thin Segal, D., Carmi, I., 1996. Rehovot radiocarbon date list V. 'Atiqot 29, 79e106. Sections. Soil Science Society of America, Madison, Wisconsin. Segal, D., Carmi, I., 2004. Rehovot radiocarbon date list VI. 'Atiqot 48, 123e148. Wintle, A.G., Murray, A.S., 2006. A review of quartz optically stimulated lumines- Shahack-Gross, R., Finkelstein, I., 2015. Settlement oscillations in the Negev High- cence characteristics and their relevance in single-aliquot regeneration dating lands revisited: the impact of microarchaeological methods. Radiocarbon 57, protocols. Radiat. Meas. 41, 369e391. 253e264. Wood, W.R., Johnson, D.L., 1978. A survey of disturbance processes in archaeological Shahack-Gross, R., Boaretto, E., Cabanes, D., Katz, O., Finkelstein, I., 2014. Subsis- site formation. Adv. Archaeol. Method Theory 1, 315e381. tence economy in the Negev Highlands: the Iron Age and the Byzantine/Early Yizhaq, M., Mintz, G., Cohen, I., Khalaily, H., Weiner, S., Boaretto, E., 2005. Quality Islamic period. Levant 46, 98e117. controlled radiocarbon dating of bones and charcoal from the early Pre-Pottery Shanan, L., 2000. Runoff, erosion, and the sustainability of ancient irrigation sys- Neolithic B (PPNB) of Motza (Israel). Radiocarbon 47, 193e206. tems in the central Negev desert. In: Hassan, M.A., Slaymaker, O.,

7.3 Subsistence economies and trading systems of IBA Negev settlement (Dunseth et al. 2016, 2018)

92 Journal of Archaeological Science: Reports 19 (2018) 712–726

Contents lists available at ScienceDirect

Journal of Archaeological Science: Reports

journal homepage: www.elsevier.com/locate/jasrep

Intermediate Bronze Age subsistence practices in the Negev Highlands, T Israel: Macro- and microarchaeological results from the sites of Ein Ziq and Nahal Boqer 66

⁎ ⁎ ⁎ Zachary C. Dunsetha,b, , Israel Finkelsteina, ,1, Ruth Shahack-Grossb, ,1 a Institute of Archaeology, Tel Aviv University, Tel Aviv 6997801, Israel b Department of Maritime Civilizations, University of Haifa, Haifa 3498838, Israel

ARTICLE INFO ABSTRACT

Keywords: This study presents new macro- and microarchaeological data on the subsistence economy of Early Bronze (c. Negev Highlands 3500–2500 BCE) and Intermediate Bronze Age (c. 2500–1950 BCE) settlements in the arid Negev Highlands in Intermediate Bronze Age southern Israel. The data originates from two sites: Nahal Boqer 66, a small Early Bronze/Intermediate Bronze Subsistence practices site, and Ein Ziq, the largest central Intermediate Bronze Age settlement in the region. At Nahal Boqer 66 we Phytoliths identified ceramic evidence for mainly domestic cooking activities, clear microarchaeological evidence for Dung spatial division of human activity and penning livestock, and no macro- or microarchaeological evidence for Ash Copper cereal agriculture. At Ein Ziq, the ceramic assemblage suggests a strong connection to trade networks and spatial division of activity, while the microarchaeological data shows no indication of direct food production—neither herding nor agriculture—and no trace of copper processing activities, previously considered an important supplemental subsistence strategy at many Negev Intermediate Bronze Age sites. We interpret the small Negev sites, such as Nahal Boqer 66, as representing the indigenous pastoral population, and the central sites as trading posts on the way to the coastal plain and Egypt. We explain the Early Bronze and Intermediate Bronze Age settlement patterns in the Negev Highlands on the background of contemporary geo-political transformations in the Levant and Egypt.

1. Introduction In this study we deal with the settlement history of the Negev Highlands in the third millennium BCE, with special emphasis on the The last 5000 years in the arid Negev Highlands (southern Israel) IBA. In the Mediterranean region, the EB/IBA transition is characterized are characterized by sharp settlement oscillations; several periods fea- by the decline and abandonment of urban centers and the system of ture strong evidence for human activity, while others lack or have city-states, and a shift to rural subsistence (for possible causes and scarce human remains (e.g., Rosen, 1987, 2011a, 2016; Finkelstein, discussion see e.g., Dever, 1989; Esse, 1991; Rosen, 1995; de 1995; Shahack-Gross and Finkelstein, 2015). The former include a Miroschedji, 2009; Greenberg, 2017). However, at the same time the phase in the Early Bronze Age (hereafter EB, commonly identified as the arid Negev Highlands display evidence for prosperity rather than de- EB II, c. 3000–2900 BCE, for the dates see Regev et al., 2012), Inter- cline. The IBA settlement peak in the Negev Highlands features two mediate Bronze Age (hereafter IBA, also known as EB IV, c. main site types: four large ‘central’ sites (Fig. 1) that are composed of 2500–1950 BCE, Regev et al., 2012), Iron Age IIA (c. 940–780 BCE, dozens to hundreds of circular stone structures and open spaces be- Finkelstein, 2014), the Nabatean period (c. 170 BCE–100 CE, Erickson- tween them, and hundreds of small sites composed of several stone built Gini, 2010) and Byzantine/Early Islamic period (c. 325–900 CE, rooms set around open enclosures (e.g., Finkelstein, 1995; Haiman, Magness, 2003). The cause/s for these transformations—climate or 1996; Dunseth et al., 2016). human induced—have been discussed by many scholars (e.g., Rosen, Paradigms concerning subsistence strategies in the Negev Highlands 1987; Finkelstein, 1995; Palumbo, 2001); the current paradigm seems have been based on a combination of circumstantial evidence and ar- to exclude climate as a major factor behind the settlement history of the chaeological factors. The former includes: 1) location of sites an- region (Shahack-Gross and Finkelstein, 2015). d—specifically—proximity to ancient terraced wadi systems (e.g.,

⁎ Corresponding authors. E-mail addresses: [email protected] (Z.C. Dunseth), fi[email protected] (I. Finkelstein), [email protected] (R. Shahack-Gross). 1 Project co-directors. https://doi.org/10.1016/j.jasrep.2018.03.025 Received 9 October 2017; Received in revised form 6 March 2018; Accepted 26 March 2018 2352-409X/ © 2018 Elsevier Ltd. All rights reserved. Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

Fig. 1. Aerial photographs showing the study area and selected sites mentioned in the text. (A) Satellite image from Google Earth (Landsat/Copernicus, dated 31.12.2016). (B) Archaeological sites and modern settlements for reference. The sites reported in this study—Nahal Boqer 66 and Ein Ziq—are marked black in the center.

Evenari et al., 1958; Cohen and Dever, 1981; Dever, 2014); and 2) Indirect indicators of subsistence strategies such as lithics and assumptions based on 19th–20th century Bedouin practices (see ground stone items have also been studied from EB and IBA sites in the Dunseth et al., 2016 for summary and considerations). The archae- region. Sickle blades form a miniscule proportion of IBA lithic assem- ological evidence is comprised of: 1) architectural similarities to pre- blages (Vardi et al., 2007; Rosen et al., 2014a; also, Saidel, 2002a; modern Bedouin tents and semi-permanent structures (Finkelstein, Saidel et al., 2006). Sickle gloss forms from abrasive activities and has 1995; Haiman, 1996); 2) scarce zooarchaeological remains (Hakker- been shown experimentally to form through the repeated cutting or Orion, 1999; Saidel, 2002a; Saidel et al., 2006); and 3) lithic material, reaping of any plant material (e.g., Anderson, 1999; recently Rosen namely the presence of blades with ‘sickle gloss’ and grinding stones et al., 2014b contra Anderson, 1980) thus it is not a valid indicator for (e.g., Finkelstein, 1989; Cohen, 1992; Haiman, 1996). the processing of cultivated plants. Assumptions have also been made Haiman (1996) emphasized the importance of copper trade for the based on the layout of sites and their parallels to premodern Bedouin IBA inhabitants of the Negev Highlands, especially in light of ingot settlements; mainly that internal enclosures represent livestock pens. hoards found at central Negev sites (e.g., Har Yeruham, Kochavi, 1967; However, the presence of animal dung in these enclosures has not been Be'er Resisim, Cohen and Dever, 1980; Ein Ziq, Cohen, 1999: 137–188). determined (cf. Rosen, 2011b: 203). Based on spatial and architectural affinities, Haiman (1996) postulated Recent ethno- and geo-ethnoarchaeological studies have demon- the existence of two complementary socioeconomic phenomena in the strated the importance of animal dung and its constituents to re- Negev, attributing central sites to copper processing/production activ- constructing human subsistence practices (e.g., Brochier et al., 1992; ities, and the smaller sites to herding-based economy. Subsequent stu- Reddy, 1999; Shahack-Gross et al., 2003; Valamoti and Charles, 2005). dies have supported Haiman's copper paradigm, based on new large- Specifically, durable inorganic remains of animal dung—namely, cal- scale excavations at the copper production site of Hamra Ifdan (Adams, citic dung spherulites and opaline phytoliths—have been found to be 2000; Levy et al., 2002; Gidding, 2016; Ben-Yosef et al., 2016) as well effective indicators of herding and grazing practices, as well as fod- as lead isotope provenance studies of ingots found at Negev sites (Segal dering strategies (Shahack-Gross, 2011 and references therein). et al., 1996–1997; Hauptmann et al., 2015). However, data on other Earlier work in the Negev Highlands examined micro-remains from forms of subsistence practices—i.e., herding and agriculture—- Iron Age (Shahack-Gross and Finkelstein, 2008) and late Byzantine- continued to be based on indirect evidence and interpretation of scarce Early Islamic sites (Shahack-Gross et al., 2014), comparing them to data remains (more below). collected in modern and sub-modern Bedouin agro-pastoralist en- In this report we present the results of our work at the small EB–IBA campments. These studies, as well as works on dung identified in sites site of Nahal Boqer and the central IBA site of Ein Ziq. Integrating from areas with higher biomass and annual precipitation (e.g., the macroscopic and microscopic evidence, we discuss the subsistence Mediterranean coastal site of Tel Dor, Albert et al., 2008), indicated the practices in the Negev Highlands in the third millennium BCE. possibility to distinguish between the practices of free-grazing (pure pastoralism) and raising animals with the addition of agricultural by- 1.1. Identifying EB–IBA subsistence strategies in the Negev Highlands products (agro-pastoralism) using phytolith assemblages in well-dated dung deposits (also Shahack-Gross, 2011, 2017). Animal dung from Understanding subsistence strategies in the Negev Highlands has free-grazing herds in desert areas is typified by low phytolith con- been based on studies of meager zooarchaeological and botanical re- centrations and less than 1% presence of dendritic morphotypes, as mains. In general, at central IBA sites the faunal assemblage is domi- animals primarily feed on phytolith-poor local shrub vegetation nated by meat-bearing parts of young animals, suggesting a consumer (Shahack-Gross and Finkelstein, 2008). Animal dung related to agro- economy (Hakker-Orion, 1999) and mature animals at EB–IBA small pastoral activities in the Negev is characterized by higher phytolith sites, suggesting a producer economy (Saidel et al., 2006). Archae- concentrations associated with at least 3–4% of dendritic morphotypes, obotanical studies have been limited to wood charcoal from hearth as animals feed on both phytolith-poor local shrub vegetation and the contexts (e.g., Warnock, 1991; Baruch, 1999). Pollen analysis from straw and chaff byproducts of domestic cereals (Shahack-Gross et al., degraded dung from the Atzmaut rock-shelter highlighted seasonal 2014). Using this approach, it was asserted that Iron IIA livestock herds grazing of caprine herds (Babenko et al., 2007). in the Negev Highlands were free grazing, with no evidence for the

713 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726 practice of dry farming. For the Byzantine-Early Islamic period, phy- (PPNB) activity (Noy and Cohen, 1974), while the structures were tolith analysis supported earlier textual evidence for agro-pastoral ac- dated by ceramic typology to the EB II and the IBA (Cohen, 1985: 40; tivities (Shahack-Gross and Finkelstein, 2015, 2017). Cohen, 1999:61–62). Our radiocarbon investigation at the site pro- Dunseth et al. (2016) recently studied the central IBA site of Ma- vided evidence for activity from the EBIb through the first half of the shabe Sade and a smaller site in its immediate vicinity (Cohen, 1999: IBA (c. 3300–2350 BCE) in the (probable) form of discontinuous, short- 137). Based on the microarchaeological signatures, the two sites term visitations (Dunseth et al., 2017). The distribution of pottery (and showed a marked difference in subsistence strategies, namely the ac- to a lesser extent, minor architectural changes) identified in the earlier cumulation of ancient animal dung at the smaller site, and a complete excavation showed two distinct occupations, with the excavated por- absence of dung at the central site. An optically stimulated lumines- tions of the Northern Complex dominated by IBA sherds, and the cence (OSL) investigation showed that while sediment accumulation at Southern Complex representing a mix of EB and IBA sherds (Cohen, the central site immediately postdated the IBA (c. 1950–1650 BCE), the 1999: Fig. 41). No faunal or botanical remains were reported from the dung accumulation at the smaller site occurred during the Iron Age (c. earlier excavations. 900–500 BCE) (Junge et al., 2016). We conducted two short excavations at Nahal Boqer 66 in 2013 and The results raised two questions: 2016. Twelve large probes (ranging from 0.5 × 1 m to 1.5 × 3 m) and six small probes (less than c. 0.5 × 0.5 m) were dug in the two com- 1) Is the absence of evidence for food production (herding, agriculture) plexes, totaling c. 25 m2 of exposure (~1% of the site's area) (Fig. 2). An a phenomenon particular to Mashabe Sade, or characteristic of IBA additional rectangular structure was probed on the ridge immediately central sites in general? southwest of the site (L. 16/NB/11). Emphasis was given to exploring 2) Are small and central sites of the IBA contemporary, or do they areas partially excavated by Cohen (for direct comparison), as well as represent separate phases of settlement during the half-millennium- sampling unexcavated structures and the large central courtyards to long IBA? highlight spatial differences across the site. The excavations also fo- cused on the collection of all possible botanical, faunal, lithic and Addressing the chronology question, the radiocarbon investigations ceramic material: all sediments were sifted through 5 mm mesh; grey in Dunseth et al. (2017) indicated that a small site—Nahal Boqer sediments, shown to be often related to dung and/or ash deposits 66—shows discontinuous activity throughout the EB to the IBA (c. (Shahack-Gross et al., 2014), were further sieved through 1 mm mesh. 3300–2300 BCE), while a central site—Ein Ziq—is limited to the early A standard 5 × 5 m grid was laid over the site, and points were re- part of the IBA (c. 2450–2200 BCE). This showed that both the small corded on the New Israeli Grid in three dimensions with a Leica TS06 and the central sites were contemporary, and were active only during Total Station. Photogrammetric 3D models were created for each ex- the first part of the IBA (c. 2450–2200 BCE). cavated locus using Agisoft PhotoScan following the procedures out- This paper deals with the question of subsistence practices. It pre- lined by Prins (2016), and site plans were produced from the geor- sents the macro- and micro-archaeological results from the small site of ectified models. Nahal Boqer 66 and the central site of Ein Ziq. Sediment samples (c. 10 g) were collected in plastic vials according to macroscopic differences in color and texture, mainly from vertical 2. Fieldwork sections. Control samples were taken from ridges and wadis beyond the limits of the site in the immediate vicinity. A total of 62 sediment 2.1. Nahal Boqer 66 samples were collected; 46 were used for reconstructing subsistence practices and 6 as controls (Appendix 1). Nahal Boqer 66 is a small site (0.2 ha) c. 4 km north of modern Sede Boqer (New Israeli Grid [NIG]: 179900/0535400, 521 m a.s.l.; Fig. 1). 2.2. Ein Ziq The site is located in a small saddle between two low Turonian lime- stone ridges of the Nezer Formation (Avni and Weiler, 2013). It is Ein Ziq (NIG: 186170/523900, 325 m a.s.l.; Figs. 1, 2) is the largest composed of two stone-built complexes typical to the Negev Highlands (c. 2 ha) IBA site in the Negev Highlands, spreading over a pair of al- (see Finkelstein, 1995:37–49; Haiman, 1996), each with oval or rec- luvial terraces above the dry wadi bed of Nahal Ziq. Two natural tangular rooms (c. 3–5 m in diameter) attached to large courtyards springs—Ein Ziq and Ein Shaviv—are in the immediate vicinity, c. 1 km (8–14 m in diameter) (Fig. 2). Most walls are preserved only to 1–2 southwest of the site. courses (c. 50–60 cm), and are made of limestone blocks from the im- The site was extensively excavated (110 rooms/structures out of mediate vicinity. over 200 visible on the surface; Fig. 3) in the early 1980s by Cohen The site was excavated in the 1970s by R. Cohen (1985: 40–41, (1999: 137–188). Structures at the site (c. 2–4 m in diameter) are all 1999: 60–62). Lithic evidence suggests earlier Pre-Pottery Neolithic B stone-built, preserved in some areas up to 1 m. Ceramics found in

Fig. 2. Site plan of Nahal Boqer 66. Excavated areas are indicated in red; sam- pled sections are marked by black lines. Red letters correspond to field photos in Fig. 4. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) Plan adapted and redrawn from Cohen (1999): Fig. 41.

714 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

Fig. 3. Site plan of Ein Ziq. New excavated areas and recleaned sections are marked in red. Loci in areas A, C and F are labeled in blue (detailed plans of Areas J and K are presented in Fig. 7); green numbers indicate original loci as ex- cavated by R. Cohen (1999). Red letters corre- spond to field photos in Fig. 8. (For interpreta- tion of the references to color in this figure legend, the reader is referred to the web version of this article.) Plan adapted and redrawn from Cohen (1999): Fig. 88.

excavation showed one main period of activity, during the IBA, with samples, including 11 controls; 73 samples were used for reconstructing some reuse of structures as Nabatean tombs (e.g., Cohen, 1999: 142), subsistence practices (Appendix 2). and a lime kiln radiocarbon-dated to the Early Islamic period (Segal and Carmi, 1996: 95). Radiocarbon dates from the IBA structures in both 3. Laboratory methods Cohen's and our excavations show almost exclusively a short range of activity limited to the first half of the IBA (c. 2450–2200 BCE, Dunseth 3.1. Fourier transform infrared (FTIR) spectroscopy et al., 2017). Faunal remains recovered in Cohen's excavations, though limited FTIR analysis of all samples was conducted to determine the mi- (n = 242 NISP), showed a preference for meat-bearing parts of younger neralogical composition of sediments and to evaluate if sediments had ovicaprines (c. 88%), as well as a smaller component (c. 10%) of wild been exposed to heat. Samples were prepared following the conven- desert animals including birds, hares and gazelle (Hakker-Orion, 1999: tional KBr method (Weiner, 2010), and analyzed between 4000 and − − 329–330). Macroscopic botanical remains from selected loci were re- 250/400 cm 1 at a 4 cm 1 resolution using either a Thermo Scientific ported; although contexts are unclear, all charred and uncharred spe- Nicolet 380 or a Nicolet iS5 spectrometer with Omnic 9.3 software. cies are local to the immediate environment and none are economic Spectra were compared to an extensive reference library (courtesy of plants (Baruch, 1999:7*–11*). Phytolith analysis of five sediment the Kimmel Center for Archaeological Science, Weizmann Institute of samples, each yielding only around 30 phytoliths, was inconclusive Science) and experimental data (Berna et al., 2007; Regev et al., 2010; (Rosen, 1999). Friesem et al., 2013; Forget et al., 2015). The analysis of lithic material from Ein Ziq was limited by older collection methodologies (Vardi et al., 2007: 103). Nevertheless, Vardi 3.2. X-Ray fluorescence (XRF) et al. (2007: 112–114) concluded that: 1) the lithic assemblage suggests ad hoc domestic activity, including production and utilization of a Archaeological (e.g., ash, building collapse, floors) and control se- variety of materials (organic, leather, bone, shell, etc.), and 2) the five diments (n = 63) from Ein Ziq were collected to evaluate the possibility ‘sickle blades’ are insufficient evidence for agricultural activity. High of metallurgical (copper) activities at the site. Elemental composition quantities of hammerstones and grinding stones (photographed though was retrieved using a Spectro-XEPOS Energy Dispersive X-Ray unpublished) led Haiman (1996) to suggest large-scale copper proces- Fluorescence unit (ED-XRF) at the Kimmel Center for Archaeological sing activities, (note that Ein Ziq is the central IBA site closest to the Science, Weizmann Institute of Science. Samples were ground and copper production center in Wadi Faynan; Fig. 1), although Vardi et al. homogenized to 425 μm through sieving, following analytical proce- (2007: 113) questioned this identification due to the lack of other dures for powder samples outlined by Eliyahu-Behar et al. (2012: 258). copper production items (e.g., crucibles, molds, etc.). A known soil standard (GSS-1, see Xie et al., 1985, 1989) was run with We conducted two short seasons of investigation at the site in 2014 each tray of 11 samples to calculate measurement error for each ele- and 2015 (Fig. 3). We excavated both structures and open areas (the ment (Appendix 3). latter were generally neglected in the previous excavation) and cleaned and sampled sections left from Cohen's excavations (Area A; area des- 3.3. Microremains: phytoliths, dung spherulites and ash pseudomorphs ignators are unchanged for easy comparison to the earlier excavation). Five 4 × 4 m squares in two new areas (J and K, Fig. 3) were opened in Phytoliths were extracted from bulk sediment samples following the the center of the site. In addition, three smaller 1–2 × 2 m probes were rapid extraction method of Katz et al. (2010). Phytoliths were counted opened across the site. A total of c. 75 m2 (~0.3% of the site's area) at 200× over 16 random fields of view using a petrographic microscope were excavated by us. Controls were collected from beyond the limits of (Nikon Eclipse 50i POL). In samples with multi-cell phytoliths, all in- the site, including the local Taqiya Formation, alluvial sediments from dividual phytoliths were counted. Phytolith morphologies were iden- the wadi south of the site, the surface deposits immediately east of the tified following standard literature (Twiss et al., 1969; Rapp and site and subsurface sediments in areas beyond the limit of the site. Mulholland, 1992; Albert and Weiner, 2001; Madella et al., 2005; Excavation and recording strategies were the same as those described Piperno, 2006) and a reference collection of Negev plants (at the La- above for Nahal Boqer 66. We collected a total of 214 sediment boratory for Sedimentary Archaeology, University of Haifa).

715 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

Morphological quantification was conducted by identifying at least 200 the living surface of most rooms (Fig. 4). Small pockets of greyish de- phytoliths with consistent morphologies at 400× under PPL (Albert posits (ash?) were identified on bedrock in L. 16/NB/3, L. 16/NB/4, and Weiner, 2001) for samples with concentrations higher than c. 2.0 and L. 16/NB/8. million/1 g of sediment. Phytolith morphotypes are described ac- cording to ICPN guidelines (Madella et al., 2005). 4.1.2. Artifacts Calcitic dung spherulites (Canti, 1997) and ash pseudomorphs Macroscopic artifacts were exceptionally scarce. Only scant frag- (Canti, 2003; Brochier and Thinon, 2003; Shahack-Gross and Ayalon, ments of animal bones from medium-sized mammals were identified 2013) were analyzed using the sodium polytungstate (SPT) procedure (Dunseth et al., 2017: Table 1). A few lithics (notably, no sickle blades) outlined in Gur-Arieh et al. (2013). All calcitic microremains were and small fragments of grinding stones (n = 3) were found. Three counted at 400× in 16 random fields of view, utilizing both plane unusual groundstone limestone artifacts (c. 8 × 5 × 5 cm, oval, sphe- polarized (PPL) and cross polarized light (XPL). rical and rhombohedral respectively) were found in Structure 16/NB/ All microremain concentrations are reported in millions per gram of 11. A small quantity of sherds (n = 37; diagnostic = 8) was found at the sediment. Sample descriptions can be found in Appendices 1–2. site, mainly in the Southern Complex (n = 31/37). The characteristic coarse fabric and evidence of one-sided sooting on nearly half of all 4. Results sherds (n = 14/37) suggest that most were related to domestic cooking activities, although only three rims from holemouth cooking vessels 4.1. Nahal Boqer 66 were identified. The remaining diagnostic sherds were indicative of handmade closed vessels, mainly bases of medium flat-bottomed sto- 4.1.1. Field observations rage jars (n = 3), the notch-decorated neck of one medium storage jar, Wall features at Nahal Boqer 66 are visible on the surface, so our and one teapot spout. However, almost all diagnostic material could be work revealed little architecture not apparent in the earlier excavation. limited to types common to the Negev throughout the late 3rd mil- We confirmed the presence of two phases (by stratigraphic division and lennium BCE; only the decorated neck sherd, found on the surface in the changes in building techniques) previously identified by Cohen (1999: Northern Complex (L. 16/NB/5), clearly dates to the IBA. In the pre- 60–61) in the Southern Complex (L. 16/NB/1 and L.16/NB/2, Fig. 2; vious excavation (Cohen, 1986, 1999), sherd counts are not reported, see also Cohen, 1999: 60, Fig. 24, though note substantial confusion in and ceramics from only two loci were published, limiting the possibility his text between described loci and the site plan). of comparison. However, the diagnostic assemblage ascribed to the EB All probes revealed c. 10–40 cm of sediment accumulation. All is limited to holemouth cooking vessels (see Cohen, 1986: Pl. 8: 15–17), profiles were composed of an upper loose yellow-brown sediment (c. while the IBA assemblages are dominated by necked and holemouth 10–15 cm thick) covering a more compact grey-brown sediment that storage jars (Southern Complex, L. 9 and L. 17, Cohen, 1986: Pl. 46: continued down to limestone bedrock, which appears to have served as 7–22). Similar assemblages dominated by holemouth vessels (cooking

Fig. 4. Selected field photos from Nahal Boqer 66. Red tags on excavation sections indicate sampled locations and numbers (see Appendix 2 for details). A) L. 16/NB/1, looking north. Note clear grey-brown sediment continuing under the wall. B) L. 16/NB/2, looking west; note con- tinuation of grey-brown material underlying the wall-stones to the south. C) Courtyard 16/NB/3 looking west; note fragmented bedrock surface and sloping stone collapse from wall to the north. D) Sediment accumulation in Room 16/ NB/4, looking east; note thin grey sediment overlying bedrock, sampled as NB-4.3–5. E) Corner of Room 16/NB/8, looking southeast; note the thin accumulation above bedrock left from the previous excavation. F) Room 16/NB/ 9, looking northwest. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this ar- ticle.)

716 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726 and storage) are reported at many third millennium BCE small sites in the Negev and Sinai, including the Camel Site (EB and IBA, Saidel, 2011:67–79, Fig. 5.1–5.3), Rekhes Nafha 396 (EB and IBA, Saidel, 2002a:51–54, Fig. 13); Sheikh Awad, Har Horesha, and Ramat Matred 3 (EB: Saidel, 2002b: 183–188) and others (e.g., EB and IBA: Kadesh Barnea and Western Negev Highland sites, Saidel and Haiman, 2014: 99–134, Fig. 4.30).

4.1.3. Mineralogical observations Mineralogical composition of the sediments (n = 63) shows that most of them are dominated by calcite, unheated clay (mixture of kaolinite and montmorillonite) and quartz. Gypsum was identified in a third of the samples (n = 21/63). Carbonated hydroxylapatite—a phosphate mineral—was identified in two floor contexts (L. 16/NB/4 and L. 16/NB/8). Minor changes in the clays were noted only in the floor context of L. 16/NB/4, suggesting exposure to relatively low heat (~400–500 °C, cf. Berna et al., 2007; Forget et al., 2015).

4.1.4. Dung spherulite concentrations Dung spherulite concentrations in control sediments ranged from 0.00–0.74 million/1 g of sediment. Dung spherulite concentrations in the archaeological sediments were almost universally elevated, ranging from 0.9–195 million/1 g of sediment (Fig. 5A). The highest con- centrations were found in the loci associated with the earlier phases (EBIb-EBII; L. 16/NB/1 and 16/NB/2) and in the grey (ash?) deposit in L. 16/NB/3, although samples from almost all contexts at the site showed relatively high concentrations (including the EBIII-IBA context at L.16/NB/8). Notably, no dung spherulites were identified in samples from L. 16/NB/11, excavated within a one-room structure located on a ridge southeast of the site. Within the main complex itself, lower con- centrations were noted in a room, L. 16/NB/4. Aeolian sediments from the top part of sediment profiles include dung spherulites, yet the concentrations are generally much lower than in most of the archae- ological sediments.

4.1.5. Phytolith concentrations and morphologies Phytolith concentrations ranged between 0.13 and 0.32 million/1 g of sediment in the controls taken from the area around the site (Fig. 5B). Phytolith concentrations are elevated in all archaeological samples relative to the controls, averaging c. 1.1 million/1 g of sedi- ment. The highest phytolith concentrations were found just above bedrock in the earlier phase (EBIb-EBII; L. 16/NB/1 and 16/NB/2), averaging 2.4 million/1 g of sediment, as well as the grey (ash?) deposit from the later phase (EBIII-IBA; L.16/NB/8), averaging c. 3.5 million/ 1 g of sediment—the highest phytolith concentration at the site. Samples with the highest phytolith concentrations from the two radiocarbon-dated contexts (L. 16/NB/2 and L. 16/NB/8) were ana- lyzed morphologically. Phytolith assemblages from these contexts are dominated by morphologies indicative of woody shrubs, making up 71% and 66% of the assemblage respectively (Fig. 6). Not a single dendritic phytolith, a morphotype abundant in the inflorescence of domestic cereals (Albert et al., 2008), was observed in any of the 46 analyzed samples. Phytoliths were well-preserved in both dated contexts, with be- tween 3 and 4% weathered morphologies. No melted or heat-affected phytoliths were noted in the assemblage. The samples showed 22 and 47% of phytoliths in anatomical connection (i.e. multicells), also sug- gesting good preservation. Preliminary palynological analysis of samples from L. 16/NB/2 showed a dominance of Chenopodiaceae pollen clusters, while samples from L. 16/NB/8 were dominated by clusters of Artemisia pollen (D. Langgut personal comm.), flora typical to the arid Negev environment. The pollen in both contexts showed evidence for being partially di- (caption on next page) gested (D. Langgut personal comm.), i.e., in accordance with its asso- ciation with dung spherulites.

717 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

Fig. 5. Boxplots of microremain results from Nahal Boqer 66, by context. A) identify as collapse and degradation of construction material (details Dung spherulite concentrations; B) phytolith concentrations; C) ash pseudo- below) and aeolian dust filled all structures excavated. Notably within morph concentrations. Sample size (n) is provided below labels. Boxplots this material were fiber impressions (probably of vegetal origin), as well constructed using IBM SPSS Statistics 24, following standard conventions: boxes as large gypsum crystals and hematite-pyrite concretions up to 5 cm represent interquartile (IQR) range (lower box: quartile 1, 25th percentile; solid long. All structures and open areas were sealed by c. 10–25 cm of line: median, 50th percentile; upper box: quartile 3, 75th percentile); whiskers yellow-brown aeolian sediment and stone collapse. indicate the maximum and minimum datapoints within 1.5× interquartile range; dots are suspected outliers, asterisks are extreme outliers. Note clear In Square L21, the remains of Structure 14/J/21 were partially presence of dung indicated by high dung spherulite concentrations in nearly all excavated. The structure had collapsed outward, and was preserved samples coupled with elevated phytolith concentrations (significantly above only to one course. Within the structure a beaten earth floor was un- controls). covered (Fig. 7B), and along the eastern section a sandstone grinding stone and a nearly complete medium storage jar were found next to a small hearth, sealed by grey-colored sediment including evidence for plant impressions (Fig. 8C). Outside the structure, a small hearth was identified, as were pockets of ash sealed by the collapse (L. 14/J/30). In the open area exposed in Square M21, a c. 10 cm thick homo- genous ashy deposit (L. 14/J/16) was uncovered (Figs. 7B, 8D). This material was rich in charcoal fragments (mainly twigs of local desert species; Table 1 in Dunseth et al., 2017). Small copper fragments, a few burnt sherds, animal bones and a chert hammerstone were found within the ashy material. The sediment below the deposit was completely sterile of archaeological remains. Above, the material was sealed by c. 20 cm accumulation of yellow-brown aeolian deposit (L. 14/J/10). In Squares M22 and M23, the remains of Structure 14/J/19 were preserved up to three courses, c. 60 cm high. Structure 14/J/19 was partitioned by a thin single-course dividing-wall built of vertical thin limestone slabs (Fig. 7B), somewhat similar to Cohen's L. 154 (Cohen, 1999: Figs. 94–95, see also Building 3c–d at Be'er Resisim, Dever, 2014: Fig. 11.9). The structure also contained a bin-like installation (L. 15/J/ 4) common in many structures previously excavated at Ein Ziq. Throughout the structure the floor could not be clearly defined. The structure was filled with the uniform grey-brown collapse, similar to that found in Structures 14/J/21 and 15/K/5. A probe (L. 15/J/7) was cut north of Wall 17/J/2 to provide a full stratigraphic sequence for sampling. To the east of Structure 14/J/19, Wall 15/J/2 was partially exposed (Fig. 7B). It is unclear if the remains belong to an open structure or Fig. 6. Results of phytolith morphological analysis from the dated dung accu- another room. Between Structure 14/J/19 and Wall 14/J/1, Pavement mulations in L. 16/NB/2 and L. 16/NB/8. Grass phytoliths comprise 20–30% of 15/J/5 was built of angular cobbles. The extent of the pavement was the assemblage while phytoliths from woody shrubs dominate the assemblage, not explored. Sealed below the westward collapse of Structure 14/J/19 fl re ecting livestock diet based on free-grazing in an arid environment. were a few ash pockets (L. 14/J/31). In Area K, Structure 15/K/5 was partially exposed (Fig. 7A). It was 4.1.6. Ash pseudomorph concentrations and PSR values filled with c. 50 cm of grey-brown sediment interpreted as collapse. Most locations containing ash pseudomorphs are in the Southern Below this relatively sterile collapse a small hearth was found in the Complex, and higher concentrations are found in accumulations di- section. Outside of the structure three small hearths were identified, rectly above bedrock, i.e., floors (Fig. 5C). However, the higher con- including stone-lined Hearth 15/K/8 (Fig. 8E) sealed by the collapse of centrations are those in the Northern Complex (L. 16/NB/3 and L. 16/ Structure 15/K/5. Charred twigs from the hearth were radiocarbon- NB/4), with the highest concentrations on the floor of L. 16/NB/4. dated (Dunseth et al., 2017: Table 1). Pseudomorph to spherulite ratio (PSR) values range from 0.001–0.778 No horizontal development (i.e. addition of abutting rooms, change (Appendix 3), mainly falling within the range of dung-rich ash (Gur- in structure use, etc.) was observed in the limited excavated areas. Arieh et al., 2014: Fig. 6). Three probes were dug across the site in open areas to 1) check preservation of microremains, and 2) explore areas where copper 4.2. Ein Ziq hoards had been previously found (Fig. 3). Probes 15/B/TP1 and 15/C/ TP2 showed homogenous sediment profiles and very few material re- 4.2.1. Field observations mains. Probe 15/F/TP3 in the lower terrace revealed an ash-rich de- In Area A, four sections were cleared in structures partially ex- posit c. 10 cm thick immediately next to Structure 73 (Fig. 8F). cavated by Cohen (Figs. 3, 7, 8). The structures in this area were pre- served to a height of between 20 and 40 cm. All structures were con- 4.2.2. Artifact assemblage from the new excavations structed by digging c. 5–10 cm into the alluvial terrace sediments, In our small excavations, a sizable number of sherds (n = 541; di- forming a ‘pit house’, as identified during the previous excavation agnostic = 58) and two complete vessels (one cup and one medium- (Cohen, 1999: 138–141). A small hearth was present in the south- sized storage jar) were uncovered. Other than a single Nabatean sherd eastern section of Structure 14/A/2 (Fig. 8A). Charcoal was also iden- from the surface, all pottery belongs to the IBA. The majority of sherds tified in the section of Structure 14/A/4 (Fig. 8B). were found outside of the structures (75%, n = 405/541). Due to the In the two new areas opened by us, three structures were exposed scarcity of rims, only a basic classification of the material has been (14/J/21, 14/J/19 and 15/K/5) (Fig. 7). Again, all structures, built made (see Fig. 9). Diagnostic sherds from closed storage vessels (am- from rounded limestone boulders, were dug c. 5–10 cm below the an- phoriskoi, storage jars, pithoi) dominate the assemblage (c. 64%). cient activity surface. A 30–60 cm grey-brown sediment of what we Thirty-one small sherds (6% of the total assemblage; diagnostic

718 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

Fig. 7. Detailed plan of new excavations at Ein Ziq. A) Area K and, B) Area J. Sampled sections and contexts are marked in red; findspots of copper scrap in green; full grinding stones in light blue; and sandstone grinding stone fragments in orange. Red letters correspond to sections presented in Fig. 8. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) n = 7) were red- or brown-slipped and/or burnished fragments of excavation belong to Family TR. The rest of the identifiable assemblage Dever's Family TR, associated with sites from Transjordan (Dever, 1980, at Ein Ziq can be associated with ceramic types common to the Negev 2014; for petrographic evidence, see Goren, 1996) and classified as and Sinai (known as Family S: Dever, 1971, 1980; Group A or Southern representing an early component of the IBA ceramic assemblage (Dever, Group in Amiran, 1969). 1980 contra D'Andrea, 2012). Diagnostic sherds of closed storage ves- Diagnostic sherds from cooking vessels were relatively rare (n = 9, sels make up the largest percentage of this very small assemblage 16%). Still, this is higher than the percentages reported by earlier ex- (n = 3/7), although sherds of open forms, including one lamp, one cavators at Ein Ziq (10% holemouth vessels, n = 18) and other central bowl and two cups, were also found. Two additional red-slipped and IBA sites (Cohen, 1999: 232–235); but note that the sample size is ra- burnished sherds from non-diagnostic open forms were identified. No- ther small. tably, nearly half of diagnostic open forms (4/9, 44%) found in our Not a single sickle blade was identified in our excavations. This

719 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

Fig. 8. Selected sampled locations at Ein Ziq. A) Hearth 14/A/2 looking west; note charred twigs. B) West section of L. 14/A/4 within a structure. The grey-green sediment is a mixture of Taqiya marl and aeolian dust. Round black samples are for OSL analysis (results appear in Dunseth et al., 2017). C) Hearth 15/J/14 inside a structure, looking east; note grey sediment, sealed by a grinding stone, covered by grey- green collapse. The grey-green collapse includes marl of the geological Taqiya Formation. D) Refuse deposit 14/J/16 looking north; note the sloping light grey ash-rich sediment sealed by yellow-brown aeolian dust. Carved features are for micromorphological analysis (results appear in Dunseth et al., 2017). The dotted white line represents limit of the locus. E) Stone-lined Hearth 15/K/8 in an open area, looking north; note dark grey sediment sealed by the collapse of Structure 15/K/5 above. F) The lower portion of a secondary ash deposit in 15/F/TP3; note grey appearance dotted by charcoal remains. The dotted white line represents the limit of the locus. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

follows a meager five found in Cohen's much larger exposure (Vardi grinding stone surfaces. et al., 2007: 110). The groundstone artifacts are mainly sandstone A few unidentifiable copper fragments and a copper pin were found grinding stones (see Fig. 7 for findspots). Analysis of sediments ad- in open spaces outside of the structures (see Fig. 7 for locations). hering to the grinding stone fragments found in Structure 14/J/19 noted very low quantities of blocky and polyhedral multi-cell dicoty- 4.2.3. Mineralogical and elemental results – ledonous phytoliths (0.1 0.5 millions/1 g of sediment), within the Mineralogical data from the samples was obtained though FTIR range of control sediments. Starch granules were absent from the analysis. Most sediments showed similar spectra, dominated by calcite,

Fig. 9. Analysis of the pottery assemblage from the new excavations at Ein Ziq. A) Diagnostic sherds and complete vessels; B) open vs. closed vessels in the whole as- semblage. Note the dominance of closed storage vessels (shades of blue) associated with trade in contrast to open forms (green). (For interpretation of the refer- ences to color in this figure legend, the reader is referred to the web version of this article.)

720 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726 unaltered clay (montmorillonite with traces of kaolinite) and quartz. Distinguishing secondary ash deposits from in situ hearths was aided by the identification of heat-altered clays. Micromorphology served as another line of evidence (data in Dunseth et al., 2017). Al- tered clays were identified only in two contexts, in rubified sediments below Hearth 15/J/14, and in one sample from Ash Deposit 14/J/16. All ash features showed pyrogenic calcite, the presence of gypsum and carbonated hydroxylapatite. The latter mineral also appears in the samples from the floor in Structure 14/J/21. Sediments interpreted macroscopically as collapse are composed of geogenic calcite, kaolinite and gypsum. Coupled with the identification of marl in micromorphological samples (Dunseth et al., 2017), as well as the large gypsum crystals and pyrite-hematite concretions, we sug- gest that these sediments are made from the marly Taqiya Formation (cf. Sass et al., 1965; Ben-Tor, 1966), which is exposed c. 200 m south of the site (see Avni and Weiler, 2013). To evaluate the possibility of metallurgical processing activities at the site (e.g., grinding of slag, hammering, etc. cf. Haiman, 1996), se- diment samples from ash, collapse and floor contexts were subjected to ED-XRF. The XRF results are presented in Appendix 4. Notably, neither copper nor other elements related to metallurgical activity (enrichment of Mn, Si, Fe in slags, ash and sediment; see, Hauptmann, 2007: 186–199; Shilstein et al., 2014) were enriched in any archaeological collapse, floor, surface or ash deposits relative to controls. Additionally, the elemental data shows enrichment of sulfur in ash- containing features, likely related to the presence of sulfur in some of the combusted plant species, e.g., Tamarix sp. (Shahack-Gross and Finkelstein, 2008).

4.2.4. Dung spherulite concentrations The dung spherulite concentrations are very low in all controls and archaeological samples from Ein Ziq and its vicinity. All sediment samples range from 0.0–0.35 million dung spherulites per g sediment (Fig. 10A). There is no microremain evidence for accumulations of dung at Ein Ziq.

4.2.5. Phytolith concentrations and morphologies The phytolith concentrations in the controls range from 0.0–0.3 million/1 g sediment. Phytolith concentrations in subsurface sediments, construction material from building collapse, and aeolian sediments are within the range of the controls. Only hearth and ash deposits have significantly elevated (by an order of magnitude) phytolith concentra- tions (Fig. 10B). Morphologically, the assemblages from both Hearths 15/J/14 and 15/K/5 were dominated by phytoliths indicative of woody plants (82% and 68% respectively), with a smaller component of types indicative of grasses (13% and 29% respectively) (see Fig. 11). The assemblage in Hearth 15/J/14 is very similar to the phytolith assemblage found in the nearby Ash Deposit 14/J/16 (woody shrubs: 77%; grasses: 20%). No- tably, dendritic phytoliths made up 3% of the assemblage in Hearth 15/ K/5. Phytolith concentrations of consistent morphologies were too low in Hearth 15/K/8 and the Ash Deposit 15/F/TP3 for a statistically significant analysis. Phytoliths were generally well-preserved in the hearth deposits, with c. 5–17% weathered morphologies. This is comparable to rela- tively well-preserved phytolith assemblages from the Iron Age site of Izbet Sartah (Cabanes et al., 2012: Fig. 7), and within the range of open ash features reported from Philistine pebble-hearths at Tell es-Safi/Gath (c. 5–10% weathered morphotypes, Gur-Arieh et al., 2014: Fig. 4). Fig. 10. Boxplots of microremain results from Ein Ziq, by context. A) Dung 4.2.6. Ash pseudomorph concentrations spherulite concentrations; B) phytolith concentrations; C) ash pseudomorph Enrichment in ash pseudomorph concentrations relative to controls concentrations. Statistical representation as in Fig. 5. Low levels of micro- was noted in two ash deposits (L. 15/F/TP3 and L.14/J/16) and three remains are represented in all contexts except for ash (refuse) and hearth fea- hearths (Hearths 15/J/14, 15/K/5 and 15/K/8), as expected from the tures (shown at same scale to emphasize the significant differences by orders of nature of these features (Fig. 10C). These are also features that have magnitude between the archaeological contexts); the extremely low dung — — significantly elevated values of phytoliths. spherulite signature below levels of controls in nearly all samples shows there is no evidence for animal penning in any excavated context at the site.

721 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

5.2. Ein Ziq

5.2.1. Construction techniques Our excavations clarify some construction techniques utilized at Ein Ziq. In his earlier excavations, Cohen already identified the practice of digging shallow pits into the alluvial terrace and lining it with field- stones. He reconstructed roofing techniques according to the discovery of pillars and radiating collapsed beams in a few of the rooms (Cohen, 1992, 1999: 277–279). This differs from the stone roofing slabs com- monly found at other central IBA sites in the region, including Mashabe Sade (Cohen, 1999: 117–130, Fig. 74) and Be'er Resisim (Dever, 2014: 161–163). In addition, Cohen and Dever (1981: 61) stated that they identified a mixture of mud, charcoal and lime used as weatherproofing at Be'er Resisim, although they present little information. We can add mineralogical and micromorphological evidence that the structures at Ein Ziq included marl from the local Taqiya Formation (this study and Dunseth et al., 2017). Use of marl from the Taqiya Formation—though not necessarily at Ein Ziq specifically—as a clay source for ceramics is attested petro- graphically from the Chalcolithic (Boness et al., 2016), EB (Porat, 1989) and IBA (Goren, 1996). However, this is the first time we can observe it Fig. 11. Analysis of phytolith morphologies from selected ash deposit and specifically used as a construction material, without being subject to hearth contexts at Ein Ziq. The phytolith signals in all three contexts suggest pyrotechnology. woody shrubs were the primary fuel source for all hearths.

5.2.2. Fuel sources in hearths 5. Discussion Phytolith assemblages from hearth contexts have been extensively investigated in Israel, Jordan and Syria in both prehistoric and historic In the arid Negev Highlands, phytoliths and calcitic microremains, sites (e.g., Albert et al., 2003, 2008; Jenkins et al., 2011; Portillo et al., namely dung spherulites and wood ash pseudomorph crystals, tend to 2014; Gur-Arieh et al., 2014) informing on the types of fuel used. be well preserved (Cabanes and Shahack-Gross, 2015; Shahack-Gross The phytolith morphological analysis of two hearths and an ashy and Finkelstein, 2015). Thus microremain concentrations and refuse deposit at Ein Ziq showed a clear dominance of woody dicots as a morphologies are expected to provide information on herding and fuel source. According to the dung spherulite concentrations, there is no agricultural practices, the location of animal enclosures and use of dung evidence for use or addition of animal dung as a fuel source in the for fuel. hearths or the ash deposit. Preliminary charcoal analysis suggests preference for local shrubs including Anabasis sp. and Retama raetam, all found within ash deposits 5.1. Nahal Boqer 66 or hearths (see Table 1 in Dunseth et al., 2017). The presence of gypsum in all hearth and ash deposits may suggest that Tamarix sp. was also a Dung spherulite and phytolith concentrations are elevated above common fuel source (cf. Shahack-Gross and Finkelstein, 2008)—an — controls in almost every context sampled unequivocal evidence that observation that goes against modern Bedouin practices, who avoid animal stabling took place at the site. Analysis of phytoliths in dung burning the plant due to its noxious smoke (Bailey and Danin, 1981: deposits showed no evidence for dendritic phytoliths that may be in- 146). dicative of cereal agriculture. Preliminary palynological analysis also did not identify cereal pollen grains. Both the phytolith and pollen 5.2.3. Uses of space analyses indicate that woody dicotyledonous species make up the ma- Based on the artifact assemblages, ash deposits, and numerous jority of the ancient animal diet throughout the history of Nahal Boqer hearths, activity appears to have been concentrated outside of struc- 66. This puts doubt on the assumed role of seasonal or opportunistic tures. We identified hearths and ashy refuse deposits in open areas. Ash agriculture in these and other small EB and IBA sites (e.g., Glueck, Deposit 14/J/16 is composed of a mixture of ash and heated and un- – 1939; Evenari et al., 1958; Kochavi, 1967: 253 254; Cohen and Dever, heated clay, which we reconstruct to be the result of dumping following 1981: 70; Cohen, 1999; Dever, 2014: 220). sweeping or cleaning activities from both outside and within the Lower concentration of dung spherulites in L.16/NB/4, together structures (see also the micromorphological evidence in Dunseth et al., with the presence of ash pseudomorphs and mineralogical evidence for 2017: Fig. 9). That activity remains are prevalent in open spaces at Ein phosphate and low-temperature heat-altered clays in the sediments Ziq accords with the earlier arguments of Dever (1985: 19–21), that above bedrock, provide non-ceramic evidence for domestic (perhaps most activity at Be'er Resisim, another central IBA site in the Negev cooking) activities at the site, possibly using dung as fuel. Exceptionally Highlands occurred in shared open areas. low phytolith and ash pseudomorph concentrations in conjunction with Internal divisions, hearths, grinding stones and complete vessels in the complete absence of dung spherulites suggest that animal stabling some structures in our and Cohen's excavations at Ein Ziq all suggest or use of dung for fuel or construction was not conducted in L. 16/NB/ that human activity was also conducted within the structures, contra 11. The unusual stone artifacts found in this locus may also hint at some Dever's (1985) assertion that structures at the central site of Be'er Re- ff sort of di erent (cultic?) activity versus the rest of Nahal Boqer 66. sisim were used primarily for sleeping. Given the dominance of storage Scarcity of macroscopic material in general is common at many small vessels at Ein Ziq, we propose that at least some of the round structures sites in arid regions (Banning and Köhler-Rollefson, 1992; Saidel and could be interpreted as storage structures. Erickson-Gini, 2014) and suggests either planned abandonment(s) or ff later scavenging activity (cf. LaMotta and Schi er, 1999). 5.2.4. Subsistence practices In comparison to Nahal Boqer, there is no evidence for stabling of animals at Ein Ziq. This echoes the same negative evidence from

722 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

Mashabe Sade (Dunseth et al., 2016), and portrays it as a phenomenon inside circular structures at large IBA sites has been missed in previous characteristic of IBA central settlements. excavations, which focused on the structures. The function of the cir- Although dendritic phytoliths at a relatively high percentage (3%) cular structures may not necessarily reflect strict domestic units, espe- were identified in one hearth context (Hearth 15/K/5), they are prob- cially in light of the lack of evidence for food production. On the other ably the remains of grass inflorescence used as tinder, as hearths were hand, the layout of the small IBA sites is different from that of the large not fueled by dung. The meager phytolith assemblages from all contexts sites, in the sense of the existence of large enclosures with an abun- show no sign for domestic cereals. This evidence—or more accurately, dance of dung deposits, pointing to nomadic pastoralism. the lack thereof—together with the good state of phytolith preservation Evidence for long-distance trade at both large and small sites in- suggests that food production was not practiced at Ein Ziq. This too is cludes ingot hoards and finished copper objects from Faynan (Segal similar to our finding at Mashabe Sade, hinting that this phenomenon is et al., 1996–1997; Hauptmann et al., 2015), marine shells from the Red also characteristic of central IBA sites. Sea (Bar-Yosef, 1999; Saidel, 2002a; Dunseth et al., 2016), and the The elemental and mineralogical analysis of sediments in Areas J dominance of (non-local) storage vessels in the ceramic assemblages and K revealed no evidence for copper or slag processing activities. (Goren, 1996; Cohen, 1999; Saidel, 2002b; Gidding, 2016 and this Similar negative results were reported from one ash dump context at paper). For the large sites, evidence for long-distance trade, the com- Mashabe Sade (Dunseth et al., 2016: Table 3). Copper items—including plete absence of macro- or microarchaeological signals for pastoral ingots, finished weapons, tools and scrap—were moved through Ein Ziq activity or food production, as well as their location in defensible lo- and other Negev central sites (Be'er Resisim, Cohen, 1999: 200–224, calities or near water sources (Haiman, 1996), can best be understood Dever, 2014: 166–168, Fig. 11.64: 4–5; Mashabe Sade, Cohen, 1999: as pointing to their function as trading hubs with storage facilities (see 118, Dunseth et al., 2016: 55; Har Zayyad, Cohen, 1999:96–98; Har already Finkelstein, 1989: 134). If so, they may have been populated Yeruham, Kochavi, 1967: 108–118), as well as some of the smaller sites selectively (by gender, age, social status), rather than by whole families, (Rekhes Nafha 396, Saidel, 2002a; Rogem Be'erotayim, Saidel et al., an observation supported by the composition of the ceramic assem- 2006). Active IBA copper processing is indeed evident at Arabah copper blage. The radiocarbon data from Ein Ziq spans some 300 years in nine production sites (Hamra Ifdan, Adams, 2000; Levy et al., 2002; overlapping determinations (Dunseth et al., 2017). We propose that the Hauptmann et al., 2015; Gidding, 2016; Ben-Yosef et al., 2016;En site is not entirely contemporaneous, and clusters of structures are Yahav, Yekutieli et al., 2005; Vardi et al., 2008; possibly in Timna, possibly incorporating diachronic development or ‘horizontal strati- Rothenberg and Shaw, 1990). However, the data does not support graphy’, likely associated with the addition of clans/families into the processing of material at central sites, as argued by Haiman (1996: trade network. The population in the large sites could have been—at 18–22). least partially—external to the Negev Highlands. At small sites, dung remains and the ceramic assemblage composi- 5.3. An updated model for the IBA in the Negev Highlands tion (dominated by cooking vessels) indicate primarily domestic ac- tivity, probably by whole families. Moreover, the longevity of these The prevalent model for the IBA in the Negev Highlands argues for sites throughout the EB and IBA (Dunseth et al., 2017) indicates that two populations, one at small sites subsisting on herding and oppor- nomadic pastoralists was practiced in the area for many centuries, tunistic cereal cultivation, and another dwelling in large sites, engaged meaning that we are dealing with an indigenous desert population. in copper processing and trade, whose direct subsistence base in unclear Although the material culture (e.g., ceramics, lithics and copper) at (Haiman, 1996). Two major deviations from this model emerge from small and large sites is similar and recent radiocarbon data (Dunseth the research presented here: et al., 2017) shows that in the IBA the sites were inhabited con- temporaneously, it is still unclear whether (or how) the large and small 1) There is no evidence for opportunistic cereal cultivation at small settlements interacted with each other. Assuming that small sites were sites (the pattern reported here from Nahal Boqer 66 is repeated in inhabited by pastoral desert groups and that the large sites were visited another small site we recently excavated—west of Mitzpe Ramon, by (local or non-local) traders en route, the presence of copper items and unpublished data). This is supported by the lithic evidence from Red Sea shells at small sites may reflect gift giving and/or exchange other excavated sites (Saidel, 2002a; Saidel et al., 2006; Rosen, relations between these populations. Furthermore, we note that 2011b; Vardi, 2014). zooarchaeological evidence, though scarce, indicates consumption of 2) There is no evidence for copper processing (contra Haiman, 1996)or meat-rich portions at large sites. It is therefore possible that pastoral food production (in support of Haiman, 1996, but contra Cohen, nomadic groups living at small sites supplied inhabitants of large sites 1992, 1999 and Dever, 1985, 2014) at large sites (both Ein Ziq and with meat, milk and other herd products (already Haiman, 1996: Mashabe Sade, Dunseth et al., 2016). 24–25). Turning to the broader picture, work at the site of Hamra Ifdan In addition, based on radiocarbon results, we have recently shown (Adams, 2000; Levy et al., 2002; Gidding, 2016; Ben-Yosef et al., 2016) that: and copper items found at (mainly large) sites, show that third mil- lennium BCE activity in the Negev was strongly associated with copper A) Nomadic pastoralism was practiced at Nahal Boqer 66 throughout production and transportation. In an earlier phase of this network, the late third millennium BCE (Dunseth et al., 2017). during the EBII–III, copper production was arguably controlled by B) IBA activity at both small and large sites is confined to the first part gateway communities such as Arad (Ilan and Sebbane, 1989; Gidding, of the period (c. 2500–2200 BCE) (Dunseth et al., 2017). 2016) and Bab edh-Dra', which administered the distribution of copper towards destinations in Egypt and northern Canaan. The collapse of With all the new information at hand, from two large and two small urban Canaan at the end of the EB III c. 2500 BCE brought about a sites, we can now propose an alternative model to explain the IBA change. During the ensuing IBA copper production in the Arabah and settlement peak in the Negev Highlands. transportation of copper were operated by desert groups, which pre- The structure of the large IBA sites is characteristic of clan-based ferred using 'inner' desert routes to Egypt (cf. Haiman, 1996). The nomadic groups, which are typically composed of clusters of small production and transportation probably intensified because of growing habitation units surrounded by open spaces where most activity takes demand during the peak period of the Old Kingdom—the time of the place (Shahack-Gross, 2002 and references therein on pastoral nomads; 5th and 6th Dynasties. Given the nature of the finds at Negev and Sinai Friesem and Lavi, 2017 and references therein on hunter-gatherers). sites (defined by some as the ‘Timnian’, cf. Rosen, 2011a), and the The abundance of activity remains in open spaces in comparison to scarcity of Egyptian material at these sites (cf. Cohen, 1999; also Goren,

723 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

1996), it is clear that the system was not commanded by Egypt directly, Babenko, A.N., Kiseleva, N.K., Plahkt, I., Rosen, S., Savinetskii, A.B., Khasanov, B.F., but arguably was instead controlled by the indigenous desert popula- 2007. Reconstruction of the Holocene vegetation in the central Negev Desert, Israel, on the basis of palynological data on the Atzmaut zoogenic deposit. Russ. J. Ecol. 38, tion. 417–426. As demonstrated in Dunseth et al. (2017), IBA activity in the Negev Bailey, C., Danin, A., 1981. Bedouin plant utilization in Sinai and the Negev. Econ. Bot. covers only the first half of the period, c. 2500–2200 BCE. It is rea- 35, 145–162. Banning, E., Köhler-Rollefson, I., 1992. Ethnographic lessons for the pastoral past: camp sonable to propose that prosperity in the south came to an end with the locations and material remains near Beidha, Southern Jordan. In: Bar-Yosef, O., decline of Old Kingdom Egypt c. 2200 BCE and the resulting decay of Khazanov, A. (Eds.), Pastoralism in the Levant: Archaeological Materials in demand for Arabah copper (Ben-Yosef et al., 2016: 80; Dunseth et al., Anthropological Perspective. Monographs in World Archaeology 10. pp. 181–204 2016). Madison. Baruch, U., 1999. Identification of wood remains from Horbat 'En Ziq. In: Cohen, R. (Ed.), Ancient Settlement of the Central Negev, Vol. I: The Chalcolithic, the Early Bronze 6. Conclusion Age and the Middle Bronze Age I pp. 7–11 Jerusalem. Bar-Yosef, D., 1999. Shells from three Middle Bronze Age I sites in the Negev Highlands. In: Cohen, R. (Ed.), Ancient Settlements of the Central Negev, Vol. I: The Chalcolithic The high-resolution dating of Nahal Boqer 66 and Ein Ziq (Dunseth Period, the Early Bronze Age and the Middle Bronze Age I pp. 322–326 Jerusalem. et al., 2017) and the microarchaeological evidence for Nahal Boqer 66, Ben-Tor, Y.K., 1966. The Clays of Israel. Jerusalem. Ein Ziq (this article) and Mashabe Sade (Dunseth et al., 2016), make it Ben-Yosef, E., Gidding, A., Tauxe, L., Davidovich, U., Najjar, M., Levy, T.E., 2016. Early fl ff Bronze Age copper production systems in northern Arabah Valley: new insights from possible to propose that the two IBA site types re ect di erent sub- archaeomagnetic study of slag deposits in Jordan and Israel. J. Archaeol. Sci. 72, sistence strategies. The first—small sites characterized by a set of rooms 71–84. surrounding open enclosures (in our work Nahal Boqer 66 and the yet Berna, F., Behar, A., Shahack-Gross, R., Berg, J., Boaretto, E., Gilboa, A., Sharon, I., — Shalev, S., Shilstein, S., Yahalom-Mack, N., Zorn, J.R., Weiner, S., 2007. Sediments unpublished Nahal Nizzana 332) represents desert nomadic pastor- exposed to high temperatures: reconstructing pyrotechnological processes in Late alism; they had already been active in the Early Bronze Age. The second Bronze and Iron Age strata at Tel Dor (Israel). J. Archaeol. Sci. 34, 358–373. site-type—the central IBA sites (in our project Mashabe Sade and Ein Boness, D., Scheftelowitz, N., Fabian, P., Gilead, I., Goren, Y., 2016. Petrographic study of the pottery assemblages from Horvat Qarqar South, a Ghassulian Chalcolithic cem- Ziq)—shows no connection to food production or metallurgical activ- etery in the Southern Levant. Bull. Am. Sch. Orient. Res. (375), 185–213. ities and should probably be understood as trading hubs. The Negev Brochier, J.E., Thinon, M., 2003. Calcite crystals, starch grains aggregates or…POCC? network was closely related to the copper industry in the Arabah. IBA Comment on ‘calcite crystals inside archaeological plant tissues’ J. Archaeol. Sci. 30, – prosperity in the Negev Highlands was fueled by growing demand for 1211 1214. Brochier, J.E., Villa, P., Giacomarra, M., 1992. Shepherds and sediments: geo-eth- copper during the time of the 5th and 6th Dynasties in Egypt. The IBA noarchaeology of pastoral sites. J. Anthropol. Archaeol. 11, 47–102. settlement system decayed with the decline of the Old Kingdom in Cabanes, D., Shahack-Gross, R., 2015. Understanding fossil phytolith preservation: the Egypt. role of partial dissolution in paleoecology and archaeology. PLoS ONE 10, e0125532. http://dx.doi.org/10.1371/journal.pone.0125532. Appendices and supplementary data to this article can be found Cabanes, D., Gadot, Y., Cabanes, M., Finkelstein, I., Weiner, S., Shahack-Gross, R., 2012. online at https://doi.org/10.1016/j.jasrep.2018.03.025. Habitation around settlement sites: a phytolith and mineralogical study for assessing boundaries, phytolith preservation, and implications for spatial reconstructions using plant remains. J. Archaeol. Sci. 39, 2697–2705. Acknowledgements Canti, M.G., 1997. An investigation into microscopic calcareous spherulites from herbi- vore dungs. J. Archaeol. Sci. 23, 219–231. This work was supported by the German-Israeli Foundation for Canti, M.G., 2003. Aspects of the chemical and microscopic characteristics of plant ashes – fi found in archaeological soils. Catena 54, 339 361. Scienti c Research and Development (GIF; Grant No. I-1244-107.4/ Cohen, R., 1985. Archaeological Survey of Israel – Map of Sede Boqer-West (167). 2014) to R.S-G. and Markus Fuchs (Justus-Liebig-University Gießen) as Jerusalem. principle investigators and I.F. as co-investigator. We thank all volun- Cohen, R., 1986. The Settlement of the Central Negev in the Light of Archaeological and Literary Sources During the 4th–1st Millennia B.C.E. (PhD. dissertation). Hebrew teers that helped in the excavation of the two sites. Special thanks to the University of Jerusalem, Jerusalem (Hebrew with English summary). area supervisors at Ein Ziq, Erin Hall and Adam Kaplan; to Maayan Mor, Cohen, R., 1992. The nomadic or semi-nomadic Middle Bronze Age I settlements in the Sivan Einhorn, Paula Waiman-Barak, and Alon Shavit and Boaz Gross of Central Negev. In: Bar-Yosef, O., Khazanov, A. (Eds.), Pastoralism in the Levant: the Israeli Institute of Archaeology for logistical support; to Shira Gur- Archaeological Materials in Anthropological Perspectives. Monographs in World Archaeology 10. pp. 105–131 Madison. Arieh, David Friesem, Lior Regev, Yotam Asscher, Michael Toffolo and Cohen, R., 1999. Ancient Settlement of the Central Negev, Vol. 1: The Chalcolithic Period, Adi Eliyahu-Behar for sharing their expertise; and to Steve Weiner and the Early Bronze Age and the Middle Bronze Age I (Israel Antiquities Authority the Kimmel Center for Archaeological Science, Weizmann Institute for Reports, No. 6). Jerusalem (Hebrew with English summary). Cohen, R., Dever, W.G., 1980. Preliminary report of the second season of the “Central access to labs and equipment in the early phase of this research Negev Highlands Project” Bull. Am. Sch. Orient. Res. 232, 29–45. (2014–2015). Cohen, R., Dever, W.G., 1981. Preliminary report of the third and final season of the “Central Negev Highlands Project” Bull. Am. Sch. Orient. Res. (243), 55–77. D'Andrea, M., 2012. The Early Bronze IV period in South-Central Transjordan: re- References considering chronology through ceramic technology. Levant 44, 17–50. de Miroschedji, P., 2009. Rise and collapse in the Southern Levant in the Early Bronze Adams, R.B., 2000. The Early Bronze Age III–IV transition in Southern Jordan: evidence Age. In: Cardarelli, A., Cazzella, A., Frangipane, M., Peroni, R. (Eds.), Reasons for from Khirbet Hamra Ifdan. In: Philip, G., Baird, D. (Eds.), Ceramics and Change in the Change: Birth, Decline and Collapse of Societies From the End of the Fourth to the – Early Bronze Age of the Southern Levant, pp. 397–401 Sheffield. Beginning of the First Millennium B.C. pp. 101 129 Rome. Albert, R.M., Weiner, S., 2001. Study of phytoliths in prehistoric ash layers from Kebara Dever, W.G., 1971. The peoples of Palestine in the Middle Bronze I period. Harv. Theol. – and Tabun Caves using a quantitative approach. In: Meunier, J.D., Colins, F. (Eds.), Rev. 64, 197 226. “ ” – Phytoliths: Applications in Earth Sciences and Human History, pp. 251–266 Lisse. Dever, W.G., 1980. New vistas on the EB IV ( MB I ) horizon in Syria Palestine. Bull. Am. – Albert, R.M., Bar-Yosef, O., Meignen, L., Weiner, S., 2003. Quantitative phytolith study of Sch. Orient. Res. (237), 35 64. – hearths from Natufian and Middle Palaeolithic levels of Hayonim Cave (Galilee, Dever, W.G., 1985. Village planning at Be'er Resism and socio economic structure in – Israel). J. Archaeol. Sci. 30, 461–480. Early Bronze Age IV Palestine. Eretz Israel 18, 18* 28*. Albert, R.M., Shahack-Gross, R., Cabanes, D., Gilboa, A., Lev-Yadun, S., Portillo, M., Dever, W.G., 1989. The collapse of the urban Early Bronze Age in Palestine: towards a Sharon, I., Boaretto, E., Weiner, S., 2008. Phytolith-rich layers from the Late Bronze systemic analysis. In: de Miroschedji, P. (Ed.), L'urbanisation de la Palestine à l'âge du – and Iron Ages at Tel Dor (Israel): mode of formation and archaeological significance. Bronze ancien. British Archaeological Reports International Series 527. pp. 225 246 J. Archaeol. Sci. 35, 57–75. Oxford. Amiran, R., 1969. Ancient Pottery of the Holy Land. Ramat Gan. Dever, W.G., 2014. Excavations at the Early Bronze Age IV sites of Jebel Qa'aqir and Be'er Anderson, P.C., 1980. A testimony of prehistoric tasks: diagnostic residues on stone tool Resisim. Eisenbrauns, Winona Lake. working edges. World Archaeol. (2), 181–194. Dunseth, Z.C., Junge, A., Fuchs, M., Finkelstein, I., Shahack-Gross, R., 2016. Anderson, P.C., 1999. Experimental cultivation, harvest, and threshing of wild cereals. In: Geoarchaeological investigation in the Intermediate Bronze Age site of Mashabe – Anderson, P.C. (Ed.), Prehistory of Agriculture: New Experimental and Ethnographic Sade, the Negev Highlands. Tel Aviv 43, 43 75. Approaches. Monograph of the Institute of Archaeology, University of California, Los Dunseth, Z.C., Junge, A., Lomax, J., Boaretto, E., Finkelstein, I., Fuchs, M., Shahack-Gross, Angeles 40. pp. 118–144 Los Angeles. R., 2017. Dating archaeological sites in an arid environment: a multi-method case – Avni, Y., Weiler, N., 2013. Geological Map of Israel: Sede Boqer (Sheet 18-IV) (1:50,000). study in the Negev Highlands, Israel. J. Arid Environ. 144, 156 169. (Jerusalem). Eliyahu-Behar, A., Yahalom-Mack, N., Shilstein, S., Zukerman, A., Shafer-Elliott, C.,

724 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

Maeir, A.M., Boaretto, E., Finkelstein, I., Weiner, S., 2012. Iron and bronze produc- Paleoecologists. Lanham. tion in Iron Age IIA Philistia: new evidence from Tell es-Safi/Gath, Israel. J. Archaeol. Porat, N., 1989. Composition of Pottery—Application to the Study of the Interrelations Sci. 39, 255–267. Between Canaan and Egypt During the 3rd Millennium B.C. (PhD. dissertation). Erickson-Gini, T., 2010. Nabatean Settlement and Self-Organized Economy in the Central Hebrew University of Jerusalem, Jerusalem. Negev: Crisis and Renewal. Oxford. Portillo, M., Kadowski, S., Nishiaki, Y., Albert, R.M., 2014. Early Neolithic household Esse, D.L., 1991. Subsistence, Trade and Social Change in Bronze Age Palestine (Studies in behavior at Tell Seker al-Aheimar (Upper Khabur, Syria): a comparison to eth- Ancient Oriental Civilizations 50). In: 50 Chicago. noarchaeological study of phytoliths and dung spherulites. J. Archaeol. Sci. 42, Evenari, M., Aharoni, Y., Shanan, L., Tadmor, N.H., 1958. The ancient desert agriculture 107–118. of the Negev: III. Early beginnings. Isr. Explor. J. 8, 231–268. Prins, A.B., 2016. 3D modelling for archaeological documentation: using the JVRP Finkelstein, I., 1989. Further observations on the socio–demographic structure of the method to record archaeological excavations with millimeter-accuracy (version 3.0). Intermediate Bronze Age. Levant 21, 129–140. In: JVRP White Papers in Archaeological Technology, . www. Finkelstein, I., 1995. Living on the Fringe: the Archaeology and History of the Negev, jezreelvalleyregionalproject.com/3d-modeling.html. Sinai and Neighboring Regions during the Early Bronze and Iron Ages. Sheffield Rapp, G., Mulholland, S. Eds, 1992. Phytolith Systematics: Emerging Issues. New York. Academic Press, Sheffield. Reddy, S.N., 1999. Fueling the hearths in India: the role of dung in plaeoethnobotanical Finkelstein, I., 2014. The southern steppe of the Levant ca. 1050–750 BCE: a framework interpretation. Paléorient 24, 61–70. for a territorial history. Palest. Explor. Q. 146, 89–104. Regev, L., Poduska, K.M., Addadi, L., Weiner, S., Boaretto, E., 2010. Distinguishing be- Forget, M.C.L., Regev, L., Friesem, D.E., Shahack-Gross, R., 2015. Physical and miner- tween calcites formed by different mechanisms using infrared spectrometry: ar- alogical properties of experimentally heated chaff-tempered mud bricks: implications chaeological applications. J. Archaeol. Sci. 37, 3022–3029. for reconstruction of environmental factors influencing the appearance of mud bricks Regev, J., de Miroschedji, P., Greenberg, R., Braun, E., Greenhut, Z., Boaretto, E., 2012. in archaeological conflagration events. J. Archaeol. Sci. Rep. 2, 80–93. Chronology of the Early Bronze Age in the Southern Levant: new analysis for a high Friesem, D.E., Lavi, N., 2017. Foragers, tropical forests and the formation of archae- chronology. Radiocarbon 54, 525–566. ological evidences: an ethnoarchaeological view from South India. Quat. Int. 448, Rosen, A.M., 1995. The social response to environmental change in Early Bronze Age 117–128. Canaan. J. Anthropol. Archaeol. 14, 26–44. Friesem, D.E., Tsartsidou, G., Karkanas, P., Shahack-Gross, R., 2013. Where are the roofs? Rosen, A.M., 1999. Past environments and farming potential at Horbat 'En Ziq. In: Cohen, A geo-ethnoarchaeological study of mud brick structures and their collapse processes, R. (Ed.), Ancient Settlements of the Central Negev, Vol. I: The Chalcolithic Period, the focusing on the identification of roofs. Archaeol. Anthropol. Sci. 6, 73–92. Early Bronze Age and the Middle Bronze Age I. pp. 17–21 Jerusalem. Gidding, A., 2016. Archaeological Approaches to Commodity Production and Rosen, S.A., 1987. Demographic trends in the Negev Highlands: preliminary results from Distribution: An Example From the Early Bronze Age of Jordan (PhD. dissertation). the emergency survey. Bull. Am. Sch. Orient. Res. (266), 45–58. University of California San Diego, San Diego. Rosen, S.A., 2011a. The desert and the pastoralist: an archaeological perspective on Glueck, N., 1939. Explorations in Eastern Palestine III. American Schools of Oriental human-landscape interaction in the Negev over the millennia. Ann. Arid Zone 50, Research, Cambridge, Massachusetts. 1–15. Goren, Y., 1996. The Southern Levant in the Early Bronze Age IV: the petrographic Rosen, S.A., 2011b. An Investigation into Early Desert Pastoralism: Excavations at the perspective. Bull. Am. Sch. Orient. Res. (303), 33–72. Camel Site, Negev. Cotsen Institute of Archaeology, Los Angeles. Greenberg, R., 2017. No collapse: transmutations of Early Bronze Age urbanism in the Rosen, S.A., 2016. Revolutions in the Desert: The Rise of Mobile Pastoralism in the Southern Levant. In: Höflmayer, F. (Ed.), The Late Third Millennium in the Ancient Southern Levant. London. Near East: Chronology, C14, and Climate Change. Oriental Institute Seminars 11. pp. Rosen, S.A., Vardi, J., 2014a. Appendix IIH: the chipped stone assemblage from Be'er 31–58 Chicago. Resisim: A final report. In: Dever, W.G. (Ed.), Excavations at the Early Bronze Age IV Gur-Arieh, S., Mintz, E., Boaretto, E., Shahack-Gross, R., 2013. An ethnoarchaeological sites of Jebel Qa'aqir and Be'er Resisim, pp. 327–338 Winona Lake. study of cooking installations in rural Uzbekistan: development of a new method for Rosen, S.A., Shugar, A., Vardi, J., 2014b. Function and value in sickle segment analysis: identification of fuel sources. J. Archaeol. Sci. 40, 4331–4347. Odellian perspectives. In: Shott, M.J. (Ed.), Works in Stone: Contemporary Gur-Arieh, S., Shahack-Gross, R., Maeir, A.M., Lehmann, G., Hitchcock, L.A., Boaretto, E., Perspectives on Lithic Analysis, pp. 116–130 Salt Lake City. 2014. The taphonomy and preservation of wood and dung ashes found in archae- Rothenberg, B., Shaw, C.T., 1990. The discovery of a copper mine and smelter from the ological cooking installations: case studies from Iron Age Israel. J. Archaeol. Sci. 46, end of the Early Bronze Age (EBIV) in the Timna Valley. Institute of Archaeo- 25–45. Metallurgical Studies 15–16, 1–8. Haiman, M., 1996. Early Bronze Age IV settlement pattern of the Negev and Sinai Deserts: Saidel, B.A., 2002a. The excavations at Rekhes Nafha 396 in the Negev Highlands, Israel. view from small marginal temporary sites. Bull. Am. Sch. Orient. Res. (303), 1–32. Bull. Am. Sch. Orient. Res. (325), 37–63. Hakker-Orion, D., 1999. Faunal remains from Middle Bronze Age I sites in the Negev Saidel, B.A., 2002b. Pot luck? Variation and function in the ceramic assemblages of pre- Highlands. In: Cohen, R. (Ed.), Ancient Settlements of the Central Negev, Vol. I: The camel pastoralists in the Negev Highlands, Israel. Mitekufat Haeven 32, 175–196. Chalcolithic Period, the Early Bronze Age and the Middle Bronze Age I. pp. 327–335 Saidel, B.A., 2011. The pottery from the Camel Site. In: Rosen, S.A. (Ed.), An Investigation Jerusalem. into Early Desert Pastoralism: Excavations at the Camel Site, Negev. Cotsen Institute Hauptmann, A., 2007. The Archaeometallurgy of Copper. Berlin. of Archaeology, Los Angeles, pp. 67–79. Hauptmann, A., Schmitt-Strecker, S., Levy, T.E., Begemann, F., 2015. On Early Bronze Saidel, B.A., Erickson-Gini, T., 2014. A note on the excavation of an Ottoman and British Age copper bar ingots from the Southern Levant. Bull. Am. Sch. Orient. Res. (373), Mandate period Bedouin campground at Nahal Be'erotayim West in the Negev desert, 1–24. Israel. Arab. Archaeol. Epigr. 25, 138–145. Ilan, O., Sebbane, M., 1989. Copper metallurgy, trade and the urbanization of southern Saidel, B.A., Haiman, M., 2014. The ceramic assemblages. In: Saidel, B.A., Haiman, M. Canaan in the Chalcolithic and Early Bronze Age. In: de Miroschedji, P. (Ed.), (Eds.), Excavations in the Western Negev Highlands. British Archaeological Reports L'urbanisation de la Palestine à l'âge du Bronze ancien. British Archaeological Reports International Series 2684. pp. 99–136 Oxford. International Series 527. pp. 139–159 Oxford. Saidel, B.A., Erickson-Gini, T., Vardi, J., Rosen, S.A., Maher, E., Greenfield, H., 2006. Test Jenkins, E., Baker, A., Elliot, S., 2011. Past plant use in Jordan as revealed by archae- excavations at Rogem Be'erotayim in western Negev. Mitekufat Haeven 36, 201–229. ological and ethnoarchaeological phytolith signatures. In: Mithen, S., Black, E. (Eds.), Sass, E., Nathan, Y., Nissenbaum, A., 1965. Mineralogy of certain pyrite concretions from Water, Life and Civilization: Climate, Environment and Society in the Jordan Valley, Israel and their alteration products. Mineral. Mag. 35, 84–87. pp. 381–400 Cambridge, UK. Segal, D., Carmi, I., 1996. Rehovot radiocarbon date list V. 'Atiqot 29, 79–106. Junge, A., Lomax, J., Shahack-Gross, R., Dunseth, Z.C., Finkelstein, I., Fuchs, M., 2016. Segal, I., Roman, I., Cohen, R., Brenner, I.B., 1996–1997. Chemical and metallurgical OSL age determination of archaeological structures using trapped aeolian sediments: study of 'Ein Ziq and Be'er Resisim ingots. Arx 2–3, 43–51. a case study from the Negev Highlands, Israel. Geoarchaeology 31, 550–563. Shahack-Gross, R., 2002. Taphonomy of Abandoned Maasai Settlements in Southern Katz, O., Cabanes, D., Weiner, S., Maeir, A.M., Boaretto, E., Shahack-Gross, R., 2010. Kenya With Special Reference to Livestock Enclosures: Implications for the Spread of Rapid phytolith extraction for analysis of phytolith concentrations and assemblages Pastoralism in East Africa (PhD. dissertation). Washington University, St. Louis, during an excavation: an application at Tell es-Safi/Gath, Israel. J. Archaeol. Sci. 37, Missouri. 1557–1563. Shahack-Gross, R., 2011. Herbivorous livestock dung: formation, taphonomy, methods Kochavi, M., 1967. The Settlement of the Negev in the Middle Bronze (Canaanite) I Age for identification and archaeological significance. J. Archaeol. Sci. 38, 205–218. (PhD. dissertation). Hebrew University of Jerusalem, Jerusalem (Hebrew with Shahack-Gross, R., 2017. Animal gathering enclosures. In: Nicosia, C., Stoops, G. (Eds.), English summary). Archaeological Soil and Sediment Micromorphology, pp. 265–280 Hoboken. LaMotta, V., Schiffer, M.B., 1999. Formation processes of house floor assemblages. In: Shahack-Gross, R., Ayalon, A., 2013. Stable carbon and oxygen isotopic compositions of Allison, P.M. (Ed.), Archaeology of Household Activities, pp. 19–29 London. wood ash: an experimental study with archaeological implications. J. Archaeol. Sci. Levy, T.E., Adams, R.B., Hauptmann, A., Prange, M., Schmitt-Strecker, S., Najjar, M., 40, 570–578. 2002. Early Bronze Age metallurgy: a newly discovered copper manufactory in Shahack-Gross, R., Finkelstein, I., 2008. Subsistence practices in an arid environment: a southern Jordan. Antiquity 75, 425–437. geoarchaeological investigation in an Iron Age site, the Negev Highlands, Israel. J. Madella, M., Alexandre, A., Ball, T., 2005. International code for phytolith nomenclature Archaeol. Sci. 35, 965–982. 1.0. Ann. Bot. 96, 253–260. Shahack-Gross, R., Finkelstein, I., 2015. Settlement oscillations in the Negev Highlands Magness, J., 2003. The Archaeology of the Early Islamic Settlement in Palestine. Winona revisited: the impact of microarchaeological methods. Radiocarbon 57, 253–264. Lake. Shahack-Gross, R., Finkelstein, I., 2017. Iron Age agriculture in the Negev Highlands? Noy, T., Cohen, R., 1974. Nahal Boqer: an early Pre-Pottery Neolithic B site. Mitekufat Methodological and factual comments on Bruins and van der Plicht 2017a Haeven 13, 15–25 (Hebrew). (Radiocarbon Vol. 59, NR. 1). Radiocarbon 59, 1227–1231. Palumbo, G., 2001. The Early Bronze Age IV. In: MacDonald, B., Adams, R.B., Bienkowski, Shahack-Gross, R., Marshall, F., Weiner, S., 2003. Geo-ethnoarchaeology of pastoral sites: P. (Eds.), The Archaeology of Jordan, pp. 233–269 Sheffield. the identification of livestock enclosures in abandoned Maasai settlements. J. Piperno, D., 2006. Phytoliths: A Comprehensive Guide for Archaeologists and Archaeol. Sci. 30, 439–459.

725 Z.C. Dunseth et al. Journal of Archaeological Science: Reports 19 (2018) 712–726

Shahack-Gross, R., Boaretto, E., Cabanes, D., Katz, O., Finkelstein, I., 2014. Subsistence Agricultural Societies. British Archaeological Reports International Series 1657. pp. economy in the Negev Highlands: the Iron Age and the Byzantine/Early Islamic 99–116 London. Periods. Levant 46, 98–117. Vardi, J., Shilstein, S., Shalev, S., Yekutieli, Y., 2008. The Early Bronze Age IV chipped Shilstein, S., Shalev, S., Yekutieli, Y., 2014. Appendix: XRF study of archaeological and and ground stone assemblages of 'En Yahav and its relation to copper smelting ac- metallurgical material from copper smelting sites in Timna. In: Tebes, J.M. (Ed.), tivities. Mitekufat Haeven 38, 1–20. Unearthing the Wilderness: Studies on the History and Archaeology of the Negev and Warnock, P.J., 1991. Analysis of Wood Charcoal from Be'er Resisim, the ‘Central Negev Edom in the Iron Age, pp. 85–102 Leuven. Highlands Project,’ Israel: Implications for Economic and Cultural Strategies (M.A. Twiss, P.C., Suess, E., Smith, R.M., 1969. Morphological classification of grass phytoliths. thesis). Texas A&M University, College Station, Texas. Soil Sci. Soc. Am. Proc. 33, 109–115. Weiner, S., 2010. Microarchaeology: Beyond the Visible Archaeological Record. Valamoti, S.M., Charles, M., 2005. Distinguishing food from fodder through the study of Cambridge, UK. charred plant remains: an experimental approach to dung-derived chaff. Veg. Hist. Xie, X., Yan, N., Li, L., Shen, H., 1985. Usable values for Chinese standard reference Archaeobotany 14, 528–533. samples of stream sediments, soils and rocks: GSD 9-12, GSS 1-8 and GSR 1-6. Vardi, J., 2014. The lithic assemblages. In: Saidel, B.A., Haiman, M. (Eds.), Excavations in Geostand. Newslett. 9, 277–280. the Western Negev Highlands (British Archaeological Reports International Series Xie, X., Yan, M., Wang, C., 1989. Geochemical standard reference samples GSD 9-12, GSS 2684), pp. 59–98. 1-8 and GSR 1-6. Geostand. Newslett. 13, 83–179. Vardi, J., Rosen, S.A., Hermon, S., 2007. The economy of the Early Bronze Age IV period Yekutieli, Y., Shilstein, S., Shalev, S., 2005. 'En Yahav: a copper smelting site in the (c. 2200–2000): the lithic evidence. In: Saidel, B.A., van der Steen, E.J. (Eds.), On the 'Arava. Bull. Am. Sch. Orient. Res. 340, 1–21. Fringe of Society: Archaeological and Ethnographic Perspectives on Pastoral and

726 Appendix 1: Nahal Boqer 66: mineralogical and microremain results for samples utilized in the study (n = 54). Key to mineralogical results: Ca = calcite; Cl = clay, (ua) = unaltered, (a) = altered; Q = quartz; G = gypsum; P = carbonated hydroxylapatite; Ar = aragonite; SN = sodium nitrate. All microremains are reported in millions / 1 g of sediment. Two samples were analyzed in duplicate thus showing average and standard deviation values. PSR = pseudomorph to spherulite ratio; 14C = samples associated with radiocarbon determinations (Dunseth et al. 2017: Table 3).

Dung Ash Phytolith Spherulite Pseudomorph Area Context Locus Interpretation Sample Color Description Mineralogy (106 / g of PSR 14C (106 / g of (106 / g of sediment) sediment) sediment)

Controls Ridge south of Aeolian dust NB-C1 Yellow- Ridge south of site Ca > Cl (ua) Q G 0.13 0.74 0.00 - site brown Ridge southwest Aeolian dust NB-C2 Yellow- Ridge southwest of site Ca > Cl (ua) Q 0.14 0.00 0.00 - of site brown

Wadi west of site Wadi sediments NB-C3 Yellow- Wadi west of site Ca > Cl (ua) Q 0.16 0.00 0.00 - (aeolian + alluvial brown sediment) Ridge northwest Aeolian dust NB-C4 Yellow- Ridge northwest of site Ca > Cl (ua) Q 0.14 0.50 0.00 - of site brown

Ridge northeast of Aeolian dust NB-C5 Yellow- Ridge northeast of site Ca > Cl (ua) Q G 0.17 0.00 0.00 - site brown

Mid-slope Aeolian dust + NB-C6 Yellow- Mid-slope between site Ca > Cl (ua) Q G 0.32 0.00 0.00 - between site and alluvium brown and wadi wadi Southern Open area (?) 16/NB/1 Aeolian dust NB-1.8 Yellow- Yellow brown aeolian Ca > Cl (ua) Q G P? 0.49 23.06 0.00 - Complex (Fig. 4A) brown sediment NB-1.11 Yellow- Yellow brown aeolian Ca > Cl (ua) Q G P 0.75 54.79 0.00 - brown sediment Ash feature (?) NB-1.5 Red- Red-brown feature Ca > Cl (ua) Q 1.70 32.71 0.21 0.006 brown NB-1.6 Red- Red-brown feature Ca > Cl (ua) Q 0.75 - - - brown Ash feature (?) NB-1.1 Yellow- Yellow-grey feature Ca > Cl (ua) Q P 1.97 - - - grey with burnt pebbles and charcoal Dung accumulation NB-1.2 Yellow- Yellow-grey sediment Ca > Cl (ua) Q P 0.89 75.06 0.00 - grey with charcoal flecks NB-1.3 Yellow- Yellow-grey sediment Ca > Cl (ua) Q 1.13 69.23 0.00 - grey with charcoal flecks below wall stones Dung Ash Phytolith Spherulite Pseudomorph Area Context Locus Interpretation Sample Color Description Mineralogy (106 / g of PSR 14C (106 / g of (106 / g of sediment) sediment) sediment) NB-1.7 Yellow- Yellow-grey sediment Ca > Cl (ua) Q P 1.18 - - - grey Living surface (on NB-1.1 Yellow- Yellow-grey on Ca > Cl (ua) Q G 1.27 51.51 0.00 - bedrock) grey bedrock NB-1.4 Yellow- Yellow-brown Ca > Cl (ua) Q P 0.59 4.97 0.00 - brown sediment on bedrock NB-1.9 Yellow- Yellow-brown Ca > Cl (ua) Q P 0.82 - - - brown sediment on bedrock Southern Room 16/NB/2 Aeolian dust NB-2.5 Yellow- Yellow brown aeolian Ca > Cl (ua) Q P 1.04 19.97 0.00 - × Complex (Fig. 4B) brown sediment NB-2.4 Yellow- Yellow-brown Ca > Cl (ua) Q G P 1.03 20.34 0.09 - × brown sediment within stone collapse Dung accumulation NB-2.3 Yellow- Yellow-grey compact Cl (ua) > Ca Q G 2.16 16.24 0.00 - × grey sediment NB-2.7 Yellow- Yellow-brown below Ca > Cl (ua) Q P 1.38 89.63 0.10 - × brown wall stones Living surface (on NB-2.2 Yellow- Fine compact yellow- Ca > Cl (ua) Q G P? 3.90 ± 0.86 47.66 0.08 - × bedrock) grey brown sediment NB-2.6 Yellow- Yellow-grey sediment G > Ca > Cl (ua) 3.13 131.84 0.00 - × grey on bedrock NB-2.8 Yellow- Yellow-grey sediment Ca > Cl (ua) Q G P 1.57 47.42 0.26 - × grey on bedrock Southern Room 16/NB/6 Aeolian dust NB-6.5 Yellow- Yellow-brown aeolian Ca > Cl (ua) Q P 0.54 8.08 0.03 - Complex brown sediment NB-6.3 Yellow- Compact yellow- Ca > Cl (ua) Q 0.56 6.07 0.10 - brown brown aeolian sediment Dung accumulation NB-6.4 Grey- Grey-yellow on Ca > Cl (ua) Q P 0.34 14.81 0.10 - yellow bedrock NB-6.1 Yellow- Yellow-brown on Ca > Cl (ua) G Q 0.61 27.48 0.00 - brown bedrock Southern Room 16/NB/8 Dung accumulation NB-8.1 Yellow- Yellow-grey on Ca > Cl (ua) Q P 3.76 39.74 0.24 - × Complex (Fig. 4E) grey bedrock NB-8.2 Yellow- Yellow-grey on Ca > Cl (ua) Q P 4.51 38.29 0.00 - × grey bedrock Southern Internal enclosure 16/NB/9 Aeolian dust NB-9.3 Yellow- Yellow-brown aeolian Ca > Cl (ua) Q P 0.15 6.84 0.00 - Complex (Fig. 4F) brown sediment Dung accumulation NB-9.1 Yellow- Yellow-grey on Ca > Cl (ua) Q G P 0.48 7.75 0.00 - grey bedrock NB-9.2 Yellow- Yellow-grey on Ca > Cl (ua) G Q P 0.71 31.89 0.00 - grey bedrock Dung Ash Phytolith Spherulite Pseudomorph Area Context Locus Interpretation Sample Color Description Mineralogy (106 / g of PSR 14C (106 / g of (106 / g of sediment) sediment) sediment)

Southern Internal enclosure 16/NB/12 Aeolian dust NB-12.3 Yellow- Yellow-brown aeolian Ca > Cl (ua) Q P? 1.06 3.60 0.00 - Complex brown sediment

Dung accumulation NB-12.2 Yellow- Yellow-grey Ca > Cl (ua) Q P 0.47 49.09 0.30 - grey accumulation (with snail shell inclusions)

NB-12.1 Yellow- Yellow-grey sediment Ca > Cl (ua) Q P 1.73 19.25 0.00 - grey on bedrock (with snail shell inclusions)

Northern Internal enclosure 16/NB/3 Dung accumulation NB-3.5 Yellow- Yellow-brown above Ca > Cl (ua) G Q P 1.74 16.78 0.60 - Complex (Fig. 4C) brown bedrock

NB-3.4 Yellow- Yellow-brown above Ca > Cl (ua) Q P? 0.34 30.14 0.47 - grey bedrock with snail shell inclusions

Dung (ash?) NB-3.1 Grey- Grey-yellow below Ca > Cl (ua) G Q P 1.67 125.61 0.10 0.001 yellow stone collapse

NB-3.2 Grey- Grey-brown below Ca > Cl (ua) Q P 1.32 123.90 0.00 0.000 brown stone collapse

NB-3.3 Grey- Grey-brown on Ca > Cl (ua) Q P 0.57 195.68 0.60 0.003 brown bedrock NB-3.7 Grey- Grey-brown on Ca > Cl (ua) Q G 0.70 86.52 0.11 0.001 brown bedrock Northern Room 16/NB/4 Aeolian dust NB-4.7 Yellow- Yellow brown aeolian Ca > Cl (ua) Q P 0.18 14.40 0.24 - Complex (Fig. 4D) brown sediment

Dung accumulation NB-4.6 Yellow- Yellow-brown Ca > Cl (ua) Q G P 0.27 1.84 0.14 0.417 brown accumulation

NB-4.3 Grey- Grey-brown on Ca > Cl (ua) Q G P 0.45 3.08 0.57 0.184 brown bedrock NB-4.5 Grey- Grey-brown on Ca > Cl (400-500 °C) 0.20 2.29 0.95 0.778 brown bedrock Ar G?

Northern Room 16/NB/7 Aeolian dust NB-7.2 Yellow- Crumbly compact Ca > Cl (ua) Q 0.27 2.99 0.12 - Complex brown yellow-brown sediment near surface

Dung accumulation NB-7.1 Yellow- Yellow-brown on Ca > Cl (ua) Q 0.53 14.77 0.00 - Brown bedrock

NB-7.3 Brown- Brown-grey soft Ca > Cl (ua) Q P 1.24 15.32 0.00 - grey sediment on bedrock (+ root turbation)

Structure Rectangular 16/NB/11 Aeolian dust NB-11.4 Yellow- Yellow-brown aeolian Ca > Cl (ua) Q 0.04 0.00 0.00 - 16/NB/11 structure brown sediment Dung Ash Phytolith Spherulite Pseudomorph Area Context Locus Interpretation Sample Color Description Mineralogy (106 / g of PSR 14C (106 / g of (106 / g of sediment) sediment) sediment)

Accumulation on NB-11.5 Yellow- Yellow-grey sediment Ca > Cl (ua) Q 0.04 0.07 0.00 - living surface grey between and below (bedrock) wall stones

NB-11.3 Light Light grey sediment 5 Ca > Cl (ua) Q G 0.04 0.00 0.00 - grey cm above NB-11.2

NB-11.2 Light Light grey sediment on Ca > Cl (ua) Q 0.07 0.00 0.07 - grey bedrock

NB-11.1 Grey Grey-sandy sediment Ca > Cl (ua) Q G 0.11 0.00 0.00 - on bedrock

Appendix 2: Ein Ziq: mineralogical and microremain results for samples utilized in the study (n = 81). Key to mineralogical results: Ca = calcite, Cl = clay, (ua) = unaltered, (a) = altered; Q = quartz; G = gypsum; P = carbonated hydroxylapatite; Ar = aragonite; SN = sodium nitrate. All microremains are reported in millions / 1 g of sediment. Samples analyzed in duplicate show average and standard deviation values. Associated geoarchaeological analyses are noted: XRF (this study); MB = micromorphology block sample (Dunseth et al. 2017: Fig. 8); OSL = optically stimulated luminescence sample (Dunseth et al. 2017: Table 2); 14C = radiocarbon dated sample (previously reported in Dunseth et al. 2017: Table 1).

Dung Ash Phytolith Spherulite Pseudomorph Area Context Locus Interpretation Sample Color Description Mineralogy (106 / g of PSR MB XRF OSL 14C (106 / g of (106 / g of sediment) sediment) sediment) Controls Taqiya Formation Geological EZ-C1 Green-grey Taqiya Formation Ca > Cl (kaolinite) 0.00 0.00 0.00 - × formation Q Hematite Nahal Zin Wadi sediment EZ-C2 Yellow- Alluvial wadi Ca > Cl (ua) Q 0.00 0.00 0.00 - × brown sediment Nahal Zin Wadi sediment EZ-C3 White- Alluvium + marl Ca > Cl (ua) Q 0.07 0.04 0.00 - × yellow wadi sediment Nahal Zin Flint EZ-C4 Dark brown Flint from wadi Flint - - - - (Nahal Zin) Nahal Zin Weathered EZ-C5 Yellow- Weathered Ca - - - - limestone white limestone (Nahal Zin) Nahal Zin Limestone EZ-C6 White Hard limestone Ca - - - - Nahal Zin Limestone EZ-C7 Yellow- Sandy limestone Ca Q - - - - white Taqiya Formation Pyrite concretion EZ-C9 Red-orange Pyrite concretion Pyrite - - - - from Taqiya Formation c. 50 m east of site Aeolian dust EZ-C10 Yellow- Yellow desert Ca > Cl (ua) Q 0.11 0.00 0.00 - × brown pavement crust c. 50 m east of site Aeolian dust EZ-C11 Yellow- Yellow desert Ca > Cl (ua) Q 0.15 ± 0.21 0.02 ± 0.03 0.00 - × brown pavement dust c. 50 m east of site Alluvial terrace EZ-C12 White- Alluvial terrace Ca > Cl (ua) Q 0.05 0.00 0.00 - × yellow sediment c. 50 m east of site Alluvial terrace EZ-C13 White- Alluvial terrace Ca > Cl (ua) Q 0.05 0.00 0.00 - × yellow sediment c. 50 m east of site Alluvial terrace EZ-C14 White- Alluvial terrace Ca > Cl (ua) Q 0.00 0.05 ± 0.02 0.00 - × yellow sediment c. 50 m east of site Aeolian dust EZ-C15 Yellow- Yellow desert Ca > Cl (ua) Q 0.12 0.00 0.00 - × brown pavement A Structure 14/A/2 14/A/2 Aeolian dust EZ-65 Yellow- Yellow-brown Ca > Cl (ua) Q 0.03 0.00 0.00 - × (Fig. 8A) brown sediment between stone collapse Collapse EZ-64 Grey-white Upper grey-white Ca > Cl (ua) Q G 0.19 0.09 0.05 - × sediment Hearth EZ-68 Light grey Grey sediment Ca > Cl (ua) Q G P 0.25 0.16 0.00 - × × associated with charcoal EZ-63 Red-brown Red brown Ca > Cl (ua) Q G P 0.26 0.05 0.00 - × sediment below hearth Subsurface EZ-62 White White sediment Ca > Cl (ua) Q G 0.09 0.00 0.00 - × (alluvial terrace) below hearth EZ-61 White Pebbly sediment Ca > Cl (ua) Q 0.13 0.00 0.00 - below hearth A Structure 14/A/4 14/A/4 Mix of aeolian EZ-103 Grey-brown Compact grey- Ca > Cl (ua) Q G 0.07 0.13 0.09 - × × × (Fig. 8B) dust, collapse of brown sediment construction associated with material charcoal EZ-104 Grey-brown Compact grey- Ca > Cl (ua) Q G 0.05 0.20 0.04 - × × × brown sediment associated with micromorphology block EZB-4 F Ash deposit 15/F/TP3 Aeolian dust EZ-392 Yellow- Yellow-brown Ca > Cl (ua) Q G 0.23 0.04 0.00 - × 15/F/TP3 brown aeolian dust (Fig. 8F) Ash deposit EZ-394 Grey Bulk sediment of Ca > Cl (ua) Q G 0.29 0.00 0.65 0.31 × × ash EZ-393 Grey Soft grey ashy Ca > Cl (ua) Q G 0.55 0.00 0.31 0.65 × sediment EZ-391 Grey Soft grey ashy Ca > Cl (ua) Q G 0.26 0.04 0.40 9.00 × × sediment EZ-390 Dark grey Dark grey ashy Ca > Cl (ua) Q G 0.94 0.06 0.06 1.00 × sediment SN Alluvial terrace EZ-389 Yellow- Yellow-brown Ca > Cl (ua) Q G 0.07 0.07 0.00 - × brown compact sediment from alluvial terrace EZ-388 Grey-brown Grey-brown Ca > Cl (ua) Q 0.04 0.00 0.13 - × crumbly sediment with white nodules in alluvial terrace J Structure 14/J/21 14/J/9 Aeolian dust EZ-108 Yellow- Yellow-brown Ca > Cl (ua) Q G 0.05 0.00 0.00 - × brown aeolian dust (South Section) Bedded dust EZ-107 Light grey Bedded light grey Ca > Cl (ua) Q G 0.02 ± 0.01 0.00 0.00 - × (Fig. 8C) sediments fine sediments 14/J/21 Collapse EZ-115 Grey-green Grey-green Ca > Cl (ua) Q G 0.02 ± 0.02 0.03 ± 0.04 0.00 - × collapse sediment EZ-106 Grey-green Grey-green Ca > Cl (ua) Q G P 0.02 ± 0.01 0.00 0.05 - × collapse sediment Accumulation on EZ-105 Brown-grey Brown-grey Ca > Cl (ua) Q G P 0.06 ± 0.02 0.05 0.05 - × beaten earth floor sediments immediately on floor Floor EZ-81 Grey-brown Grey-brown Ca > Cl (ua) G Q P 0.03 0.00 0.31 - × × compact floor below collapse EZ-82 Grey-brown Grey-brown Ca > Cl (ua) G Q P 0.19 0.00 0.05 - × × compact floor below collapse EZ-84 Grey-brown Grey-brown floor Ca > Cl (ua) G Q P 0.07 ± 0.05 0.04 ± 0.06 0.12 ± 0.06 - × × with possible plant impressions J Structure 14/J/21 14/J/9 Aeolian dust EZ-342 Yellow- Yellow-brown Ca > Cl (ua) Q G 0.05 0.08 0.00 - × (East Section) brown sediment above P? (Fig. 8C) collapse 14/J/21 Collapse EZ-347 Grey-green Grey-green Ca > Cl (ua) Q G 0.02 0.08 0.00 - × collapse sediment P? EZ-346 Grey-green Grey-green Ca > Cl (ua) Q G 0.10 0.00 0.05 - × × collapse sediment P? EZ-341 Grey-green Grey-green Ca > Cl (ua) Q G 0.00 0.00 0.00 - × collapse sediment P? 15/J/14 Dark grey ash EZ-344 Dark grey Dark grey ashy Ca > Cl (400-500 7.41 0.25 2.90 11.40 × × sediment °C) Q G P EZ-340 Dark grey Dark grey ashy Ca > Cl (ua) Q G P 7.63 0.04 8.55 229.00 × sediment Light grey ash EZ-345 Light grey Light grey ashy Ca > Cl (400-500 0.52 ± 0.38 0.00 0.69 ± 0.15 18 × sediment °C) Q G P? Dark brown ash EZ-339 Dark brown Dark brown soft Ca > Cl (400-500 3.00 0.04 2.47 69.00 ashy sediment °C) Q G P Aeolian dust (?) EZ-338 Yellow- Yellow soft Ca > Cl (ua) Q 1.11 0.08 0.54 6.50 [Hiatus?] brown sediment directly below dark brown ash EZ-336 Yellow- Yellow soft Ca > Cl (ua) Q G 0.03 0.18 0.00 0.00 × brown sediment directly below dark brown ash Brown (ash?) EZ-337 Brown-grey Dark brown Ca > Cl (ua) Q P? 1.55 0.04 1.71 41.00 sediment sediment directly below yellow sediment Heated substrate EZ-343 Red-yellow Red-yellow soft Ca > Cl (400-500 0.27 ± 0.11 0.08 ± 0.11 0.03 ± 0.04 - × sediment directly °C) Q P below hearth Alluvial terrace EZ-335 Red-yellow Soft red sediment Ca > Cl (ua) Q G 0.02 0.00 0.00 - × rich in angular pebbles below structure J North Section Sq. 14/J/10 Aeolian dust EZ-41 Yellow- Yellow-brown Ca > Cl (ua) Q G 0.17 0.00 0.00 - × × × × M21 (Open Area) brown aeolian dust (Fig. 8D) EZ-40 Yellow- Yellow-brown Ca > Cl (ua) Q G 0.15 0.04 0.00 - × × × × brown aeolian dust 14/J/16 Ash deposit EZ-39 Light grey Grey ash lens Ca > Cl (400-500 5.95 0.03 2.56 10.55 × × × × °C) Q G P? EZ-38 Light grey Grey ash lens Ca > Cl (ua) Q G 1.89 0.09 0.97 77.00 × × × × P? 14/J/20 Alluvial terrace EZ-37 Red-orange Red-orange alluvial Ca > Cl (ua) Q G 0.59 0.37 0.00 - × × × × terrace P? EZ-36 Red-orange Red-orange alluvial Ca > Cl (ua) Q G 0.34 0.04 0.13 - × × × × terrace Op? P? J Structure 14/J/19 14/J/11 Aeolian dust EZ-334 Yellow- Yellow-brown Ca > Cl (ua) Q G 0.08 0.00 0.00 - brown aeolian dust 14/J/19 Collapse EZ-333 Green-grey Soft green-grey Ca > Cl (ua) Q G 0.03 0.00 0.00 - sandy collapse sediments EZ-332 Green-grey Green grey collapse Ca > Cl (ua) Q G 0.10 0.00 0.00 - sediments, small P? pebble inclusions EZ-331 Green-grey Green-grey Ca > Cl (ua) Q G 0.14 0.00 0.00 - collapse sediments, P? small angular pebbles 15/J/3 Surface EZ-330 Grey-brown Accumulation on Ca > Cl (ua) Q G 0.20 ± 0.00 0.00 0.00 - surface EZ-329 Grey-brown Grey-brown Ca > Cl (ua) Q G 0.33 0.00 0.00 - surface sediment P? Alluvial terrace EZ-328 Light Light brown Ca > Cl (ua) Q G? 0.04 0.00 0.00 - brown alluvial sediment P? (below structure) J Structure 14/J/19 (Grinding stone EZ-208 Light Sediment brushed - 0.10 - - - sediment) brown from grinding stone (AR 1) (Grinding stone EZ-209 Light Sediment brushed - 0.54 - - - sediment) brown and washed from grinding stone with distilled water (AR 1) J South Section Sq. 14/J/13 Aeolian dust EZ-327 Yellow- Yellow-brown Ca > Cl (ua) Q G? 0.03 0.04 0.04 - M22 (Open Area) brown aeolian dust P? Collapse EZ-326 Grey-brown Grey-brown Ca > Cl (ua) Q G? 0.06 0.00 0.00 - collapse mixed P? with large stones EZ-325 Grey-brown Grey-brown Ca > Cl (ua) Q G 0.02 0.05 0.00 - collapse mixed with large stones 14/J/17 Surface EZ-324 Brown-red Red-brown surface Ca > Cl (ua) Q G 0.02 0.00 0.00 - P? 14/J/22 Alluvial terrace EZ-323 Yellow- Alluvial terrace Ca > Cl (ua) Q P? 0.07 0.00 0.00 - brown sediment K Structure 15/K/5 15/K/1 Aeolian dust EZ-377 Yellow- Bedded yellow- Ca > Cl (ua) Q G 0.21 0.00 0.00 - brown brown fine sediment EZ-376 Yellow- Yellow-brown Ca > Cl (ua) Q G P 0.09 0.00 0.00 - brown aeolian dust 15/K/5 Collapse EZ-375 Grey-green Crumbly grey- Ca > Cl (ua) Q G 0.05 0.00 0.00 - brown collapse P? sediment EZ-372 Grey-green Crumbly grey- Ca > Cl (ua) Q 0.10 0.00 0.00 - brown collapse sediment Hearth EZ-379 Grey Grey fine ash Ca > Cl (400-500 1.26 0.21 0.99 4.75 °C) Q EZ-374 Grey Grey fine ash with Ca > Cl (ua) Q G 0.32 0.39 0.11 0.27 charcoal fragments P? EZ-371 Grey Grey fine ash with Ca > Cl (ua) Q G P 5.35 0.06 1.80 29.00 charcoal fragments Alluvial terrace EZ-373 Red-yellow Red soft sediment Ca > Cl (ua) Q G 0.03 0.00 0.00 - directly below hearth EZ-370 Yellow- Yellow-brown Ca > Cl (ua) Q G 0.03 0.00 0.00 - brown alluvial sediment K Hearth 15/K/8 15/K/8 Hearth EZ-348 Dark grey Dark grey ash from Ca > Cl (ua) Q 0.75 0.00 0.08 2.00 × × (Fig. 8E) stone-lined hearth EZ-349 Dark grey Dark grey ash from Ca > Cl (ua) Q 0.38 0.19 0.34 1.75 × × stone-lined hearth Appendix 3: Accuracy tests: XRF results of GSS-1 standards (n = 7) compared to certified reference values. δ signifies difference between average measured value and reference value, expressed as a percent (%) and absolute (abs) value.

Al O SiO P O S K O CaO TiO MnO Fe O Cu Zn Sn Pb Sum Date Method Sample Description 2 3 2 2 5 2 2 2 3 Notes (%) (%) (%) (%) (%) (%) (%) (%) (%) (ppm) (ppm) (ppm) (ppm) (%) - Reference GSS-1 Geological 14.18 62.60 0.17 0.08 2.59 1.72 0.81 0.23 5.19 21 680 6 98 - Xie et al. Value Soil 1985; Xie et Standard al. 1989: Table 9-10 27/05/2014 Powder GSS-1 Geological 17.78 61.87 0.22 0.05 2.53 1.87 0.85 0.24 4.97 26 802 15 95 90.62 Homogenized Soil to 425 µm Standard 28/05/2014 Powder GSS-1 Geological 18.01 63.07 0.23 0.04 2.59 1.91 0.87 0.24 5.21 29 849 17 98 92.43 Homogenized Soil to 425 µm Standard 17/09/2015 Powder GSS-1 Geological 17.56 59.49 0.21 0.07 2.54 1.90 0.85 0.24 5.12 30 831 16 96 88.23 Homogenized Soil to 425 µm Standard 20/09/2015 Powder GSS-1 Geological 17.54 59.40 0.21 0.07 2.52 1.89 0.84 0.24 5.14 24 833 18 97 88.11 Homogenized Soil to 425 µm Standard 30/09/2015 Powder GSS-1 Geological 17.57 59.82 0.21 0.07 2.54 1.90 0.85 0.24 5.11 29 831 18 95 88.58 Homogenized Soil to 425 µm Standard 06/10/2015 Powder GSS-1 Geological 17.70 60.21 0.22 0.06 2.56 1.92 0.86 0.24 5.19 26 842 20 98 89.21 Homogenized Soil to 425 µm Standard 09/10/2015 Powder GSS-1 Geological 17.68 59.73 0.21 0.07 2.56 1.91 0.86 0.24 5.18 24 839 20 97 88.69 Homogenized Soil to 425 µm Standard AVERAGE 17.69 60.51 0.22 0.06 2.55 1.90 0.85 0.24 5.13 27 832 18 96 89.41 STDEV 0.17 1.41 0.01 0.01 0.02 0.02 0.01 0.00 0.08 2 15 2 1 1.58 δ (%) 124.76 96.67 127.38 76.15 98.43 110.45 105.30 104.39 98.86 128 122 290 98 δ (abs) 3.51 -2.09 0.05 -0.02 -0.04 0.18 0.04 0.01 -0.06 6 152 12 -2 Appendix 4: XRF results of sediment, geological and control samples (n = 63) from Ein Ziq. Note all samples are color-coded for convenience: light blue: Taqiya Formation; green = alluvial terrace sediments; yellow = aeolian dust; light- grey = collapse; dark grey = ash; brown = floor or surface sediments; red = pyrite nodule from collapse.

Date Method Sq. Locus Sample Context Al2O3 SiO2 P2O5 SO3 K2O CaO TiO2 MnO Fe2O3 Cu Zn Sn Pb Sum LOI (%) (%) (%) (%) (%) (%) (%) (%) (%) (ppm) (ppm) (ppm) (ppm) (%) 27/05/2014 Powder - Control EZC-01 Taqiya Formation 10.79 23.09 0.60 0.52 0.79 31.21 0.42 0.09 3.69 2 193 13 17 71.97 56.44 17/09/2015 Powder - Control EZC-01b Taqiya Formation 10.23 21.83 0.57 0.53 0.79 30.89 0.41 0.08 3.60 32 189 15 16 69.66 30/09/2015 Powder - Control EZC-02 Alluvial wadi sediment (Nahal Zin) 4.29 15.78 0.56 0.39 0.59 40.90 0.33 0.04 2.07 27 117 16 8 66.01 57.53

30/09/2015 Powder - Control EZC-03 Alluvium + marl wadi sediment (Nahal Zin) 5.13 15.66 0.61 0.70 0.56 42.81 0.26 0.03 2.14 27 113 10 10 68.44 56.18 27/05/2014 Powder - Control EZC-10 Yellow desert pavement 7.68 27.12 0.49 0.17 1.06 30.89 0.72 0.06 3.19 23 86 12 9 72.70 56.03 09/10/2015 Powder - Control EZC-10b Yellow desert pavement 7.13 25.35 0.46 0.15 1.02 29.84 0.70 0.06 3.20 27 88 16 11 69.08 27/05/2014 Powder - Control EZC-11 Yellow aeolian dust 4.30 18.36 0.41 0.13 0.67 37.10 0.49 0.04 2.25 32 87 12 14 65.40 58.32 17/09/2015 Powder - Control EZC-11b Yellow aeolian dust 3.76 16.68 0.40 0.12 0.65 36.93 0.43 0.04 2.24 28 88 13 15 62.67 30/09/2015 Powder - Control EZC-12 Alluvial terrace 2.52 12.69 0.40 0.75 0.47 41.87 0.30 0.03 1.85 24 73 18 5 62.16 59.20 30/09/2015 Powder - Control EZC-13 Alluvial terrace 3.17 14.08 0.42 1.42 0.56 39.87 0.37 0.03 2.09 25 80 14 5 63.49 58.87 30/09/2015 Powder - Control EZC-14 Alluvial terrace 4.38 17.93 0.36 0.57 0.65 31.12 0.50 0.04 2.83 32 70 14 8 59.71 62.55 30/09/2015 Powder - Control EZC-15 Desert pavement 8.15 26.62 0.49 0.14 1.18 28.05 0.71 0.07 3.80 34 107 18 9 70.18 57.88 30/09/2015 Powder - 14/A/2 EZ-065 Aeolian deposit (modern surface) 0.78 5.17 0.43 0.10 0.22 52.50 0.16 0.02 0.94 11 38 16 4 60.53 58.27 30/09/2015 Powder - 14/A/2 EZ-064 White-grey collapse (Structure 14/A/2) 2.70 12.77 0.43 4.17 0.52 36.69 0.29 0.03 1.52 17 75 19 9 60.18 61.80 20/09/2015 Powder - 14/A/2 EZ-068 Grey ash (Hearth 14/A/2) 3.27 14.63 0.53 4.08 0.58 41.02 0.32 0.04 1.70 21 88 11 6 67.20 30/09/2015 Powder - 14/A/2 EZ-068b Grey ash (Hearth 14/A/2) 2.86 13.06 0.52 3.28 0.52 37.37 0.28 0.04 1.45 19 92 13 8 60.43 61.47 30/09/2015 Powder - 14/A/2 EZ-063 Red-brown sediment (Hearth 14/A/2) 3.32 14.60 0.49 1.54 0.60 38.26 0.32 0.04 1.63 26 79 17 10 61.88 60.34 30/09/2015 Powder - 14/A/2 EZ-062 Sterile white sediment (alluvial terrace) 3.29 14.85 0.47 5.01 0.61 41.01 0.35 0.04 1.93 19 82 12 7 68.45 56.85 27/05/2014 Powder M21 14/J/10 EZ-040 Aeolian deposit (modern surface) 4.90 20.33 0.49 0.31 0.84 32.13 0.52 0.05 2.26 26 81 12 11 63.35 60.49 09/10/2015 Powder M21 14/J/10 EZ-040b Aeolian deposit (modern surface) 4.93 20.53 0.51 0.27 0.89 34.26 0.51 0.05 2.43 27 87 12 11 65.79 28/05/2014 Powder M21 14/J/10 EZ-041 Aeolian deposit (modern surface) 4.44 22.07 0.51 0.31 0.81 34.29 0.49 0.05 2.31 22 83 9 8 67.10 58.08 09/10/2015 Powder M21 14/J/10 EZ-041b Aeolian deposit (modern surface) 4.29 20.75 0.48 0.27 0.81 34.05 0.49 0.05 2.36 26 83 8 11 65.23 27/05/2014 Powder M21 14/J/16 EZ-038 Secondary ash deposit 2.26 12.24 0.95 12.93 0.41 39.34 0.22 0.05 1.28 51 153 6 9 71.23 56.47 (Ash Deposit 14/J/16) 09/10/2015 Powder M21 14/J/16 EZ-038b Secondary ash deposit 2.10 11.42 0.90 12.16 0.39 38.50 0.22 0.05 1.24 49 149 10 10 68.36 (Ash Deposit 14/J/16) 28/05/2014 Powder M21 14/J/16 EZ-039 Secondary ash deposit 1.81 10.12 0.80 8.93 0.30 32.26 0.16 0.04 0.95 35 119 10 5 56.51 65.18 (Ash Deposit 14/J/16) 09/10/2015 Powder M21 14/J/16 EZ-039b Secondary ash deposit 7.78 10.00 0.85 9.35 0.33 37.76 0.19 0.04 1.02 39 131 11 8 62.45 (Ash Deposit 14/J/16) 27/05/2014 Powder M21 14/J/20 EZ-036 Sterile alluvial terrace sediment 3.96 17.64 1.23 4.50 0.63 39.33 0.37 0.05 1.90 31 117 8 8 71.38 55.37 17/09/2015 Powder M21 14/J/20 EZ-036b Sterile alluvial terrace sediment 3.69 16.59 1.18 4.15 0.62 38.77 0.38 0.05 1.91 28 119 7 7 68.96 28/05/2014 Powder M21 14/J/20 EZ-037 Sterile alluvial terrace sediment 2.70 13.22 1.03 8.14 0.44 36.65 0.24 0.04 1.29 31 98 12 7 65.05 59.84 09/10/2015 Powder M21 14/J/20 EZ-037b Sterile alluvial terrace sediment 2.55 12.81 0.95 7.52 0.41 34.16 0.21 0.04 1.21 30 99 11 7 60.95 28/05/2014 Powder L21 14/J/9 EZ-107 Aeolian deposit (modern surface) 2.83 10.78 0.49 11.20 0.50 36.45 0.20 0.04 1.53 16 72 13 5 65.49 59.66 17/09/2015 Powder L21 14/J/9 EZ-107b Aeolian deposit (modern surface) 2.84 10.81 0.51 11.10 0.52 38.13 0.20 0.04 1.56 20 76 12 7 66.90 28/05/2014 Powder L21 14/J/9 EZ-108 Aeolian deposit (modern surface) 4.42 15.59 0.57 3.82 0.71 35.25 0.30 0.05 2.28 25 93 9 11 64.55 59.49 17/09/2015 Powder L21 14/J/9 EZ-108b Aeolian deposit (modern surface) 4.60 16.03 0.59 3.54 0.75 36.32 0.33 0.05 2.36 27 97 9 11 66.04 28/05/2014 Powder L21 14/J/21 EZ-115 Collapse (Structure 14/J/21) 3.53 12.64 0.55 5.10 0.51 36.49 0.22 0.04 1.69 19 74 14 8 62.03 60.86 17/09/2015 Powder L21 14/J/21 EZ-115b Collapse (Structure 14/J/21) 3.40 12.17 0.54 4.89 0.50 36.64 0.23 0.04 1.84 28 79 13 8 61.40 28/05/2014 Powder L21 14/J/21 EZ-106 Collapse (Structure 14/J/21) 5.18 15.75 0.75 1.16 0.71 38.68 0.30 0.05 2.16 27 99 10 13 66.18 57.74 09/10/2015 Powder L21 14/J/21 EZ-106b Collapse (Structure 14/J/21) 5.14 15.60 0.76 1.06 0.72 39.60 0.31 0.05 2.23 27 101 10 13 66.71 28/05/2014 Powder L21 14/J/21 EZ-105 Collapse (Structure 14/J/21) 4.91 16.50 0.71 1.64 0.68 37.90 0.32 0.06 2.21 32 106 14 11 66.33 57.93 09/10/2015 Powder L21 14/J/21 EZ-105b Collapse (Structure 14/J/21) 4.95 16.74 0.75 1.79 0.73 40.38 0.32 0.06 2.34 27 110 11 12 69.21 28/05/2014 Powder L21 14/J/21 EZ-081 Floor (Structure 14/J/21) 5.08 14.36 0.84 1.12 0.65 38.31 0.26 0.06 1.95 24 99 11 11 63.87 59.09 09/10/2015 Powder L21 14/J/21 EZ-081b Floor (Structure 14/J/21) 5.18 14.54 0.82 1.07 0.67 38.66 0.28 0.06 1.99 26 100 16 11 64.37 28/05/2014 Powder L21 14/J/21 EZ-082 Floor (Structure 14/J/21) 3.48 14.54 0.58 1.13 0.54 37.83 0.33 0.04 1.79 31 82 11 11 61.32 60.66 09/10/2015 Powder L21 14/J/21 EZ-082b Floor (Structure 14/J/21) 3.22 13.22 0.54 1.04 0.51 35.07 0.31 0.04 1.58 26 82 11 10 56.37 28/05/2014 Powder L21 14/J/21 EZ-084 Floor (Structure 14/J/21) 4.38 13.52 0.63 1.58 0.58 34.30 0.20 0.04 1.68 21 97 10 11 58.00 63.08 09/10/2015 Powder L21 14/J/21 EZ-084b Floor (Structure 14/J/21) 4.68 14.19 0.66 1.57 0.62 36.50 0.22 0.04 1.86 24 91 14 11 61.35 06/10/2015 Powder L21 14/J/9 EZ-342 Aeolian deposit (modern surface) 5.84 18.30 0.70 0.54 0.88 38.14 0.37 0.05 2.61 29 113 13 13 68.13 57.12 06/10/2015 Powder L21 14/J/21 EZ-347 Collapse (Structure 14/J/21) 5.37 15.77 0.83 2.22 0.69 35.15 0.23 0.04 1.90 23 85 11 9 62.94 60.64 06/10/2015 Powder L21 14/J/21 EZ-346 Collapse (Structure 14/J/21) 3.88 13.62 0.57 5.58 0.62 37.09 0.26 0.04 1.94 21 87 8 6 65.28 58.89 06/10/2015 Powder L21 14/J/21 EZ-341 Collapse (Structure 14/J/21) 4.31 12.67 0.74 2.16 0.63 36.35 0.23 0.05 2.17 31 105 11 14 61.32 60.44 06/10/2015 Powder L21 15/J/14 EZ-344 Dark grey ash (Hearth 15/J/14) 2.30 11.10 0.90 2.54 0.48 35.87 0.25 0.05 1.41 33 147 18 10 57.37 62.55 06/10/2015 Powder L21 15/J/14 EZ-340 Light grey ash (Hearth 15/J/14) 1.80 8.50 0.75 0.55 0.41 32.00 0.20 0.05 1.31 36 134 15 17 47.47 68.40 06/10/2015 Powder L21 15/J/14 EZ-336 Aeolian dust? [Hiatus?] (Hearth 15/J/14) 2.81 12.75 0.44 2.56 0.53 37.68 0.30 0.04 1.62 20 106 12 8 60.02 61.37 20/09/2015 Powder P20 15/K/4 EZ-316 Ash deposit (Open Area 15/K/4) 2.46 12.62 0.45 0.31 0.54 40.04 0.29 0.03 1.62 26 70 12 7 60.52 20/09/2015 Powder P20 15/K/4 EZ-317 Ash deposit (Open Area 15/K/4) 2.70 13.66 0.46 0.26 0.58 37.16 0.29 0.04 1.73 31 76 23 12 59.24 20/09/2015 Powder P20 15/K/5 15/K/5/LB1 Pyrite nodule from collapse 0.00 0.68 0.00 21.08 0.80 15.14 0.01 0.00 33.13 4 73 13 11 71.46 (Structure 15/K/5) 20/09/2015 Powder P20 15/K/8 EZ-348 Ash from stone-lined hearth (Hearth 15/K/8) 3.08 13.94 0.50 0.35 0.62 38.76 0.32 0.04 1.96 29 85 12 5 61.30 20/09/2015 Powder P20 15/K/8 EZ-349 Ash from stone-lined hearth (Hearth 15/K/8) 2.45 12.42 0.49 0.28 0.53 39.08 0.29 0.03 1.57 26 76 9 6 59.10 06/10/2015 Powder TP3 15/F/TP3 EZ-392 Aeolian deposit (modern surface) 3.69 15.43 0.52 0.47 0.75 37.86 0.36 0.04 2.10 28 87 12 10 62.97 59.33 06/10/2015 Powder TP3 15/F/TP3 EZ-391 Ash deposit (Open Area 15/F/TP3) 2.73 12.32 0.57 4.86 0.54 36.92 0.24 0.04 2.30 29 120 10 9 62.00 60.53 20/09/2015 Powder TP3 15/F/TP3 EZ-394 Ash deposit (Open Area 15/F/TP3) 2.24 10.75 0.58 6.35 0.46 43.06 0.23 0.04 1.35 28 86 13 8 66.64 - 06/10/2015 Powder TP3 15/F/TP3 EZ-389 Sterile alluvial terrace sediment 2.49 12.42 0.52 5.17 0.49 43.49 0.24 0.04 1.56 29 90 10 8 67.54 56.74 06/10/2015 Powder TP3 15/F/TP3 EZ-388 Sterile alluvial terrace sediment 3.47 15.86 0.45 0.46 0.59 41.20 0.44 0.05 2.13 23 88 7 6 65.85 57.47 TEL AVIV Vol. 43, 2016 43–75

Geoarchaeological Investigation at the Intermediate Bronze Age Negev Highlands Site of Mashabe Sade

Zachary C. Dunseth1, Andrea Junge3, Markus Fuchs3, Israel Finkelstein1* and Ruth Shahack-Gross2,4* 1Tel Aviv University, 2University of Haifa, 3Justus-Liebig-University, Giessen, 4Weizmann Institute of Science

Massive settlement activity characterizes the arid Negev Highlands during the Intermediate Bronze Age (ca. 2500–1950 BCE). However, the underlying subsistence basis of this population is poorly understood. Recent microarchaeological work at Iron Age sites in the Negev Highlands has shown the potential for recovering direct evidence for subsistence practices through analysis of the microscopic plant remains in degraded animal dung. Following these methods, this paper reports new macro- and micro- archaeological results of two sites near Mashabe Sade: a central Intermediate Bronze Age site, and for comparison, an ephemeral site in the immediate vicinity. At the central site, dated to the Intermediate Bronze Age by pottery and Optically Stimulated Luminescence (OSL), evidence is absent for any sort of food production. In contrast, identification of ancient livestock dung at the ephemeral site suggests that it was sustained by animal husbandry— yet the OSL results suggest these degraded dung deposits date to the Iron Age. Taken together, the Intermediate Bronze Age results from Mashabe Sade bolster arguments suggesting that central sites were supported mainly by trade and other alternative subsistence practices.

Keywords Mashabe Sade, Negev Highlands, Intermediate Bronze Age, Microarchaeology

After nearly a century of research, the Intermediate Bronze Age is still an enigmatic period in the settlement history of the southern Levant. Traditionally dated to ca. 2300–2000 BCE (e.g., Palumbo 2001), the Intermediate Bronze Age is characterized by the cessation of

* Co-directors of the Negev Highlands Project.

© The Institute of Archaeology of Tel Aviv University 2016 DOI 10.1080/03344355.2016.1161372 44 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

urban activity throughout the region, and a unique wave of settlement in the arid zones, especially the Negev Highlands. The impetus behind this dramatic settlement oscillation is unclear, though scholars have variously associated the phenomenon with changing political (e.g., Kochavi 1967), demographic (e.g., Kenyon 1951; Prag 1985), socio-economic (e.g., Dever 1980), and climatic conditions (e.g., Rosen 1987; Frumkin 2009). Until very recently, debate over the nature of the Intermediate Bronze Age had mellowed into a general consensus that the period was characterized in the north of modern Israel by a shift from urban life to rural agro-pastoralism (Finkelstein 1995: 88), and in the Negev, to subsistence practices characterized by animal husbandry and opportunistic agriculture (Dever 1985; Cohen 1999; Palumbo 2001). Hoards of copper ingots found at sites throughout the southern Levant, and the recent discovery of an Intermediate Bronze Age copper production centre in Wadi Faynan, suggest that a developed copper trade economy supported the Negev settlement system (Goren 1996; Haiman 1996; Adams 2000; Levy et al. 2002; Hauptmann, Schmitt- Strecker, Levy and Begemann 2015). These assumptions have been based primarily on macroarchaeological assemblages from a few excavated sites (e.g., mainly Kochavi 1967; Cohen and Dever 1978, 1979, 1981; Cohen 1999; Saidel 2002; Saidel et al. 2006; Dever 2014) and on ethnographic parallels to pre-modern Bedouin mode of life (e.g., Palmer 1871; Musil 1908; Marx 1967; Ginguld, Perevolotsky and Ungar 1997). Recently, an analysis of radiocarbon determinations from secure archaeological contexts redated the Intermediate Bronze Age to ca. 2500–1950 BCE (Regev et al. 2012). This changes the equation between the Intermediate Bronze Age and Egyptian history and hence forces a re-evaluation of the Intermediate Bronze Age in general, and the role of Negev settlement systems within it in particular. Simultaneously, developments in microarchaeology (sensu Weiner 2010) have given researchers new tools to explore aspects of human activity in the past which are not visible to the naked eye. Microarchaeological investigations have recently provided insight into subsistence practices and settlement organization through the study of sediments, i.e., geoarchaeology (e.g., Shahack-Gross, Marshall and Weiner 2003; Shahack-Gross et al. 2014; Shahack-Gross and Finkelstein 2008, 2015 for the Negev Highlands), paleoenvironment through the study of pollen grains (e.g., Langgut et al. 2015), and preservation and post-depositional processes through taphonomic studies (e.g., Courty, Goldberg and Macphail 1989; Cabanes et al. 2012; Gur-Arieh, Mintz, Boaretto and Shahack-Gross 2013). In addition, systematic numerical dating projects provided insight into the role of terrace agriculture in the Levant (e.g., Davidovich et al. 2012), including the Negev (Avni, Porat and Avni 2012, 2013). Having generally exhausted the interpretive potential of the macroarchaeological assemblage from the Negev Highlands, microarchaeological and radiometric dating methods can provide new approaches to reassess old assumptions. While many of the aforementioned microarchaeological and numerical dating studies focused on the Iron Age, no such studies have been carried out in order to better understand the Intermediate Bronze Age settlement system in the Negev Highlands. Here we present the first results of a systematic geoarchaeological study at Mashabe Sade—one of the largest and richest Intermediate Bronze Age sites in the region. Geoarchaeological Investigation at the Site of Mashabe Sade 45

Background to intermediate Bronze Age site typologies The central settlements are limited to the northern Negev Highlands (Fig. 1). They tend to be located near permanent water sources (Cohen 1992: 127; Haiman 1996: 4, Fig. 1), and, ostensibly, ancient trading routes (ibid.: 11). They lack courtyards or open features (ibid.: 3), and appear to have no clear settlement plan (Cohen 1992; Haiman 1996 contra Dever 1985). The permanency of activity at the central sites is debated (Haiman 1996 contra Dever 2014). The most well-known central settlements have been excavated and provide a relatively substantial assemblage of ceramic, lithic and architectural material. They include: Ein Ziq (Cohen 1999: 137–188), Be’er Resisim (Cohen and Dever 1978, 1979, 1981; Cohen 1999: 200–224; Dever 2014), Har Yeruham (Kochavi 1967; Cohen 1999: 111–116), and the subject of this study, Mashabe Sade (Cohen 1999: 117–130). Ephemeral sites make up the vast majority of the Intermediate Bronze Age phenomenon in the Negev Highlands (Cohen 1999), Sinai (Clamer and Sass 1977; Oren and Yekutieli 1990; Rothenberg 1999; Beit Arieh 2003) and Jordan (MacDonald 1992). They share a few characteristics, including number of structures (1–20), relatively low quality of construction, and the presence/dominance of courtyards interpreted as animal pens (Haiman 1996: 5–14). Unlike the central settlements, these sites are often located quite far from water sources; indeed, the number of sites increases in proportion to the distance from water sources (Haiman 1996). Ephemeral sites are generally assumed to relate to animal husbandry, though minor shell-working and lithic tool industries have been identified at a

Mediterranean Sea Lachish Dead Sea Hebron Hills

Har Dimon Har Zayyad Mashabe Sade Horvat Avnon Har Yeruham

Rogem Be’erotayim Ein Ziq Nahal Nizzana Be’er Resisim Khirbet Hamra Ifdan Rehkes Nafha 396

HaGamal Site ‘En Yahav 0 20 40 60 km

Figure 1 Sites mentioned in the text. Orange signifies excavated central sites, while dark green signifies excavated ephemeral sites. The larger light green circles highlight sites where Intermediate Bronze Age ingots or fragments have been discovered. Satellite photo adapted from Google Earth. 46 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

few of them (e.g., Rekhes Nafha 396; Saidel 2002; and Early Bronze/Intermediate Bronze Age HaGamal Site, Rosen 2011). Unlike the central settlements, burial cairns and tumuli are commonly found in the vicinity of ephemeral settlements (see Haiman 1992). It has been suggested that there may be a chronologic division between the central and ephemeral sites within the Intermediate Bronze Age (Dever 2014: 227). However, with the low resolution of the Intermediate Bronze Age ceramic sequence, and only a single (possibly outlying) date from secure context at an ephemeral site (HaGamal, 2858–2582 BCE, 2 σ, Rosen 2011: Table 4.1), this idea remains unsubstantiated.

Subsistence practices: the current evidence Despite a low precipitation and high evaporation environment, the modern agricultural potential in the Negev Highlands is substantial within wadi beds, especially when augmented by terracing or water-channelling systems (Evenari, Shanan and Tadmor 1971; Avni, Porat and Avni 2013: 333–334). Recent paleoclimatic studies indicate that climatic conditions during the Intermediate Bronze Age were generally better than today—i.e., even more conducive to agriculture (Migowski et al. 2004; Frumkin 2009; Langgut et al. 2015).1 The pre-modern Bedouin practice of opportunistic dry-farming in Negev wadi beds has often been used as a mirror for ancient subsistence activities of ancient Negev inhabitants (e.g., Avner 2002), though Shahack-Gross and Finkelstein (2015) recently criticized this simplistic analogy.

Animal husbandry There is limited direct evidence concerning animal husbandry at excavated Intermediate Bronze Age sites. The assumption for most studies is that animal husbandry can be determined according to site layout—the presence/dominance of open structures often interpreted as animal pens. All published faunal assemblages at central and ephemeral Intermediate Bronze Age sites in the Negev Highlands are dominated by ovicaprine remains.2 At the central sites of Be’er Resisim and Ein Ziq, meat-bearing bones of young ovicaprines dominate (Hakker-Orion 1999: 324), suggesting meat consumption was the primary use of animals (Hakker-Orion 1999: 334). In addition to the domestic species, wild desert animals (including hares, ibex and birds) appear to have supplemented the diet of central settlement inhabitants (Hakker-Orion 1999). Few remains of draft animals (i.e., donkeys, bovines) were found. At the Intermediate Bronze Age copper production site of Khirbet Hamra Ifdan animals were also primarily used for meat consumption (Muniz 2007: 245, 254–293).

1 The Intermediate Bronze Age does appear to experience two short desiccation events around 2350 BCE and 2100 BCE (the famous ‘4.2 ka dry event’) (Langgut et al. 2015). The latter may be a factor that led to eventual abandonment of the Negev Highland sites at the end of the period. 2 Note that inherent methodological problems in collection strategies hamstring a quantitative comparison of the sites. More exacting strategies at Be’er Resisim and Rogem Be’erotayim, including some fine- and wet-sieving, resulted in much finer and larger faunal assemblages (Hakker-Orion 1999: 334; Saidel et al. 2006: 214). Geoarchaeological Investigation at the Site of Mashabe Sade 47

Mashabe Sade central site

Mashabe Sade peripheral site

0 100 m N

Figure 2 Aerial photo of Mashabe Sade, overlain with site plan of the central and ephemeral sites (site plans adapted from Cohen 1999: 118, Figs. 71–72). Aerial photo adapted from Google Earth. At the smaller sites of Rogem Be’erotayim (Saidel et al. 2006: 214–216) and Rekhes Nafha 396 (Saidel 2002: 57–58) ovicaprine bones also dominate the faunal assemblages. Mature animals dominate the Rogem Be’erotayim assemblage, suggesting herding activities and the exploitation of secondary products (i.e., dairy, hair, wool) (Saidel et al. 2006: 214–215).

Dry farming Agricultural terraces support water runoff systems that supply over 300 sq km of agricultural fields in the Negev Highlands (Avni Porat and Avni 2012: 12; 2013: 332). The proximity of some Intermediate Bronze Age sites to terraced wadi beds led scholars to suggest that Intermediate Bronze Age settlements might have been actively engaged in terraced farming (Evenari, Aharoni, Shanan and Tadmor 1958). However, systematic OSL dating of Negev terraces by Avni, Porat and Avni (2012, 2013) convincingly dated the Negev agricultural systems to the Roman through Islamic periods (3rd–11th centuries CE). There is no evidence for the utilization of runoff systems or of the use of other agricultural 48 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

installations (i.e., granaries, cisterns or threshing floors) during the Intermediate Bronze Age (for more details, see Cohen and Cohen-Amin 2004; Haiman 2012). Almost no archaeobotanical research has been conducted at Intermediate Bronze Age Negev sites; only wood charcoal remains, associated with hearths or support beams, have been investigated at Be’er Resisim (Warnock 2014) and Ein Ziq (Baruch 1999). The charcoal remains from both sites are dominated by local woody plants, especially Tamarix species, Retama raetam, Zygophyllum dumosum and Chenopodiacae (Warnock 2014: 305–309; Baruch 1999: *7–*11). A passing mention of three carbonized olive pits found at Har Yeruham (Kochavi 1963: 142) is the only published Intermediate Bronze Age material directly related to edible plants. Evidence central to the prevailing assumptions concerning dry-farming during the Intermediate Bronze Age has therefore been limited to the natural potential of wadis for growing domestic cereals and small-scale horticulture (see especially Evenari, Shanan and Tadmor 1971: 109) and the presence of grinding stones and sickle blades at a few sites (e.g., arguments in Evenari, Aharoni, Shanan and Tadmor 1958; Kochavi 1967; Cohen 1999; Dever 2014). However, these are feeble sources of evidence for the practice of agriculture. Grinding stones, for example, are also used in hide and ochre processing (e.g., Dubreuil and Grosman 2009: 942–949) as well as the extraction of metals from slag (e.g., Ben-Yosef 2010: 637). Blades with sickle gloss are rare in Intermediate Bronze Age assemblages in the Negev. Multiple blades make up a single standard composite sickle, suggesting that the meagre assemblages show a maximum number of about five total harvesters at Be’er Resisim, two–six at Ein Ziq, and one–three at Rekhes Nafha 396 (Rosen and Vardi 2014: 336). In comparison, at the small, rural northern site of Sha’ar Hagolan, 39 sickle blades were found in Intermediate Bronze Age contexts (Rosen 2012: 53–58). Accordingly, the sickle blade assemblages from the Negev Highlands remain inconclusive—the inhabitants could have used them to occasionally practice small-scale opportunistic agriculture, to gather wild plants, or to process other plant material not directly related to subsistence (e.g., Clemente and Gibaja 1998: 461; Anderson 1999).

Industry Haiman (1996) was the first to draw a connection between Intermediate Bronze Age settlement patterns in the Negev and Sinai and the copper industries of Faynan. Hoards of copper ingots are characteristic of the central settlements, and an ingot fragment was also found at the ephemeral site of Rekhes Nafha 396 (Saidel 2002: 57). Slag was identified at Ein Ziq (Segal 1999: 38*–39*), but direct evidence of copper production is absent. Copper artefacts have been found at all central sites, as well as smaller settlements such as Horvat Avnon (Cohen 1999: 108, 111, Fig. 66:3), Nahal Boqer (Cohen 1999: 131–134, Fig. 82:23) and Har Dimon (Cohen 1999: 94, Fig. 53b). North of the Negev, ingots have been found in the Hebron Hills (Dever and Tadmor 1976), Lachish (Tufnell 1958: Pl. 21:11–13) and Hazor (Yahalom-Mack et al. 2014: 22, Fig. 3:1). Intermediate Bronze Age copper production sites have been identified at Khirbet Hamra Ifdan in Wadi Faynan (Adams 2000; Levy et al. 2002), >En Yahav in the Aravah Valley (Yekutieli, Shilstein and Shalev 2005; Shalev, Shilstein, and Yekutieli 2006), and possibly Sites 149 and 250 near Timna (Rothenberg Geoarchaeological Investigation at the Site of Mashabe Sade 49

1990: 6–8). Lead isotope analyses of the copper artefacts from Khirbet Hamra Ifdan, Har Yeruham, the Hebron Hills, Be’er Resisim and Ein Ziq have shown that the ingots match ores found at Faynan (Hauptmann, Schmitt-Strecker, Levy and Begemann 2015). Petrographic analysis of ceramic material from central and ephemeral sites (Goren 1996; Saidel 2002 respectively) strengthens assumptions regarding trade connections between Faynan and the Negev and also highlights broad connections to ceramic production centres throughout the southern Levant (Goren 1996: 53–68). Red Sea marine shells and Dead Sea bitumen (e.g., Nissenbaum, Serban and Connan 1999: 12*) provide additional evidence for an expansive trade system connecting the Negev Highlands to Egypt and the sedentary parts of the southern Levant. The distribution of Red Sea shells especially highlights this system, with examples found at sites inthe Jezreel Valley (e.g., Covello-Paran 2008: 63), the Negev (Be’er Resisim, Bar-Yosef 1999: 323), Jordan Valley (e.g., Bab edh-Dhra; Sowada 2009: 95, 123, 203) as well as sites along the Nile and Sinai (e.g., Ras Budran; Mumford 2012: 108–137, Fig. 29). There is lithic evidence for a local Negev industry specializing in the manufacture of marine shell beads at Rekhes Nafha 396 (Saidel 2002: 42), Rogem Be’erotayim (Saidel et al. 2006: 213, 225) and the central site of Be’er Resisim (Rosen and Vardi 2014: 332, 336–337). To summarize the evidence known before the beginning of our project, decades of archaeological investigation in the Negev Highlands have provided clues for a flourishing Intermediate Bronze Age trade system connecting the arid region to the rest of the southern Levant. However, at least some of the theories concerning food production in the Negev Highlands are far from being proven. It has been the goal of this study to investigate direct indicators of subsistence strategies, namely microscopic animal-related items and botanical remains, in order to shed light on subsistence practices at Negev Highlands Intermediate Bronze Age sites.

Investigation at Mashabe Sade The site Mashabe Sade (map ref. NIG: 1798 5431; 428 m a.s.l.; Figs. 2, 3) covers an area of 1.2 ha, with over 200 rounded, stone-built structures visible on the surface, making it one of the largest Intermediate Bronze Age settlements in the Negev (Cohen 1992: 110–112, Fig. 5). Salvage excavations were carried out at the site in 1984 (Cohen 1999: 117). Cohen cleared 21 structures built of local fieldstones directly on bedrock. Each room had one or two central drum-built pillars supporting stone roofing slabs; walls were preserved to an approximate height of 0.6–1.2m. Intermediate Bronze Age sherds and vessels were found in most loci, including complete bowls, storage jars and amphorae. The finds also included a copper dagger, a fragment of an ingot and grinding stones. A thin layer of gray sediment above bedrock was found inside a few excavated rooms (Cohen 1999: 118). Faunal remains were few and inconclusive (Hakker-Orion 1999: 327). No open enclosures or central courtyards were identified. A smaller site (NIG: 1800 5432; 431 m a.s.l. Figs. 2, 4) was surveyed on the ridge immediately east of the central settlement (Fig. 2). This site, which had not been excavated, 50 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

includes six rooms encircling an oval central courtyard. Additionally small circular structures, some with open enclosures, surround this central structure. Based on surface finds Cohen (1999: 117, Fig. 71) dated it to the Intermediate Bronze Age. Hence, though its layout is not common in the Negev during the Intermediate Bronze Age, we contemplated the possibility that it may represent the phenomenon of ephemeral Intermediate Bronze Age sites in the region.

Research strategy The research strategy for this study continues the framework of the Negev Highlands Geoarchaeological Research Project. Previous work was conducted by two of the authors (Ruth Shahack-Gross and Israel Finkelstein) at Iron Age and Byzantine–Early Islamic sites in the Negev Highlands (Shahack-Gross and Finkelstein 2008, 2015; Shahack-Gross et al. 2014). These studies stemmed from a growing corpus of geoarchaeological and ethnoarchaeological material from pastoral and agro-pastoral settlements that emphasize the importance of livestock dung as a key material to reconstruct ancient subsistence practices (e.g., Kramer 1982; Brochier, Villa and Giacomarra 1992; Macphail, Courty, Hartner and Wattez 1997; Shahack-Gross, Marshall and Weiner 2003; Valamoti and Charles 2005; Tsartsidou, Lev-Yadun, Efstratiou and Weiner 2008). If sufficiently preserved and identified at archaeological sites, ancient livestock dung supplies a direct line of evidence for herding practices, grazing and foddering strategies, and even seasonality of settlements (Shahack-Gross 2011: 214–216). Previous work focused on small-scale excavations with systematic retrieval of macroscopic and microscopic remains through sieving, hand collection of macroscopic remains, and sampling of sediments and charred botanical remains. Aided by modern reference materials collected locally, the aforementioned Negev studies have shown that subsistence strategies are reflected in the phytolith assemblage within dung remains. Concerning the Intermediate Bronze Age, dung spherulites have also been cursorily observed in archaeological sediments at Intermediate Bronze Age ephemeral sites (e.g., Rekhes Nafha 396, Saidel 2002: 57; HaGamal, R. Shahack-Gross. personal observation). However, no methodological study has ever been attempted at Intermediate Bronze Age sites. Thus, the study reported here aims to identify the remains of degraded livestock dung, extract phytoliths from such deposits and study them systematically. Shahack-Gross et al. (2014) have shown that the important indicators of livestock dung, phytoliths and spherulites are well preserved in arid regions like the Negev Highlands, owing to low soil moisture and neutral to slightly alkaline sediment pH. Therefore, there is no risk of significant post- depositional changes to ancient phytolith assemblages from historical Negev Highlands dung deposits, and phytolith assemblages will reflect grazing and foddering practices of herd animals and thus distinguish pastoral from agro-pastoral practices.

Materials and methods Two short, limited in scope, excavation seasons (January and March 2013) were carried out by the authors at Mashabe Sade. We later returned to the site to conduct an OSL dating study of the sediments (Junge et al. forthcoming). At the central site, probes were opened within four unexcavated structures, two previously excavated Geoarchaeological Investigation at the Site of Mashabe Sade 51 structures (Cohen’s Locus 23 and Locus 24), and five 1×1 m squares in open spaces between the round structures across the site (see Fig. 3). At the smaller site on the nearby ridge two trenches were opened in opposite rooms and three 1×1 m units opened within the central courtyard (Fig. 4). Effort at both sites was expended to obtain a representative sample of sediments and the microremains within, both vertically (via their deposition/formation history) and horizontally (via the spread of the settlement, use of open areas, etc.). All units were excavated to bedrock to provide a complete stratigraphic sequence. Following observations at other Negev sites (see Shahack-Gross et al. 2014: 112–113), gray sediments were sieved systematically through 1 mm screens. All macroscopic remains were collected. Sections were photographed, and bulk sediment samples of ca. 10 g each were collected into plastic vials, according to macroscopic differences in colour and texture in a manner that represented the full stratigraphic sequence. Control samples (i.e., undisturbed loess and alluvial sediments beyond the limit of the sites, from neighbouring ridges and wadis) were also collected for comparison.

L.07 N L.08

L.06 L.03 L.02 L.04 L.01 A L.05/L.16 A L.21 0 10 m

L.17 L.13

B L.15

L.12

B 0 10 m

Figure 3 Plan of central site of Mashabe Sade, with excavated loci marked. Blue represents loci opened within structures; red, loci excavated in open areas. Dotted loci were excavated but not analyzed due to poor preservation (plan adapted from Cohen 1999: 118, Fig. 71). 52 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

L.10 107 106 Courtyard 100 L.20 100 L.19 L.09 102 L.11 104

103 L.18

0 10 m

Figure 4 Plan of the ephemeral site, with excavated loci marked. Colours follow Fig. 3 (plan adapted from Cohen 1999: 118, Fig. 72).

Sediment samples were analyzed in order to determine their composition and distinguish between sources, e.g., between geogenic and anthropogenic sources, detection of dung remains, and distinguish domestic from industrial ash. The following methods were used to address these issues: • Mineralogical analysis was performed using a Fourier Transform Infrared (FTIR) spectrometer (Nicolet 380, Thermo Electron Corp), and collected between 4000 and 250 cm-1. Samples were prepared following the KBr standard procedure (see Weiner 2010 for details), and compared with the extensive Kimmel Center for Archaeological Science (Weizmann Institute) reference library and experimental data (Berna et al. 2007; Regev et al. 2010; Weiner 2010: 275–326). • Four samples were subjected to elemental analysis by Energy-Dispersive X-Ray Fluorescence (ED-XRF) using a Spectro-XEPOS instrument. Samples for XRF were prepared following standard procedures described in Eliyahu-Behar et al. (2012). • Samples for phytolith analysis were prepared according to the rapid extraction method developed by Katz et al. (2010). Phytoliths were counted at 200× under plane polarized light (PPL); uncertain identifications were checked under crossed polarized light (XPL). Phytolith morphologies were preliminarily examined at 400× and compared to reference collections available at the laboratory of Shahack-Gross. Due to the low concentrations (see results below) of phytoliths found in the archaeological sediments at Mashabe Sade, phytolith morphologies could not be studied quantitatively. Qualitative descriptions of phytolith morphologies are given in the discussion. Geoarchaeological Investigation at the Site of Mashabe Sade 53

• In order to determine whether ash was produced from wood and/or dung fuels, bulk sediments were prepared for dung spherulite and ash pseudomorph analysis (PSR analysis) according to the procedure developed by Gur-Arieh, Mintz, Boaretto and Shahack-Gross (2013). Counting was performed at 400× under PPL for ash pseudomorphs and under XPL for dung spherulites, using a Nikon 50i POL polarizing light microscope. In the results that follow, all micro-remain concentrations are presented with the standard analytical error of ± 30% (see Albert and Weiner 2001; Gur-Arieh, Mintz, Boaretto and Shahack-Gross 2013 for more details). Results Macroarchaeological results Central settlement Excavation exposed parts of four previously unexcavated structures (Structures 1, 2, 3 and 5), as well as reinvestigated two structures partially excavated by Cohen (Structures 4 and 6, Cohen’s Locus 23 and Locus 24 respectively). Previous reconstructions of these structures remain sound: all walls were built directly on bedrock and roofed by thin local limestone slabs. Large central pillar stones (ca. 60 × 35 × 40 cm) were uncovered in Structure 1 (Loci 005/016) and Structure 6 (Locus 015), and a drum-built pillar was identified in Structures 4 (Locus 012) and 5 (Locus 013). Notably, all excavated areas had no hearths or other anthropogenic installations (e.g., worktables, kilns, cup-marks, etc.). No enclosed courtyards were identified. A common macroscopic stratigraphy was noted in the excavation of Structures 1, 4, 5 and 6. From bottom to top it included: 1. A patchy, thin (ca. 0.5–2 cm thick) layer of light gray-brown hard sediment, directly on bedrock, associated with flat-lying pottery and elevated concentrations of microremains (below). It is thus identified as a living surface. 2. Gray-brown sediments (ca. 10–30 cm thick). 3. Stone collapse, with soft yellow-brown loess sediment, roots and abundant land snail shells (ca.10–45 cm thick). In Structure 1 we identified an accumulation of gray-coloured ash (ca.15 cm thick) overlying both the bedrock and the patchy ‘living surface’ identified immediately above bedrock (Locus 021). This accumulation was rich in burnt pottery, and contained a large (ca. 5 × 5 cm) pink marine shell. This material was fully sieved. No charred material suitable for dating was found. In order to test the fuel sources of this ash, a PSR analysis was conducted, and in order to test whether it might contain copper due to activities relating to metallurgy, it was also subjected to XRF analysis (details below). Open spaces were characteristically different. The accumulation in Open Space 1 (Locus 004, 1 × 1 m) was very shallow (ca. 15 cm thick), and yielded only yellow-brown silty loess. In Open Space 2 (Locus 003, 1.5 m × 1 m; ca. 5 cm thick), a gray-brown sediment arguably contemporary with the settlement sloped up towards the northern wall of Structure 1, and was sealed by ca. 15 cm of yellow-brown loess. Unlike the other two 54 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

a)

b)

Figure 5 (a) Structure 5 at central site, north section. Note drum-built pillar outline in gray. White dotted line represents approximate division between hard brown-gray and mixed material above. (b) Excavation of L.009 in the courtyard of the ephemeral site, facing south. The dotted line highlights the approximate accumulation of degraded animal dung in the excavated area. Geoarchaeological Investigation at the Site of Mashabe Sade 55 loci, the deeper Open Space 3 (Locus 017, 1 × 1 m) showed little evidence of the yellow brown loess, and instead showed a stratigraphic sequence of a mixed stone collapse (from the structure immediately to the west), and a heavily bioturbated archaeological context. No obvious activity areas could be reconstructed macroscopically. Pottery sherds were infrequent. No faunal remains were found in any context. Though the surface of the central site was littered with lithic debris, worked tools as well as grinding stones were absent within our excavated areas. Artefacts related to trading systems were scant. A marine shell was found in Structure 1 (Locus 005), and a core of a worked Rea Sea shell (Lambis truncata sebae, H.K. Mienis, personal communication) was discovered lying on a thin gray accumulation directly above bedrock in Structure 6 (Locus 015). A single copper prill was identified in the gray-brown accumulation just above bedrock in Structure 5 (Locus 013). Additionally, a hammerstone was identified in Open Space 3 (Locus 017).

The site on the nearby ridge At the round site on the nearby ridge, suspected of representing the phenomenon of ephemeral Intermediate Bronze Age sites in the region, two rooms (Rooms 102 and 106) and the central courtyard (Courtyard 100) were excavated. The circular structure was preserved down to ca. 80 cm below surface, with walls preserved up to 3 courses. No evidence of roofing was found. Room 106 was built with large fieldstones directly on bedrock. A ca. 5–10 cm thick loess appeared to fill the entire structure. Two architectural phases were identified within Room 102. The architectural plan of the upper phase (Phase 2) is visible on the surface, and corresponds to an accumulation of yellow loess (ca. 15 cm), covering a thin (ca. 0.5 cm) gray-brown layer of anthropogenic sediment resting directly above bedrock. A few Intermediate Bronze Age sherds were found within this material. The lower phase (Phase 1) is constrained to the southeastern part of our probe (Locus 018), and appears as a collapsed structure below the southernmost wall of Room 102. The extent of this phase is not fully understood—due to the eastward collapse of the architectural feature of this earlier phase, and the minor (only 1 × 2 m) horizontal exposure—but it is nearly perpendicular to the orientation of the western walls of Room 102. Sealed below this collapse was light gray sediment with large sherds from a typical Intermediate Bronze Age ovoid band-combed storage jar. Three small 1 × 1 m probes (Loci 009, 019 and 020) were opened in Courtyard 100. Loci 019 and 020 were shallow and were made up of only yellow loessial sediments that accumulated on bedrock. However, Locus 009 had two macroscopic layers, the local yellow loessial sediments sealing an accumulation of gray-brown sediments ca. 10 cm thick (Fig. 5b).

Microarchaeological results Control sediments (n=4) Four yellow-brown sediment samples were collected around the sites, two from adjacent ridges and two from wadis. They are composed of calcite, clay and quartz. One sample 56 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

6 a)

5

4.6

4

3

2.4 2.2 2

1.3 1.3

1 Phytolith concentrations (millions/g of of sediment) (millions/g concentrations Phytolith 0.7 0.4 0.4 0.4 0.4 0.4 0.3 0.3 0.3 0.19 0.09 0.02 0.04 0.03 0.02 0.05 0.02

0 0.02 0

0.5 b)

0.3

0.28 0.26

0.11 0.1 0.11 0.11 0.09 0.09 0.04 0.05 0 0 0 0 0 0 0 0 0 0 0 0 0 0 Gray Red Brown on Bedrock on Bedrock on Bedrock on Bedrock on Bedrock on Bedrock on Bedrock Gray Brown Gray Brown Gray Brown Gray Brown Gray Brown Gray Brown Gray Brown Gray Brown Yellow Gray Yellow Controls Yellow Brown Yellow Brown Yellow Brown Yellow Brown Yellow Brown Yellow Brown Yellow Open Open Open Structure Structure Structure Structure

Dung spherulite concentratons (millions/g of sediment) Space Space Space 1 4 5 6 1 2 3

Figure 6 Microremain results from the central site. (a) phytolith concentrations; (b) dung spherulite concentrations.

also included a minor component of gypsum. Parallel to other studies at arid sites, microremains (phytoliths, dung spherulites and ash pseudomorphs) were absent or almost completely absent in these control samples, with concentrations being far lower than 1 million microremains/g of sediment (Table 1; compare to Shahack-Gross and Finkelstein 2008, Shahack-Gross et al. 2014). The presence of a very low concentration of dung spherulites in one sample is most likely from modern animal activity; indeed modern camel and sheep/goat dung was noted on the surface of the surrounding ridges, and in both the central and ephemeral sites. No ash pseudomorphs were identified in any of the Geoarchaeological Investigation at the Site of Mashabe Sade 57 controls. An anthropogenic signal is expected to be at least one order of magnitude higher than values found in the controls, considering the large analytical error, i.e., values above 1 million particles of any microremain in 1 g of sediment are significantly different from the controls and indicate anthropogenic input.

Central settlement

Structures Of the six structures excavated, four were sufficiently preserved to sample according to macroscopic differences in the sections (Structures 1, 4, 5 and 6; Table 2). The mineralogical and microremain results are as follows: 1. Light brown-gray sediment just above bedrock (Structures 1, 4, 5, 6) (n=5). These sediments, macroscopically identified as the living surfaces within the structures, are composed of calcite, clay, quartz and gypsum. Most of them include an order of magnitude higher concentration of phytoliths relative to the sediments above them, although these values are within the range of the controls (Fig. 6a). A few ash pseudomorphs were identified in samples from Structures 1 and 4. One sample from Locus 005 includes rather high concentrations of phytoliths and ash pseudomorphs. This sample probably represents infiltration (via bioturbation?) of ash material found above it. Clays or calcite in these samples are not heat- altered (as deduced from infrared spectra) thus, at least in the areas excavated, we have no evidence of fire being used inside the structures. The markedly low concentrations of dung spherulites in the living surface samples (0.00–0.11 million/g of sediment) argue against the identification of accumulations of dung within structures at the site. 2. Ash accumulation (Structure 1 only) (n=4). Mineralogically, this deposit is composed mainly of calcite and unaltered clay (i.e., clay minerals exposed to temperatures lower than 500 °C). Gypsum was present in three of the four samples analyzed, observed micro- and macroscopically as small white nodules within the material (ca. 0.1–1.0 mm). Concentrations of phytoliths in this gray sediment are high relative to controls (0.3–2.4 million/g of sediment); concentrations of dung spherulites are higher than controls but low relative to fresh dung (0.09–0.26 million/g of sediment), and ash pseudomorphs appear in concentrations that are within the range of burning of dung (0.09–0.31 million/g of sediment) (Table 2; compare to Shahack-Gross and Finkelstein 2008: Table 2; Gur-Arieh et al. 2014: Appendix 1). PSR values (1–3) indicate that this sediment is likely composed of both burnt wood and dung (Gur-Arieh et al. 2014: 49, Fig. 6). Note that the clay component is unaltered, possibly indicating that the clay filtrated into the gray material after its formation by combustion. The unstratified appearance of the sediment makes it difficult to determine whether the material is in primary (formed in situ) or secondary (ashes formed elsewhere and swept in) deposition; however, the fact that clay in the living surface below was also unaltered suggests the latter. Additionally, the presence of gypsum within this material may relate either to the burning of tamarisk wood (Shahack-Gross and Finkelstein 2008: 975) and/or to post-depositional pedogenic 58 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

gypsum. Based on the macro- and microremains found in this ash deposit, this is the strongest microarchaeological signal of anthropogenic sediments we encountered at the central site. Previous scholars have postulated that ash deposits at central sites may represent copper production activities (Haiman 1996: 18–20). In order to evaluate this hypothesis, this ashy material (n=2) was subjected to XRF analysis along with two control samples. The elements were analyzed primarily for copper content, but also for by-products of copper production (iron, manganese, lead, calcium, etc.). The levels of copper are (1) indistinguishable from controls considering analytical error, and (2) far below reported concentrations within sediments proven to be related to metallurgical activities (ca. 100–300 ppm at Tell es-Safi/Gath, Eliyahu-Behar et al. 2012; 300–600 ppm at Tel Dor, Eliyahu-Behar et al. 2008) (Table 3). Thus though the activity that resulted in this ash accumulation is uncertain, the formation of this material is not likely related to metallurgical activities. 3. Brown-gray sediment above the living surface (Structures 5, 6) (n=2). These sediments are composed of calcite, clay, quartz and gypsum. Microremain concentrations are very low, similar to the ranges in control sediments (Fig. 6). The source of gypsum in this material is unclear; it may indicate an original composition of sediment brought onto the site (construction material such as clay weatherproofing?), or post-depositional pedogenic precipitation of gypsum (cf. situation at Tel Brak, Matthews et al. 1997). 4. Yellow-brown loessial sediments overlying the collapse (Structures 1, 5, 6) (n=7). These yellow sediments are generally composed of calcite, clay and quartz, similar to the controls. No gypsum was identified in any of these samples. The microremain concentrations noted in these sediments are very low (phytoliths: 0.02–0.4 millions/g of sediment; dung spherulites: 0.0–0.11 million/g of sediment), similar to the range in control sediments.

Open areas Despite macroscopic differences in appearance of various sediments in open spaces, all samples from open areas are composed of calcite, clay and quartz, similar to the controls. Somewhat increased phytolith concentrations were found in the lower gray- brown sediment found directly on bedrock in Open Spaces 1 and 3 (Fig. 6a), possibly representing human activity in the open areas. These signals seem to be preserved in part by the collapse of stones above. Neither ash pseudomorphs nor heat-altered clay or calcite were noted in the samples, so at this time there is no evidence of use of fire outside of the structures. Dung spherulites were absent in Open Space 1 and present at very low concentrations—similar to those in controls—in Open Spaces 2 and 3 (Fig. 6b). As modern animal dung is visible on the surface of the central site, this may be the result of post-abandonment mixing. Thus, we cannot attribute this to Intermediate Bronze Age activity at the site. Geoarchaeological Investigation at the Site of Mashabe Sade 59 PSR - - - - ) † Ash pseudomorph concentration (millions/g) ( 00 0 0 0 0 0 0 ) † 3 0.3 ± 0.1 Dung spherulite concentration (millions/g) ( ) † 1

Phytolith concentration (millions/g) ( 0.4 ± 0.1 120.09 ± 0.03 3 0 0 able T Ca (ua?), Cl (ua), Q, Org Ca (ua?), Cl (ua), Q, Org Sample Mineralogy description Yellow-brown ZMS-12.008 Yellow-brown ZMS-12.009 Mineralogical results, microremain concentrations and PSR (pseudomorphs to spherulite ratios) from control sediments* Sample description Sediment Wadi control (n=2)Wadi Yellow-brownRidge control (n=2) ZMS-12.010 Yellow-brown Ca (ua), Cl Q, Org 0.028 ± 0.008 ZMS-12.015 Ca (ua), Cl Q, Org 1 0.024 ± 0.007 1 0 0 0 0 * Abbreviations: Ca = calcite; Cl = clay; Q = quartz; G = gypsum, Org = organics. (a) = altered; (sa) slightly (ua) unaltered. Microremain concentrations are reported in = organics. Abbreviations: Ca = calcite; Cl clay; Q quartz; G gypsum, Org * millions per g of sediment. (†) signifies actual number microremains counted sample. 60 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross PSR ) †

Ash pseudomorph concentration (millions/g) ( 0.0000.000 00.000 - 00.09 ± 0.03 - 0 10.27 ± 0.08 - 1 30.31 ± 0.09 3 60.31 ± 0.09 1.2 60.000 3 00.25 ± 0.07 5 - 5 Dung spherulite concentration (millions/g) ( † ) ) † 2

Phytolith concentration (millions/g) ( 4.6 ± 1.4 247 0.05 ± 0.01 1 able T Minerals CHAP? of sediments from the central site* (m asl) L.001 419.04L.001 Ca (ua), Cl Q, Org 0.4 ± 0.1 419.04S. Section 418.88 Ca (ua), Cl Q, Org 0.021 ± 0.006 Ca (ua), Cl Q, Org 17 1 0.05 ± 0.02L.005 2 ± 0.03 0.11 0.000 1 418.67L.016 0.000 Ca (a), Cl (ua), Q 0 418.66L.016 Ca (a), Cl (ua), Q, G 0 2.4 ± 0.7 418.65 1.3 ± 0.4 Ca (a), Cl (ua), Q 79 56 2.2 ± 0.7 0.09 ± 0.03 1 0.26 ± 0.08 101 5 0.10 ± 0.03 2 L.005 418.68 Ca (a), Cl (ua), Q, G 0.30 ± 0.09 14L.021 0.09 ± 0.03 1 418.63 Ca (a), Cl (ua), Q, GL.021 0.09 ± 0.03 418.63 Ca (a), Cl (ua), Q, G, 4 0.000 0 Sample Locus Height ZMS- 13.041a ZMS- 13.041b ZMS- 13.184 ZMS- 13.048 ZMS- 13.189 ZMS- 13.188 ZMS- 13.047 ZMS- 13.050 ZMS- 13.051 description Yellow-brown Yellow-brown (n=3) Gray friable (n=4) Light brown directly on bedrock (n=1) Dark gray- Brown friable directly on bedrock (n=1) Mineralogical results, microremain concentrations and PSR (pseudomorphs to spherulite ratios) Structure Sediment Structures Structure 1 (L.001, L.005, L.016, L.021) * Abbreviations: Ca = calcite; Cl = clay; Q = quartz; G = gypsum, Org = organics. (a) = altered; (sa) slightly (ua) unaltered. Microremain concentrations are reported in = organics. Abbreviations: Ca = calcite; Cl clay; Q quartz; G gypsum, Org * millions per g of sediment. (†) signifies actual number microremains counted sample. Geoarchaeological Investigation at the Site of Mashabe Sade 61 PSR ) † Ash pseudomorph concentration (millions/g) ( 0.0000.05 ± 0.01 0 1 - - 0.0000.000 00.000 - 0 - 00.000 - 0.000 00.000 - 00.000 - 0 - 0 - Dung spherulite concentration (millions/g) ( † ) ) † Phytolith concentration (millions/g) ( 0.4 ± 0.1 21 ± 0.03 0.11 2 0.025 ± 0.008 10.3 ± 0.1 0.000 16 0 0.000 0 Minerals Ca (sa?), Cl (ua), Q, G Ca (sa?), Cl (ua), Q, G Ca (sa?), Cl (ua), Q, G (m asl) L.012 424.81 Ca (ua), Cl Q, G 0.020 ± 0.006 1 0.000 0 L.015 423.70 Ca (ua), Cl Q 0.05 ± 0.01 2 0.000 0 L.012 424.75 Ca (ua), Cl Q, GL.013 0.4 ± 0.1 422.95L.013 13 Ca (ua), Cl Q 422.54L.013 Ca (ua), Cl Q, G 0.000 0.04 ± 0.01 422.36 0.000 2 0 L.015 0.000 0 423.81 Ca (ua), Cl QL.015 0 0.000 0.031 ± 0.009 423.58 1L.015 0 423.55 0.000 0 Sample Locus Height ZMS- 13.099 ZMS- 13.176 ZMS- 13.104 ZMS- 13.169 ZMS- 13.168 ZMS- 13.167 ZMS- 13.177 ZMS- 13.175 ZMS- 13.174 description Yellow-brown Yellow-brown (n=1) Yellow-brown Yellow-brown (n=1) Yellow-brown (n=2) Hard brown- Gray directly on bedrock (n=1) Hard gray- Brown (n=1) Hard gray- Brown directly on bedrock (n=1) Hard gray- Brown (n=1) Light gray directly on bedrock (n=1) Structure Sediment Structure 4 (L.012) = Cohen L.23 Structure 5 (L.013) Structure 6 (L.015) = Cohen L.24 62 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross PSR ) † Ash pseudomorph concentration (millions/g) ( 0.0000.000 00.000 - 00.000 - 00.000 - 00.000 - 0 - 00.000 - 00.000 - 0 - Dung spherulite concentration (millions/g) ( † ) ) † Phytolith concentration (millions/g) ( Minerals (m asl) L.003 418.94L.003 Ca (ua), Cl Q 418.85 Ca (ua), Cl Q 0.05 ± 0.02L.003 2 0.05 ± 0.01 418.74 2 Ca (ua), Cl Q, G ± 0.03 0.11 1.3 ± 0.4 1 0.000 54 0 0.000 0 L.003 418.74 Ca (ua), Cl QL.004 0.4 ± 0.1 418.35L.004 Ca (ua), Cl Q 19 418.14 Ca (ua), Cl QL.017 0.027 ± 0.008 0.000 2 421.88 0.19 ± 0.06 Ca (ua), Cl (ua) , Q 0 0.000 19L.017 0.3 ± 0.1 421.75 0.04 ± 0.01 0 Ca (ua), Cl Q 1 13 0.7 ± 0.2 0.09 ± 0.03 2 30 0.28 ± 0.08 6 Sample Locus Height ZMS- 13.044 ZMS- 13.045 ZMS- 13.063 ZMS- 13.064 ZMS- 13.029 ZMS- 13.028 ZMS- 13.161 ZMS- 13.160 description Yellow-brown Yellow-brown (n=2) Yellow-brown Yellow-brown (n=1) Red-gray- brown sediment (n=1) Gray directly on bedrock (n=2) Yellow-gray directly on bedrock (n=1) Dark gray-brown sediment on bedrock (n=1) Structure Sediment Open spaces Open space 1 (L.003) Open space 2 (L.004) Open space 3 (L.017) Geoarchaeological Investigation at the Site of Mashabe Sade 63 Analytical Total Pb (ppm) 6 57.27 Sn (ppm) 11 Zn (ppm) Cu (ppm) 3 O 2 MnO Fe 2 3

TiO able T O CaO 2 SK 5 O 2 P 2 XRF results from Structure 1 (L.005/L.016) In the central site SiO 3 O 2 Al Method Powder 8.7 34.76 0.31 0.12 0.99 19.53 0.68 0.05 2.68 17 924 12 23 68.7 Powder 4.38 13.3 0.62 2.88 0.98 27.63 0.48 0.04 2.14 23 49 13 6 57.9 Powder 9.35 32.05 0.34 0.13 0.98 21.95 0.9 0.06 3.25 23 68 14 28 69.2 Powder 4.4 18.04 0.7 0.92 0.59 29.41 0.48 0.04 2.06 40 50 Sample GSS-1 Powder 17.27 69.05 0.21 0.06 2.41 1.75 0.8 0.22 4.72 24 763 14 91 86.7 ZMS-10 (Wadi ZMS-10 (Wadi Control) ZMS-189 (L.005/L.016) ZMS-15 (Ridge Control) ZMS-188 (L.005/L.016) 64 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

Site on the nearby ridge Sediment accumulation in Room 106 was ca. 10–20 cm thick, sloping northeast. Though a slight macroscopic division was visible between layers in Room 106, no mineralogical division was discernible, all sediments being composed of calcite, clay and quartz, similar to local loessial sediments. Concentrations of phytoliths and dung spherulites are in the range of local loess, and no ash pseudomorphs were detected (Fig. 7). Room 102 showed evidence of two phases: Phase 1 (Locus 018), limited to the western part of the room and sealed by collapse and an organic-rich layer, and Phase 2 (Locus 011), directly above the collapse. Both phases were subdivided into two depositional events, with yellow loessial sediment overlying gray sediment. In the upper Phase 2 (Locus 011), the yellow sediments are similar mineralogically and in their microremain contents to local loess controls (Table 4; Fig. 7), while the gray material above the collapse contains slightly elevated microremain concentrations (0.3±0.1 million/g of sediment; Fig. 7). The lower phase (Locus 018) showed a gradation between its two layers: the upper yellow layer, again similar mineralogically to the yellow loess, had relatively high phytolith concentration (0.80.3± million/g of sediment) and the light gray sediment below it had very high concentrations of both phytoliths (3.81.3± million/g of sediment) and dung spherulites (1.40.4± million/g of sediment) (Table 4, Fig. 7). This material is confidently identified as degraded herbivore dung. Similar concentrations were identified in degraded dung contexts at the Byzantine/Early Islamic site of Wadi el-Mustayer (Shahack-Gross et al. 2014: Table 3). This degraded dung sediment was sampled for direct OSL dating (see below). Three sediment layers were noticed in one of the sections (Locus 009) excavated within Courtyard 100. The sediment throughout the stratigraphic profile is mineralogically similar to loess, composed of calcite, clay and quartz, yet it includes elevated concentrations of both phytoliths and dung spherulites, in one sample in its lowermost part, indicating presence of degraded herbivore dung (Table 4; Fig. 7).

Discussion There is little evidence of daily activities at Mashabe Sade. Hearths were not identified in our excavation, though previous excavations at the site showed evidence of fire within structures (Cohen 1999: 117–126). At the moment there is little evidence of activity outside of structures. The question of permanent settlement at Mashabe Sade remains unclear. On one hand, Mashabe Sade and other large central sites are well-built; obviously, significant effort was invested to construct these structures. Note that the site is built on a well-bedded limestone formation that naturally weathers into building blocks of variable sizes. This makes construction somewhat less intensive relative to building with quarried stone. The only evidence for sustained (or repeated) settlement identified in our excavation is the repurposing of Structure 1 with the disposal of secondary ash (Locus 005/Locus 016). This repurposing appears common at Mashabe Sade (likely Locus 20 and Locus 26 in Cohen 1999: 120–123) and Be’er Resisim (7C and 7E, Cohen and Dever 1979: 48; Dever 2014: 169), and does suggest some longer-term developments and possibly maintenance activities. Geoarchaeological Investigation at the Site of Mashabe Sade 65

5 a)

4 3.8

3

2 2.1

1.6

1 1.0 0.9 0.8 0.7 0.4 0.3 0.5 0.19 Phytolith concentration (million/g of of sediment) (million/g concentration Phytolith 0.23 0.3 0.02 0.06 0

3 b)

2.6

2

1.4

1

0.7 0.5 0.3 0.28 0.3 0.4 0.4 0.1 0.05 0.04 0 0 0 0 Yellow Yellow Yellow Yellow Yellow Gray Yellow Red Gray Brown Gray on Brown Gray on Brown Brown on Brown Brown Brown on Bedrock Bedrock Bedrock Bedrock Dung spherulite concentration (million/g of of sediment) (million/g concentration spherulite Dung L.010 L.011 L.018 L.009

Controls Room 106 Room 102 Courtyard 100

Figure 7 Microremain concentrations from the ephemeral site: (a) phytolith concentrations; (b) dung spherulite concentrations. 66 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross PSR ) † 0.00 0 - Ash Pseudomorph concentration (millions/g) ( 0 0.00 0 - 0 0.00 029 - ) † Dung Spherulite concentration (millions/g) ( 1.5 ± 0.4 57 0 3.8 ± 1.5 174 1.4 ± 0.4 Phytolith concentration (millions/g) 4

Org, G? Org, Ca (sa?), Cl (ua), Q able T the ephemeral site* 415.30 Ca (ua), Cl Q, overlying L.018 ZMS-13.139 L.011ZMS-13.136 L.011 415.37ZMS-13.085 directly L.011, 415.02 Ca (ua), Cl Q 0.23 ± 0.07 10 Ca (ua), Cl Q 0.3 ± 0.1 0 13 0.10 ± 0.03 2 0.00 0 - ZMS-13.148 L.018ZMS-13.147 L.018 414.99 414.81 Ca (ua), Cl Q 0.8 ± 0.3 32 0.3 ± 0.1 7 0.00 0 - ZMS-13.132 L.010ZMS-13.129 L.010ZMS-13.130 416.50 L.010 416.41 Ca (ua), Cl Q 416.37 0.3 ± 0.1 Ca (ua), Cl Q 15 0.19 ± 0.06 Ca (ua), Cl Q 6 0.05 ± 0.01 0.06 ± 0.02 3 1 0.04 ± 0.01 0.00 1 0.28 ± 0.09 0 0.00 6 0.00 0 0 - - - Sample Locus Height (m asl) Minerals Yellow brown Yellow (n=1) Yellow-gray directly on bedrock (n=1) Dark brown material organic (n=1) Yellow-brown Yellow-brown (n=1) Light gray on bedrock (n=1) Hard yellow- brown (n=1) Yellow-brown- gray directly on bedrock (n=2) Sediment description Courtyard Room 102 Rooms Room 106 Mineralogical results, microremain concentrations and PSR (pseudomorphs to spherulite ratios) of sediments from Geoarchaeological Investigation at the Site of Mashabe Sade 67 PSR ) † 0.00 0 - 0.00 0 - Ash Pseudomorph concentration (millions/g) ( 16 0 0.00 031 - ) † Dung Spherulite concentration (millions/g) ( 2.1 ± 0.6 101 0.7 ± 0.2 1.0 ± 0.3 391.6 ± 0.5 0 0.7 ± 0.2 640.9 ± 0.3 26 0.5 ± 0.11.6 ± 0.5 32 0.7 ± 0.2 5 62 0.4 ± 0.1 0.00 8 2.6 ± 0.8 0.00 0 5 0.00 0 0 - - - Phytolith concentration (millions/g) Ca (sa?), Cl (ua), Q Ca (sa?), Cl (ua), Q Ca (sa?), Cl (ua), Q Ca (sa?), Cl (ua) , Q Ca (sa?), Cl (ua), Q Ca (sa?), Cl (ua), Q ZMS-12.069b L.009 415.26 ZMS-12.069 L.009 415.26 ZMS-12.073 L.009ZMS-12.072 L.009 415.67 415.54 Ca (ua), Cl Q 0.5 ± 0.2 20 0.4 ± 0.1 2 0.00 0 - ZMS-12.071 L.009ZMS-12.070 L.009 415.35 415.28 ZMS-12.074 L.009 415.36 Sample Locus Height (m asl) Minerals Hard gray- brown on bedrock (n=2) Yellow-brown Yellow-brown (n=1) Red-brown (n=1) Hard gray- brown (n=3) Sediment description Courtyard 100 * Abbreviations: Ca = calcite; Cl = clay; Q = quartz; G = gypsum, Org = organics. (a) = altered; (sa) slightly (ua) unaltered. Microremain concentrations are reported in = organics. Abbreviations: Ca = calcite; Cl clay; Q quartz; G gypsum, Org * millions per g of sediment. (†) signifies actual number microremains counted sample. 68 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

The scarcity of macroscopic finds within a large settlement such as Mashabe Sade is puzzling. Animal bones, grinding implements and lithic tools were not found in the excavated areas. Charred botanical remains such as wood or seeds were absent as well. The latter made it impossible to conduct accurate radiometric dating. Macroscopic artefacts collected from the nearby site were even fewer. The scarcity of macroscopic artefacts that can shed light on subsistence economy highlights the need to search for microremains indicative of subsistence practices, such as dung deposits that point to herding and cereal phytoliths that designate agriculture.

Animal husbandry and agriculture The central site Dung deposits were not identified at the central site, neither in open spaces nor within structures. Low amounts of dung mixed with large amounts of wood ash were identified in an ash deposit within Structure 1 (Locus 005/Locus 016). This may be connected to the use of dung as a minor component of fuel (e.g., Kramer 1982; Miller 1984). The very low concentrations of phytoliths, with notable dominance of phytoliths originating from shrubs (e.g., ellipsoids and irregular morphotypes), indicate that grasses in general and cereals in particular are absent in the sediments studied at Mashabe Sade. In summary, macro- and microarchaeologically, no animal pens were identified, no faunal remains recovered, and sickle blades are notably absent in the lithic remains. At this time, there is no evidence for either animal husbandry or cultivation at the central site of Mashabe Sade.

The nearby site Degraded dung deposits were identified in Courtyard 100 (Locus 009) as well as in the lower phase of Room 102 (Locus 018). In both contexts, the dung signal is illustrated by relatively high phytolith concentrations. Dung spherulite concentrations are similar to concentrations at Wadi el-Mustayer, though lower than the Iron IIA sites of Atar Haroa and Nahal Boqer. Identification of the degraded dung accumulations at this site demonstrates that animal husbandry was practiced there at some point—or points—in time. Qualitative phytolith morphotype analysis from the degraded dung from Courtyard 100 (Locus 009) and Room 102 (Locus 018) suggests a mixed diet dominated by Negev woody shrubs and small amounts of grasses. This suggests that the diet of the herd animals at the site was not supplemented with agricultural by-products. In conjunction with the absence of lithics (i.e., sickle blades), there is no evidence that agriculture was practiced at this site. Several lines of evidence can be used to argue that the microremains identified at both sites are relatively well preserved, i.e., their concentrations reflect close to original input in the past by humans and animals (rather than post-depositional bias due to mixing or dissolution): (1) The mere presence of dung spherulites, a highly soluble microremain (Canti 1999), indicates good preservation. (2) Given the solubility of calcitic dung spherulites and ash pseudomorphs, the presence of both microremains within the sealed archaeological contexts of Structure 1 argues for relatively good Geoarchaeological Investigation at the Site of Mashabe Sade 69 preservation of this sediment (see Gur-Arieh et al. 2014, for a solubility test of spherulites vs. pseudomorphs). (3) Given the good preservation of the calcitic microremains, dung spherulites and ash pseudomorphs, there is no reason to assume that phytoliths are not well preserved. (4) Elevated concentrations of phytoliths have been identified associated with gypsum, a mineral that forms under neutral-alkaline conditions. Their high concentration despite this environment indicates they did not suffer much post- depositional alteration. The aridity of the Negev Highlands appears to promote the preservation of all these microremains.

Excursus: dating of archaeological sediments For the ephemeral site, in the absence of datable charred remains, it was impossible to determine with certainty whether animal penning took place during the Intermediate Bronze Age or later. It was therefore important to directly date the dung deposits. An OSL project was conducted to study some of the sections from our excavation (Junge et al. forthcoming). At the central site, OSL dates confirmed the collapse/ abandonment of the structures at the end of the Intermediate Bronze Age or a while thereafter (3.7 ± 0.3 ka; 2000–1400 BCE). At the nearby site, however, the degraded dung contexts within Room 102 (Locus 018, the lower phase) were shown to date to the Iron Age or slightly later (2.7 ± 0.2 ka; 900–500 BCE). This is contra to the original dating of the site to the Intermediate Bronze Age (Cohen 1999: 117–118) and Intermediate Bronze Age pottery found in Locus 018, but in line with its layout (similar to Iron IIA sites in the region) and two Iron Age sherds identified by us on the surface. Although OSL dating schemes cannot supply as much precision as radiocarbon dating in historical periods, the ability to merely distinguish between sediments that accumulated during the Intermediate Bronze Age or the Iron Age was useful for the study of the courtyard site. While it is clear that there was activity at the location of the ephemeral site during the Intermediate Bronze Age, the degraded dung deposits are undoubtedly later. The absence of evidence for both agricultural and herding activity in the central site turns the spotlight to other sources of livelihood.

Other possible interpretations Copper production has been previously suggested as a major component of the Intermediate Bronze Age economy at the central sites (Haiman 1996; Cohen 1999; Kochavi 1967). Previous excavation at the site uncovered four copper awls and a fragment of a copper ingot (Cohen 1999: 118, 130: 4–8). A single prill identified in Structure 5 (Locus 013) is the only additional copper artefact found in the new excavation. A hammerstone was unearthed in Open Space 3, possible evidence for a small lithic industry or copper processing (cf. Haiman 1996; Vardi, Shilstein, Shalev and Yekutieli 2008). The XRF study of the ashy material in Structure 1 indicated that it does not relate to fuelling metallurgical activities. The lack of copper artefacts cannot rule out the site as a centre for copper processing. More ashy features should be studied microarchaeologically in order to reach firmer conclusions regarding metallurgical activities at central Intermediate Bronze Age sites. 70 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

Some long-distance trade may be evident at the central site based on malacofauna. The Red Sea shell identified in Structure 6 (Locus 015), and a marine shell from Structure 1 (Locus 005), adds Mashabe Sade to the widespread marine shell trade network during the Intermediate Bronze Age. Preliminary analysis of the lithic assemblage has not identified bead-making stone tools, like those encountered at Rekhes Nafha 396 or Be’er Resisim; however, the conch material appears worked (H.K. Mienes personal communication). At the current state of research, it appears that Mashabe Sade might also be involved in the trade of raw Red Sea shell materials. The only remaining subsistence strategy that could support this central Intermediate Bronze Age settlement is trade. Based on distribution of central Intermediate Bronze Age sites in Transjordan, Negev and Sinai, Haiman convincingly argues for an east- to-west trade vector from Faynan to Egypt (1996: 23–27). New radiocarbon evidence points to the Intermediate Bronze Age commencing earlier than previously thought (ca. 2500–1950 BCE–Regev et al. 2012). This places the beginning of the Negev settlement peak within the timeframe of the powerful Old Kingdom (Fifth–Sixth Dynasties) (Malek 2000; Bard 2005; Bronk Ramsey et al. 2010). Egyptian ceramics at Be’er Resisim and Har Zayyad show palpable connections between Egypt and the Negev Highlands (Goren 1996: 56). It is conceivable that Egypt expanded economically to the east. Copper could have been one of its major interests and this may also have been the reason for the prosperity in copper production at Wadi Faynan (Haiman 1996; Adams 2000). It is noteworthy that radiocarbon dates from Khirbet Hamra Ifdan— the Intermediate Bronze Age copper production site there (Levy et al. 2002: Table 1)—and the old dates from the central site of Ein Ziq (Segal 1999: 338–339) provide determinations in the first half of the Intermediate Bronze Age, contemporary to the Fifth–Sixth Dynasties in Egypt.

Conclusions A comprehensive evaluation of the macroarchaeological data and the new microarchaeological evidence from Mashabe Sade introduces a new hypothesis regarding subsistence practices at Intermediate Bronze Age Negev Highland sites— mainly, the lack of evidence of food production in a large central settlement. This follows the lithic and faunal evidence at other central sites, and corresponds more to a population involved in alternative subsistence practices, such as trade and/or participation in the copper economy of the period, possibly related to demand in Egypt. Inhabitants at these postulated trading posts may have subsisted on dry and/or preserved commodities. The archaeozoological data (Hakker-Orion 1999) indicating meat consumption at two other central sites may support this suggestion. This further suggests that these sites may have been seasonal. Three major questions remain unresolved: 1) Was Mashabe Sade a permanent settlement? 2) Supposing that the site did not function as a place of habitation—as indicated by our results, where did the people who were active there (and at other central sites) come from? 3) Assuming that central sites such as Mashabe Sade were related to trade, what was the exact function of the hundreds of round structures? Geoarchaeological Investigation at the Site of Mashabe Sade 71

Testing the ‘copper hypothesis’ and answering these questions will be accomplished through the study of additional Intermediate Bronze Age sites, central as well as ephemeral, using macro- and microarchaeological methods.

Acknowledgments This paper is a revised and updated version of the lead author’s M.A. thesis (Dunseth 2013). The work reported here was supported by the European Research Council 2008 Advanced Grant No. 229418, “Reconstructing Ancient Israel: The Exact and Life Sciences Perspective” to I. Finkelstein and S. Weiner, where R. Shahack-Gross acted as leader of the Geoarchaeology track; the Chaim Katzman Archaeology Fund (Tel Aviv University); and a GIF (the German-Israeli Foundation for Scientific Research and Development) grant (No. I-1244-107.4/2014 to R. Shahack-Gross. and M. Fuchs as principal investigators and I. Finkelstein as co-investigator. We wish to thank S. Weiner and the team of the Kimmel Center for Archaeological Science, Weizmann Institute of Science where the analytical part of the study was carried out. Additionally, we thank M. Mor for her irreplaceable assistance in logistics and organization of the project, and the numerous Tel Aviv University students and the Mashabe Sade guesthouse for their help in carrying out the excavations.

References Adams, R.B. 2000. The Early Bronze Age III–IV Transition in Southern Jordan: Evidence from Khirbet Hamra Ifdan. In: Philip, G. and Baird, D., eds. Ceramics and Change in the Early Bronze Age of the Southern Levant. Sheffield: 397–401. Albert, R.M. and Weiner, S. 2001. Study of Phytoliths in Prehistoric Ash Layers from Kebara and Tabun Caves Using a Quantitative Approach. In: Meunier, J.D. and Colins, F., eds. Phytoliths: Applications in Earth Sciences and Human History. Lisse. Anderson, P.C. 1999. Experimental Cultivation, Harvest, and Threshing of Wild Cereals. In: Anderson, P.C., ed. Prehistory of Agriculture: New Experimental and Ethnographic Approaches (UCLA Monograph 40). Los Angeles: 118–144. Avner, U. 2002. Studies in the Material and Spiritual Culture of the Negev and Sinai Populations, During the 6th–3rd Millennia B.C.E. (Ph.D. dissertation, Hebrew University of Jerusalem). Jerusalem. Avni, G., Porat, N. and Avni, Y. 2013. Byzantine-Early Islamic Agricultural Systems in the Negev Highlands: Stages of Development as Interpreted through OSL Dating. Journal of Field Archaeology 38: 332–346. Avni, Y., Porat, N. and Avni, G. 2012. Pre-farming Environment and OSL Chronology in the Negev Highlands, Israel. Journal of Arid Environments 86: 12–27. Bard, K.A. 2005. Encyclopedia of the Archaeology of Ancient Egypt. New York. Baruch, U. 1999. Identification of Wood Remains from Horbat ‘En Ziq. In: Cohen, R. Ancient Settlement of the Central Negev, Vol. I: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports 6). Jerusalem: 7*–11*. Bar-Yosef, D. 1999. Shells from Three Middle Bronze Age I Sites in the Negev Highlands. In: Cohen, R. Ancient Settlement of the Central Negev, Vol. I: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports 6). Jerusalem: 322–326 (Hebrew). Beit-Arieh, I. 2003. Archaeology of Sinai: The Ophir Expedition (Monograph Series of the Institute of Archaeology of Tel Aviv University 21). Tel Aviv. Ben-Yosef, E. 2010. Technology and Social Process: Oscillations in Iron Age Copper Production and Power in Southern Jordan (Ph.D. dissertation, University of California-San Diego). San Diego. Berna, F., Behar, A., Shahack-Gross, R., Berg, J., Boaretto, E., Gilboa, A., Sharon, I., Shalev, S., Shilstein, S., Yahalom,-Mack, N., Zorn, J.R. and Weiner, S. 2007. Sediments Exposed to High Temperatures: Reconstructing Pyrotechnological Processes in Late Bronze and Iron Age Strata at Tel Dor (Israel). Journal of Archaeological Science 34: 358–373. 72 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

Brochier, J.E., Villa, P. and Giacomarra, M. 1992. Shepherds and Sediments: Geo-ethnoarchaeology of Pastoral Sites. Journal of Anthropological Archaeology 11: 47–102. Bronk-Ramsey, C., Dee, M.W., Rowland, J.M., Higham, T.F.G., Harris, S.A., Brock, F., Quiles, A., Wild, E.M., Marcus, E.S. and Shortland, A.J. 2010. Radiocarbon-Based Chronology for Dynastic Egypt. Science 328: 1554–1557. Cabanes, D., Gadot, Y., Cabanes, M., Finkelstein, I., Weiner, S. and Shahack-Gross, R. 2012. Human Impact Around Settlement Sites: A Phytolith and Mineralogical Study for Assessing Site Boundaries, Phytolith Preservation, and Implications for Spatial Reconstructions Using Plant Remains. Journal of Archaeological Science 39: 2697–2705. Canti, M.G. 1999. The Production and Preservation of Faecal Spherulites: Animals, Environment and Taphonomy. Journal of Archaeological Science 26: 251–258. Clamer, C. and Sass, B. 1977. Middle Bronze I. In: Bar-Yosef, O. and Philips, J.L., eds. Prehistoric Investigations in Gebel Maghara Northern Sinai (Qedem 7). Jerusalem: 245–254. Clemente, I. and Gibaja, J.F. 1998. Working Processes on Cereals: An Approach Through Microwear Analysis. Journal of Archaeological Science 25: 457–464. Cohen, R. 1992. The Nomadic or Semi-Nomadic Middle Bronze Age I Settlements in the Central Negev. In: Bar-Yosef, O. and Khazanov, A., eds. Pastoralism in the Levant: Archaeological Materials in Anthropological Perspectives (Monographs in World Archaeology 10). Madison: 105–131. Cohen, R. 1999. Ancient Settlement of the Central Negev, Vol. 1: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports 6). Jerusalem (Hebrew with English Summary). Cohen, R. and Cohen-Amin, R. 2004. Ancient Settlement of the Negev Highlands, Vol. II: The Iron Age and Persian Periods (IAA Reports 20). Jerusalem (Hebrew with English summary). Cohen, R. and Dever, W.G. 1978. Preliminary Report of the Pilot Season of the ‘Central Negev Highlands Project’. BASOR 232: 29–45. Cohen, R. and Dever, W.G. 1979. Preliminary Report of the Second Season of the ‘Central Negev Highlands Project’. BASOR 236: 41–60. Cohen, R. and Dever, W.G. 1981. Preliminary Report of the Third and Final Season of the ‘Central Negev Highlands Project’. BASOR 243: 55–77. Courty, M.A., Goldberg, P. and Macphail, R. 1989. Soils and Micromorphology in Archaeology. Cambridge. Covello-Paran, K. 2008. A Bronze Age Site at Nahal Rimmonim in the Jezreel Valley. ‘At iqot 60: 53–74. Davidovitch, U., Porat, N., Gadot, Y., Avni, Y. and Lipschits, O. 2012. Archaeological Investigations and OSL Dating of Terraces at Ramat Rahel, Israel. Journal of Field Archaeology 37: 192–208. Dever, W.G. 1980. New Vistas on the EB IV (‘MB I’) Horizon in Syria-Palestine. BASOR 237: 35–64. Dever, W.G. 1985. Village Planning at Be’er Resism and Socio-Economic Structure in Early Bronze Age IV Palestine. EI 18: 18*–28*. Dever, W.G. 2014. Excavations at the Early Bronze IV Sites of Jebel Qa>aqir and Be

Age IIA Philistia: New Evidence from Tell es-Safi/Gath, Israel.Journal of Archaeological Science 39: 255–267. Evenari, M., Aharoni, Y., Shanan, L. and Tadmor, N. 1958. The Ancient Desert Agriculture in the Negev III. Early Beginnings. IEJ 8: 231–268. Evenari, M., Shanan, L. and Tadmor, N. 1971. The Negev: The Challenge of a Desert. Cambridge, MA. Finkelstein, I. 1995. Living on the Fringe: The Archaeology and History of the Negev, Sinai and Neighboring Regions in the Bronze and Iron Ages (Monographs in Mediterranean Archaeology 6). Sheffield. Frumkin, A. 2009. Stable Isotopes of a Subfossil Tamarix Tree from the Dead Sea Region, Israel, and Their Implications for the Intermediate Bronze Age Environment. Quaternary Research 71: 319–328. Ginguld, M., Perevolotsky, A. and Ungar, E.D. 1997. Living on the Margins: Livelihood Strategies of Bedouin Herd-Owners in the Northern Negev, Israel. Human Ecology 25: 567–591. Goren, Y. 1996. The Southern Levant in the Early Bronze Age IV: The Petrographic Perspective. BASOR 303: 33–72. Gur-Arieh, S., Mintz, E., Boaretto, E. and Shahack-Gross, R. 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: 4331–4347. Gur-Arieh, S., Shahack-Gross, R., Maeir, A.M., Lehmann, G., Hitchcock, L.A. and Boaretto, E. 2014. The Taphonomy and Preservation of Wood and Dung Ashes Found in Archaeological Cooking Installations: Case Studies from Iron Age Israel. Journal of Archaeological Science 46: 50–67. Haiman, M. 1992. Cairn Burials and Cairn Fields in the Negev. BASOR 287: 25–45. Haiman, M. 1996. Early Bronze Age IV Settlement Pattern of the Negev and Sinai Deserts: View from Small Marginal Temporary Sites. BASOR 303: 1–32. Haiman, M. 2012. Geopolitical Aspects of the Negev Desert in the 11th–10th Centuries BCE. In: Galil, G., Gilboa, A., Maeir, A.M. and Kahn, D., eds. The Ancient Near East in the 12th–10th Centuries BCE: Culture and History (Proceedings of the International Conference held at the University of Haifa, 2–5 May, 2010) (Alter Orient und Altes Testament 392). Münster: 199–206. Hakker-Orion, D. 1999. Faunal Remains from Middle Bronze Age I Sites in the Negev Highlands. In: Cohen, R. Ancient Settlement of the Central Negev, Vol. I: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports 6). Jerusalem: 327–335 (Hebrew). Hauptmann, A., Schmitt-Strecker, S., Levy, T.E. and Begemann, F. 2015. On Early Bronze Age Copper Bar Ingots from the Southern Levant. BASOR 373: 1–24. Junge, A., Lomax, J., Shahack-Gross, R., Dunseth, Z.C., Finkelstein, I. and Fuchs, M. Forthcoming. OSL Age Determination of Archaeological Structures Using Trapped Aeolian Sediments: A Case Study from the Negev Highlands, Israel. Geoarchaeology. Katz, O., Cabanes, D., Weiner, S., Maeir, A.M., Boaretto, E. and Shahack-Gross, R. 2010. Rapid Phytolith Extraction for Analysis of Phytolith Concentrations and Assemblages During an Excavation: An Application at Tell es-Safi/Gath, Israel. Journal of Archaeological Science 37: 1557–1563. Kenyon, K. 1951. Some Notes on the History of Jericho in the Second Millennium B.C. PEQ 83: 101–138. Kochavi, M. 1963. Har Yeruham. Bulletin of the Israel Exploration Society 27: 284–292 (Hebrew). Kochavi, M. 1967. The Settlement of the Negev in the Middle Bronze (Canaanite) I Age (Ph.D. dissertation, Hebrew University of Jerusalem). Jerusalem (Hebrew with English summary). Kramer, C. 1982. Village Ethnoarchaeology: Rural Iran in Archaeological Perspective. New York. Langgut, D., Finkelstein, I., Litt, T., Neumann, F.H. and Stein, M. 2015. Vegetation and Climate Changes During the Bronze and Iron Ages (~3600–600 BCE) in the Southern Levant Based on Palynological Records. Radiocarbon 57: 217–235. Levy, T.E., Adams, R.B., Hauptmann, A., Prange, M., Schmitt-Strecker, S. and Najjar, M. 2002. Early Bronze Age Metallurgy: A Newly Discovered Copper Manufactory in Southern Jordan. Antiquity 75: 425–437. MacDonald, B. 1992. The Southern Ghors and Northeast ‘Arabah Archaeological Survey (SAM 5). Sheffield. 74 Zachary C. Dunseth, Andrea Junge, Markus Fuchs, Israel Finkelstein and Ruth Shahack-Gross

Macphail, R.I., Courty, M.A., Hartner, J. and Wattez, J. 1997. The Soil Micromorphological Evidence of Domestic Occupation and Stabling Activities. In: Maggi, R., ed. Arene Candide: A Functional and Environmental Assessment of the Holocene Sequences Excavated by Bernardo Brea (1940–1950) Vol. V. Umana: 55–88. Malek, J. 2000. The Old Kingdom (c. 2686–2160 BC). In: Shaw, I., ed. The Oxford History of Ancient Egypt. Oxford: 83–107. Marx, E. 1967. Bedouin of the Negev. Manchester. Matthews, W., French, C.A.I., Lawrence, T., Cutler, D.F. and Jones, M.K. 1997. Microstratigraphic Traces of Site Formation Processes and Human Activities. World Archaeology 29: 281–308. Migowski, C., Agnon, A., Bookman, R., Negendank, J.F.W. and Stein, M. 2004. Recurrence Pattern of Holocene Earthquakes along the Dead Sea Transform Revealed by Varve-counting and Radiocarbon Dating of Lacustrine Sediments. Earth and Planetary Science Letters 222: 301–314. Miller, N.F. 1984. The Use of Dung as Fuel: An Ethnographic Example and an Archaeological Application. Paléorient 10: 71–79. Mumford, G. 2012. Ras Budran and the Old Kingdom Trade in Red Sea Shells and other Exotica. British Museum Studies in Ancient Egypt and Sudan 18: 107–145. Muniz, A.A. 2007. Feeding the Periphery: Modeling Early Bronze Age Economies and the Cultural Landscape of the Faynan District, Southern Jordan (Ph.D. dissertation, University of California at San Diego). San Diego. Musil, A. 1908. Arabia Petraea III. Vienna. Nissenbaum, A., Serban, A. and Connan, J. 1999. Asphalt from Horbat En Ziq. In: Cohen, R., Ancient Settlement of the Central Negev, Vol. 1: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports 6). Jerusalem: 12–16. Oren, E.D. and Yekutieli, Y. 1990. North Sinai during the MB I Period—Pastoral Nomadism and Sedentary Settlement. Eretz-Israel 21: 6–22 (Hebrew with English summary). Palmer, E.H. 1871. Desert of the Exodus. Cambridge, UK. Palumbo, G. 2001. The Early Bronze Age IV. In MacDonald, B., Adams, R.B. and Bienkowski, P., eds. The Archaeology of Jordan. Sheffield: 233–269. Prag, K. 1985. Ancient and Modern Pastoral Migration in the Levant. Levant 17: 81–88. Regev, J., Miroschedji, de P., Greenberg, R., Braun, E., Greenhut, Z. and Boaretto, E. 2012. Chronology of the Early Bronze Age in the Southern Levant: New Analysis for a High Chronology. Radiocarbon 54: 525–566. Regev, L., Poduska, K.M., Addadi, L., Weiner, S. and Boretto, E. 2010. Distinguishing between Calcites Formed by Different Mechanisms Using Infrared Spectrometry: Archaeological Applications. Journal of Archaeological Science 37: 3022–3029. Rosen, S.A. 1987. Demographic Trends in the Negev Highlands: Preliminary Results from the Emergency Survey. BASOR 266: 45–58. Rosen, S.A. 2011. An Investigation into Early Desert Pastoralism: Excavations at the Camel Site, Negev (Cotsen Institute of Archaeology Monograph 69). Los Angeles. Rosen, S.A. 2012. The Chipped-Stone Assemblage. In: Eisenberg, E. The Bronze Age IV Site at Sha>ar Ha-Golan. >Atiqot 69: 53–59. Rosen, S.A. and Vardi, J. 2014. Appendix IIH: The Chipped Stone Assemblage From Be’er Resisim: A Final Report. In: Dever, W.G., ed. Excavations at the Early Bronze IV Sites of Jebel Qa’aqir and Be’er Resisim. Winona Lake: 327–338. Rothenberg, B. 1990. The Discovery of a Copper Mine and Smelter from the End of the Early Bronze Age (EB IV) in the Timna Valley. Institute of Archaeo-Metallurgical Studies 15–16: 1–8. Rothenberg, B. 1999. Archaeo-Metallurgical Researches in the Southern Arabah 1959–1990, Part I: Late Pottery Neolithic to Early Bronze IV. PEQ 131: 68–89. Saidel, B.A. 2002. The Excavations at Rekhes Nafha 396 in the Negev Highlands, Israel. BASOR 325: 37–63. Saidel, B.A., Erickson-Gini, T., Vardi, J., Rosen, S.A., Maher, E. and Greenfield, H. 2006. Test Excavations at Rogem Be’erotayim in Western Negev. Journal of the Israel Prehistoric Society 36: 201–229. Segal, I. 1999. Metallurgical Remains from Horbat En Ziq and Horbat Be’er Resisim, Central Negev, Israel. In: Cohen, R., ed. Ancient Settlement of the Central Negev, Vol. 1: The Geoarchaeological Investigation at the Site of Mashabe Sade 75

Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports, No. 6). Jerusalem: 38*–43*. Shahack-Gross, R. 2011. Herbivorous Livestock Dung: Formation, Taphonomy, Methods for Identification and Archaeological Significance. Journal of Archaeological Science 38: 205–218. Shahack-Gross, R., Boaretto, E., Cabanes, D., Katz, O. and Finkelstein, I. 2014. Subsistence Economy in the Negev Highlands: The Iron Age and the Byzantine/Early Islamic Period. Levant 46: 98–117. Shahack-Gross, R. and Finkelstein, I. 2008. Subsistence Practices in an Arid Environment: A Geoarchaeological Investigation in an Iron Age Site, the Negev Highlands, Israel. Journal of Archaeological Science 35: 965–982. Shahack–Gross, R. and Finkelstein, I. 2015. Settlement Oscillations in the Negev Highlands Revisited: The Impact of Microarchaeological Methods. Radiocarbon 57: 253–264. Shahack-Gross, R., Marshall, F. and Weiner, S. 2003. Geo-ethnoarchaeology of Pastoral Sites: The Identification of Livestock Enclosures in Abandoned Maasai Settlements. Journal of Archaeological Science 30: 439–459. Shalev, S., Shilstein, S. and Yekutieli, Y. 2006. XRF Study of Archaeological and Metallurgical Material from an Ancient Copper-Smelting Site near Ein-Yahav, Israel. Talanta 70: 909–913. Sowada, K.N. 2009. Egypt in the Eastern Mediterranean during the Old Kingdom: An Archaeological Perspective. Fribourg. Tsartsidou, G., Lev-Yadun, S., Efstratiou, N. and Weiner, S. 2008. Ethnoarchaeological Study of Phytolith Assemblages from an Agro-Pastoral Village in Northern Greece (Sarakini): Development and Application of a Phytolith Difference Index. Journal of Archaeological Science 35: 600–613. Tufnell, O. 1958. Lachish IV: The Bronze Age. London. Valamoti, S.M. and Charles, M. 2005. Distinguishing Food from Fodder through the Study of Charred Plant Remains: An Experimental Approach to Dung-derived Chaff. Vegetation History and Archaeobotany 14: 528–533. Vardi, J., Shilstein, S., Shalev, S. and Yekutieli, Y. 2008. The Early Bronze Age IV Chipped and Ground Stone Assemblages of En Yahav and Its Relation to Copper Smelting Activities. Journal of the Israel Prehistoric Society 38: 1–20. Warnock, P.J. 2014. Appendix IID: Analysis of Wood Charcoal. In: Dever, W.G., ed. Excavations at the Early Bronze IV Sites of Jebel Qa’aqir and Be’er Resisim. Winona Lake: 301–310. Weiner, S. 2010. Microarchaeology: Beyond the Visible Archaeological Record. Cambridge, UK. Yahalom-Mack, N., Gadot, Y., Eliyahu-Behar, A., Bechar, S., Shilstein, S. and Finkelstein, I. 2014. Metalworking at Hazor: A Long-Term Perspective. Oxford Journal of Archaeology 33: 19–45. Yekutieli, Y., Shilstein, S. and Shalev, S. 2005. ‘En Yahav: A Copper Smelting Site in the ‘Arava. BASOR 340: 1–21.

7.4 Preliminary results from Nizzana 332

The excavations at Nahal Nizzana 332 were designed to parallel the investigations into subsistence and chronology at Nahal Boqer 66 (Dunseth et al. 2018). In general, this was to test whether the subsistence strategies based on livestock rearing reconstructed at Nahal Boqer

66 are unique, or a pattern common to small sites during the IBA, and also to see when other small sites are situated in the IBA sequence. The site was specifically chosen as a case study as it represents a different geographical and microenvironment setting (above) than Nahal

Boqer 66.

In addition, given the interesting spatial conclusions of Dunseth et al. (2018), we wanted to explore the spatial distribution of archaeological assemblages at small sites at both the macro- and microscales. Spatial studies are common in prehistoric investigations, but rarer in studies of Bronze Age sites. Notable exceptions in the Negev are the works of Saidel (2002a) and

Rosen (2011a).

Negev researchers have raised questions whether investigating a few test trenches at a site is appropriate to broadly reconstruct subsistence strategies (Bruins and van der Plicht 2017), although note they too are guilty of this, Shahack-Gross and Finkelstein 2017: 1229), and indeed, multiple samples from the same context often have variable ranges of microremain assemblages (Gur-Arieh et al. 2013; Dunseth et al. 2018: Appendix 2). At Nahal Nizzana 332 we wanted to show the broad intrasite distribution of microremains to reconstruct activity areas and context as well as test whether probes are effective at evaluating subsistence strategies.

153 7.4.1 Excavation Summary

Excavations took place 12–17 April 2017 at Nahal Nizzana 332 within three small complexes

visible on the surface.1 These were designated Nizzana 332.1, 332.2 and 332.3 (Fig. 8).

Because of scant finds at Nahal Nizzana 332.2 and 332.3, the bulk of this chapter describes

Nahal Nizzana 332.1. Field observations of Nizzana 332.2 and 332.3 will be discussed below.

Nizzana 332.1

Nizzana 332.2

0 40 m

Figure 8: Schematic site plan redrawn after Haiman 1991: 128, with Nizzana 332.1 and 332.2 outlined in red. Note that Nizzana 332.3 is beyond the limit of the site plan approximately 250 m to the west.

1 Excavations were directed by the author, and assisted by Erin Hall (332.1), Omer Ze’evi (332.2) and Mark Cavanaugh (332.3), all of Tel Aviv University.

154 Due to the shallow sediment accumulation at the site, the small-scale excavation focused on

the complete collection of all macroscopic artifacts (ceramic, lithic, faunal, etc.) together with

a corresponding microarchaeological analysis of sediments to reconstruct spatial patterns. A

100 x 100 m arbitrary local grid was placed over the site. Loci (described as year

(YY)/subsite(NIZ-X)/locus(#), e.g., 2017/Nizzana-332.1/L.001 = 17/NIZ/1) were designated for the three units excavated and were dug as 1 x 1 m subunits in order to efficiently collect

horizontal spatial information (modified following Saidel 2002a; Rosen 2011a). Each unit is

described in the text as its final locus. All excavated sediments were dry-sieved through 5-mm

mesh; grey sediments from the courtyard and the structure were further sifted through a

graduated series of 2-mm, 1-mm and 0.5-mm sieves for botanical and faunal remains. Locus

boundaries, archaeological features, artifacts and sample locations were recorded in three

dimensions using a Leica TS06 total station. Georectified 3D models of the excavated units

were created using Agisoft PhotoScan 1.42 following Prins (2016).

Sediments immediately above bedrock were sampled in 10 g plastic vials systematically across

the grid to reconstruct horizontal spatial variation (n = 72). Two sections of north-south

transects were also investigated in detail to study vertical (stratigraphic) development of

microremain assemblages over space: 1) along the eastern section of the N Squares in the north-

south transect; 2) along the western section of the H Squares, inside the structure (see Fig. 9).

OSL samples were also collected from the latter section (see below). Control samples (n = 6)

were collected from the immediate landscape, about 100–300 m away from the site.

All excavated squares were backfilled after excavation, according to National Parks Authority

requests.

155 9

10 L. 17/NIZ/10 11

12

13

14

15

16

L. 17/NIZ/11 L. 17/NIZ/9 17

18

19

NMLKJIHG

Figure 9: Georectified orthophoto of Nahal Nizzana 332.1 with 1 x 1 m grid letters and numbers indicated. Red lines indicate vertical sections studied. Note scale in Squares M-N12 is 0.8 m.

156 7.4.2 Field Observations

7.4.2.1 Nahal Nizzana Complex 332.1

At Nahal Nizzana 332.1, the general plan of the structures was already visible on the surface.

The site is comprised of large central Courtyard 17/NIZ/10 (approx. 10 m in diameter) attached to two smaller rooms (Room 17/NIZ/11 and Room 17/NIZ/12, both approx. 5 x 4 m). Two

transects of the site were opened in 1 x 1 m squares: one north-south through the center of

Courtyard 17/NIZ/10, and one east-west transecting Courtyard 17/NIZ/9, attached Room

17/NIZ/11, and outside the complex (designated Open Area 17/NIZ/9). The northwestern

Room 17/NIZ/12 was not explored. A total of 36 m2 were excavated.

Walls in Room 17/NIZ/11, the southern of the two rooms, were preserved to a maximum of only 1–2 courses (c. 60–80 cm). The walls predominantly collapsed to the north, along with the downward slope, and to the east, into the courtyard. The courtyard itself is built of only a

single course of stones, generally laid on their longest side. The courtyard wall abuts Room

17/NIZ/11 and incorporates the room into its perimeter, suggesting some internal chronological

development to the site.

Sediment deposition was shallow, ranging c. 15–30 cm, sloping down north with bedrock. In

the courtyard and open area, vertical deposition was characterized in the field as two distinct

sedimentary units: yellow-grey sediments (Open Area: L. 17/NIZ/1; courtyard: L. 17/NIZ/2)

overlying compact light-grey-brown deposit over bedrock (open area: L. 17/NIZ/4+9;

courtyard: L. 17/NIZ/10). In Squares K and L, the eastward collapse of medium-large

fieldstones, interpreted as wall collapse from Room 17/NIZ/11, sealed the lower layer of light-

grey-brown sediment. Over the extent of the excavated area, no visible differences in sediment

color or texture were noted horizontally in the field.

157

In the attached room, sediment was characterized by a sequence of three archaeological layers

(from top to bottom): 1) loose yellow-grey sediment (L. 17/NIZ/3), 2) a layer of collapsed thin

limestone slabs (L. 17/NIZ/7, interpreted in the field as roofing slabs), 3) and a light-grey-

brown floor deposit overlying bedrock (L. 17/NIZ/11). In Square G16 a c. 10 x 10 x 5 cm light

grey deposit was noted below a collapsed wall stone.

7.4.2.2 Nahal Nizzana Complex 332.2

In Nizzana 332.2, a 1 x 6 m transect was made through a large central courtyard (L. 17/NIZ-

2/2) and one of the surrounding rooms (L. 17/NIZ-2/1). An additional 1 x 4 m transect was cut through an adjacent structure immediately north (L. 17/NIZ-2/3). Accumulation of sediment was relatively thick in these structures (c. 20–60 cm), although the numbers of artifacts were extremely limited. Notably, no pottery and only a few lithics (n = 12, all debitage or debris) were found in these probes. A single uncharred ovicaprine dung pellet was found near the surface (likely modern) and was not analyzed.

7.4.2.3 Nahal Nizzana Complex 332.3

At Nizzana 332.3, two 1 x 2 m small probes were made in adjacent circular structures.

Excavation was abandoned after removal of topsoil (L. 17/NIZ-3/1) revealed the characteristic stone-lined rectangular cut of Negev burials (e.g., Nahal Boqer 66: Cohen 1999; Dunseth et al.

2018, above). A number of lithics were found: a few tools (n = 3) but mainly debris and debitage (n = 41), including a small number of burnt flakes and debris (n = 3). A few sherds were collected, none diagnostic.

7.4.3 Macroarchaeological data

The remainder of this section will only deal with the findings from Nahal Nizzana 332.1.

158

7.4.3.1 Pottery

A total of 191 sherds were found in excavation, 40 of which are indicative to a certain period.

Three late (post-Roman) sherds from different vessels were found on the surface of the central

courtyard (L. 17/NZ/2): one unidentified handle, and two small sherds of Gaza Grey Ware. An

additional late intrusive sherd was found in the lower grey-brown sediment in the open area

(L. 17/NIZ/9). Two small pieces of iron shrapnel, likely from modern military activity, were

also found on the surface in Squares M14 (L. 17/NIZ/2) and H15 (L. 17/NIZ/3), and a single

piece of intrusive modern glass was found in Square N17 (L. 17/NIZ/9). Outside of the

excavation area, surface survey revealed some ribbed Byzantine–Early Islamic pottery in the

immediate vicinity.

Eleven IBA sherds were identified based on diagnostic form and/or distinctive decoration (e.g.,

band-combing and/or notch decoration). These were found in the accumulation above bedrock

in all three investigated units: the courtyard, structure and open area. An additional 25 sherds—

a few rims but mainly coarseware body sherds with one-sided sooting—of characteristic

holemouth cooking vessels, could be more generally attributed to the third millennium BCE

(Sebbane et al. 1993; see Avner 2006 for a discussion of antecedents). Based on a macroscopic analysis of fabric, manufacture and finish, the remaining non-diagnostic sherds (n = 151) are tentatively attributed to the third millennium BCE and will be considered as a coherent

assemblage as such for analysis.

Sherds from closed vessels dominated the entire assemblage (45%), while sherds from cooking and open vessels make up 13% and 10% respectively (Fig. 10A; note these have not yet been

drawn). However, of diagnostic sherds (n = 44), cooking vessels dominate the small

assemblage (n = 25, 57%), while open bowls/lamps (n = 8, 18%) and closed storage vessels

(small juglets and storage jars, n = 11, 25%) make up the remainder (Fig. 10B).

159 Nahal Nizzana 332.1 A Whole Assemblage Open Area Courtyard Structure Open vs. Closed (all sherds) L.17/NIZ/9 L.17/NIZ/10 L.17/NIZ/11 n = 191 n = 88 n = 40 n = 20 Indeterminate Open Open n = 1 n = 1 Indeterminate 5% 5% n = 19 n = 7 18% Open 10% Open n = 4 Indeterminate Cooking n = 13 10% n = 25 15% Cooking n = 61 n = 8 Cooking 32% 13% Indeterminate 9% n = 7 n = 37 17% Cooking 42% n = 8 Closed 40% Closed Closed n = 10 n = 30 n = 22 50% Closed 34% 55% n = 86 45%

Nahal Nizzana 332.1 Open Area Courtyard Structure B All Diagnostic Sherds L.17/NIZ/9 L.17/NIZ/10 L.17/NIZ/11 (Rims, handles, bases + indicative body sherds) n = 14 n = 14 n = 11 n = 44 Lamp/bowl n = 1 9% Lamp/Bowl Storage Jars n = 8 Lamp/bowl n = 2 Lamp/bowl n = 3 18% 18% n = 4 22% 29% Cooking Cooking Storage Jars Cooking n = 8 n = 7 Cooking 57% Storage Jars 50% n = 8 n = 8 n = 25 n = 2 73% 18% 57% 14%

Storage Jars n = 2 14% Juglets n = 2 Juglets 14% n = 3 7% Figure 10: Nahal Nizzana 332.1 pottery analysis. A) presents total assemblage (all sherds); B) presents diagnostic sherds only. Note that surface sherds are not included in the locus totals. 160 The general paucity of finds and the dominance of cooking vessels over storage jars and other

items follows a general pattern at EB and IBA small sites throughout the Negev (cf. Saidel

2002b; Saidel and Haiman 2014: 99–136; Saidel et al. 2006; Dunseth et al. 2018), as well as pastoral sites from the Byzantine–Early Islamic periods (cf. Oded Sites: Rosen and Avni 1997;

Rekhes Nafha 396: Saidel 2004). This is in contrast to the assemblages from large IBA sites dominated by storage vessels such as Ein Ziq, Be’er Resisim and Mashabe Sade (Cohen 1986;

1999: 230–237; Dunseth et al. 2018: Figure 9).

7.4.3.2 Spatial distribution: pottery

To control for differences in size of sherds, the distribution of sherds was normalized as weight/m2 (Fig. 11). It is clear that sherd concentrations are very low (<20 g/m2) in both

Courtyard 17/NIZ/10 and Room 17/NIZ/11. Most sherds are concentrated immediately outside the complex (approximately 100 to >300g/m2). Because the site is built on a slope, the sherd

accumulation outside the complex could theoretically be related to natural sedimentary flow

processes; however, the lack of a similar pattern both within the courtyard and the western

structure suggests that this is related to original dumping outside the structure, and then re-

deposition. This pattern is similar to refuse scatter identified through ethnoarchaeological

studies at sites across the world (e.g., Hayden and Cannon 1983; Cameron and Tomka 1993)

and most applicably at recent modern Bedouin villages in Jordan where the living quarters and

animal enclosures are kept clean and refuse is scattered around the site (e.g., Simms 1988;

Banning and Köhler-Rollefson 1992; Palmer et al. 2007).

Non-diagnostic sherds from a single open vessel were found in all three units just above

bedrock (open area: Squares K–N16; L17; courtyard: Square K15; structure: Square J16).

Because of the physical barrier of the courtyard and structure walls, this is interpreted as

movement after initial breakage, and evidence for sweeping/cleaning activities in antiquity.

161 Figure 11: Pottery distribution normalized to weight in g of ceramic material by excavated area (m2). Weights presented include both diagnostic and nondiagnostic sherds.

162 Although diagnostic sherds were only found in low quantities, some spatial patterning of types

of vessels by context can be suggested. Sherds from cooking vessels dominate all three

contexts, but overwhelmingly so inside Room 17/NIZ/11 (n = 8, 73%). Sherds from small

juglets (n = 2) were only uncovered in the central courtyard (Fig. 10B).

Two indicative cooking vessel samples were preliminarily analyzed petrographically by the

author and M.A.S. Martin (March 2018). Both samples were characterized by inclusions of

weathered granite. Similar wares are known from the Wadi Arabah (as suggested by M.A.S.

Martin, personal comm.), or south Sinai (Porat 1989).

7.4.3.3 Lithic and groundstone assemblage2

In comparison to the pottery assemblage, lithic material was abundant at the site (total assemblage n = 947, Tables 3, 4). Although the sample size is much smaller than those of the larger excavations at the small Camel Site (n = 27,757, Hermon et al. 2011) and Rekhes Nafha

396 (n = 10,617, Saidel 2002a), or the large IBA sites of Ein Ziq (n = 11,895, Vardi 2005;

Vardi et al. 2007) and Be’er Resisim (n = 12,098, Rosen and Vardi 2014), the proportions of debris, debitage and tools is similar to these assemblages, and suggests this is a good sample for analysis. The assemblage can be broadly attributed to the third millennium BCE, except for a single Epipaleolithic lunate found inside Room 17/NIZ/11 (Square J16). No tabular scrapers, associated with the EB II and earleir (Goring-Morris 1993; Rosen 1997; Rosen and Gopher

2003) were identified in the assemblage.

2 Analysis and identification of the assemblage was carried out by Bar Efrati and Dana Ackerfeld (TAU). Statistical analysis and comparison to other assemblages was conducted by the author.

163 Table 3: Whole lithic assemblage from Nizzana 332.1.

Debris n = % of group % of total assemblage Chunks 240 77.2% 25.3% Chips 71 22.8% 7.5% Total debris 311 100.0% 32.8% Debitage Flakes 274 57.8% 28.9% P.E. Flakes 80 16.9% 8.4% Micro-flakes 51 10.8% 5.4% Blades & Bladelets 37 7.8% 3.9% C.T.E. 32 6.8% 3.4% Total debitage 474 100.0% 50.0% Tools General 80 76.2% 8.4% Awls/Borers 18 17.1% 1.9% Scrapers 7 6.7% 0.7% Tool spall 20 19.0% 2.1% Total tools 105 100.0% 11.1% Cores 37 3.9% Hammers 1 0.1% Total Assemblage 948 100.0%

The raw material is generally local, made from fine-grained brown-black flint nodules common in the area, although a few non-local stones were found. The assemblage can be described as ad hoc and showing the full chaîne opératoire of production and use (B. Efrati and D.

Ackerfeld, personal comm.), including debris (n = 311, 32.8% of the whole assemblage), debitage (n = 474, 50% of the assemblage) and tools (n = 105, 11%). Debitage is dominated by flakes (n = 274, 58%), followed by primary element flakes (n = 80, 17%) and microflakes

(n = 51, 11%), allowing a general characterization of this assemblage as a flake industry (Rosen

1997).

Tools make up a large percentage of approximately 11% of the entire assemblage. In comparison, tools only make up ~2% at both the small Camel Site (Hermon et al. 2011: Table

164 6.1) and Rehkes Nafha 396 (Saidel 2002a: 47). While large amounts of tools in comparison to

debris and debitage have been associated with opportunistic collection methods, the systematic sieving and collection of microliths suggests this is a real pattern and not a methodological

artifact (see discussion in Rosen 1997: 37; Vardi 2014: 94–95). In general, the tools are overwhelmingly dominated by ad hoc variants (n = 80, 76%). Specialized tools included awls/borers (n = 18, 17%), common at many small Negev sites from the EB and IBA, as well as scrapers (n = 7, 6%). The dominance of awls/borers and drills at many large and small IBA sites in the Negev has been connected to small-scale ad hoc production of beads from ostrich egg or Red Sea marine shells (Vardi et al. 2007; Hermon et al. 2011; Rosen and Vardi 2014).

A few marine shells were found (n =3), however, they were neither drilled nor the common materials (e.g., Red Sea Lambis or Glycimeris shells) used in bead production found at other

IBA sites (cf. Bar-Yosef 1999). Although finished rectangular shell beads (n = 3) somewhat common at IBA sites in the Negev were found in the central courtyard (Squares L15 and M14, cf. Cohen 1999: Fig. 58:23, 113: 6,8, 139A: 4), this cannot be clearly associated with craft specialization at the site.

Blades were moderately common in the assemblage (n = 32, 7.8% of the debitage), but notably no sickle blades with characteristic gloss were observed (‘sickle gloss’ being formed from the repeated action of cutting through plant material, Anderson 1999; Rosen et al. 2015). This follows the general pattern at most EB and IBA Negev sites, both large and small, where sickle blades are either absent or make up a miniscule portion of the assemblage (cf. Rosen 1997;

Saidel 2002a; Saidel et al. 2006; Yekutieli et al. 2005; Vardi 2005, 2014; Vardi et al. 2007;

Hermon et al. 2011; Rosen and Vardi 2014).

165 Table 4: Lithic assemblage from Nahal Nizzana 332.1 by locus.

Debris Debitage Tools Context Locus Description Chunks Chips Flakes P.E. Flakes Micro-Flakes Blades CTE Cores Tools Awl/Borers Scrapers Tool Spall Total Courtyard L. 17/NIZ/2 Surface + Topsoil 54 15 48 11 16 5 7 10 19 0 2 6 193 Courtyard L. 17/NIZ/10 Accumulation on bedrock 27 8 32 9 5 7 4 5 8 2 0 3 110 Courtyard Total 81 23 80 20 21 12 11 15 27 2 2 9 303 Open Area L.17/NIZ/1 Surface + Topsoil 43 22 36 13 16 4 3 2 11 0 2 1 153 Open Area L. 17/NIZ/9 Accumulation on bedrock 63 16 90 32 4 12 8 10 23 4 0 7 269 Open Area Total 106 38 126 45 20 16 11 12 34 4 2 8 422 Structure 11 L.17/NIZ/3 Surface + Topsoil 19 4 10 2 2 1 3 3 5 3 1 1 54 Structure 11 L. 17/NIZ/11 Accumulation on bedrock 34 6 58 13 8 8 7 7 14 9 2 2 168 Structure 11 Total 53 10 68 15 10 9 10 10 19 12 3 3 222 Grand Total 240 71 274 80 51 37 32 37 80 18 7 20 947

166 Grinding stones (n = 5) were found in three locations: four complete large grinding stones were

found along the northern perimeter of the courtyard (Squares L10 and M9; see yellow dots in

Fig. 12), and a smaller grinding stone associated with a flint hammerstone was found within the western room (Square J15). All grinding stones are made from quartz sandstone. Similar types were found at the Camel Site and Rekhes Nafha 396, suggested to be groundstone production sites (Rosen 2011a and Saidel 2002a respectively). Unlike the other two sites, no

chips or evidence for the production of grinding stones was identified in this excavation.

7.4.3.4 Spatial distribution of lithics

Figure 12 shows concentration of lithic quantities normalized by area for the accumulation on bedrock (L. 17/NIZ/9, L. 17/NIZ/10 and L. 17/NIZ/11). The spatial distribution of lithic

material from the accumulation on bedrock generally reflects the distribution of the pottery

assemblage, with the strongest concentrations of debris, debitage and tools outside the southern

limit of the central courtyard. However, unlike the pottery distribution, there is also a marked

concentration of debitage and tools in Room 17/NIZ/11. This is not reflected in the debris and

suggests a special activity area—possibly both the use and production of tools.

By density, borers/awls especially are concentrated inside Room 17/NIZ/11, suggesting that

some sort of drilling (or piercing) activity took place there. As described above, the

concentration of debris and debitage finds outside the courtyard and living areas parallels many

ethnoarchaeological studies in the region (cf. regionally: Simms 1988; Banning and Köhler-

Rollefson 1992; Palmer et al. 2007: Fig. 12.22; more globally: Binford 1972; Hayden and

Cannon 1983).

167 A B C Debitage (qt/m2) Debris (qt/m2) 1.0 - 2.0 2.0 - 4.0 0 - 2.00 4.0 - 6.0 2.00 - 4.00 Tools (qt/m2) 6.0 - 8.0 8.0 - 10.0 4.00 - 6.00 0.00 - 1.00 10.0 - 12.0 12.0 - 14.0 6.00 - 8.00 1.00 - 2.00 14.0 - 16.0 16.0 - 18.0 8.00 - 10.00 2.00 - 3.00 18.0 - 20.0 10.00 - 12.00 3.00 - 4.00 20.0 - 22.0 22.0 - 24.0 12.00 - 14.00 4.00 - 5.00 24.0 - 26.0 26.0 - 28.0 14.00 - 16.00 5.00 - 6.00 28.0 - 30.0 16.00 - 18.00 6.00 - 7.00 30.0 - 32.0 32.0 - 34.0 18.00 - 20.00 7.00 - 7.52 34.0 - 36.0 36.0 - 38.0 20.00 - 22.00 38.0 - 40.0 40.0 - 42.0 22.00 - 24.00 42.0 - 44.0 24.00 - 26.00 44.0 - 46.0 46.0 - 46.7 26.00 - 28.00 28.00 - 29.69

Figure 12: Distribution of lithics at Nahal Nizzana 332.1 according to quantity/area. A) Tool distribution (includes all tool types); B) Debitage distribution; C) Debris distribution.

168 7.4.3.5 Botanical and faunal remains

Macrobotanical remains were exceptionally rare in the excavation. Two pieces of charcoal (one

Tamarix sp. twig, one unidentified, M. Cavanaugh personal comm.) and one unidentified unburnt seed were found and sent for radiocarbon dating. No faunal remains were found in any context.

7.4.4 Chronometric data: radiocarbon and optically stimulated luminescence

Three samples described above were sent for radiocarbon dating (Table 5, Fig. 13). The two charcoal dates were similar and show a range of ~2340–2200 BCE (1 σ). This range is comparable to the radiocarbon determinations from Nahal Boqer 66, Ein Ziq and Har Dimon

(Segal and Carmi 1996: 94; Dunseth et al. 2017: Tables 1, 3), all of which are limited to only the first half of the IBA.

The unburnt seed from the courtyard was modern (103.03 ± 0.38 pMC; 1956-1957 CE). This is unsurprising given the modern material on the surface and shallow deposition in the central courtyard.

Additionally, the north-south section inside Room 17/NIZ/11 was sampled for OSL dating by

A. Junge (Justus-Liebig University Giessen) (Fig. 13). The deposit is dated to 1.6 ± 0.2 kya

(200–600 CE). This date is incompatible with the radiocarbon dates from the same structure and the open area, and likely suggests a disturbance in this particular part of the site during the

Roman–Byzantine period. Note from the orthorectified photos that one wall stone is collapsed southward in this exact location.

169 Table 5: Radiocarbon determinations from Nahal Nizzana 332.1. Note Beta-499151 is a modern intrusion.

Lab Number Locus Field ID Type Context Species Wt. (mg) δ13C ‰ BP 1 σ (calBCE) 2 σ (calBCE) 2456 ( 5.2%) 2418 2407 ( 6.2%) 2374 2337 ( 7.1%) 2322 Beta-499150 17/NIZ/9 K16/R1 charcoal On bedrock in open area n.d. 470.5 -24.4 3830 ± 30 2368 ( 1.1%) 2356 2308 (61.1%) 2206 2351 (81.4%) 2198 2164 ( 1.5%) 2152 2398 ( 1.2%) 2384 Beta-499149 17/NIZ/11 H15/R1 charcoal On bedrock in structure Tamarix sp. 11 -21.6 3810 ± 30 2292 (68.2%) 2202 2346 (84.3%) 2190 2182 (10.0%) 2141 Beta-499151 17/NIZ/10 M12/R2 seed (unburnt) Lower dung deposit n.d. 17.2 -26 103.03 ± 0.38 pMC1956 CE 1956-1957 CE

Figure 13: Location and age ranges of radiocarbon determinations and OSL samples. Later intrusions in red.

170 7.4.5 Microarchaeological data

A total of 72 archaeological sediment samples and six controls from the immediate vicinity

were collected and analyzed by FTIR for mineralogical information, and by optical microscopy

for the analysis of phytolith, dung spherulite and ash pseudomorph quantification.

7.4.5.1 Mineral characterization

Preliminary analysis of all control and archaeological samples presented a consistent

mineralogy of calcite, unheated clay, quartz and minor peaks of dolomite in all samples.

Gypsum was occasionally present in some samples, but no discernable pattern was observed

in its distribution. The clay component in all samples showed no evidence for exposure to

heating above 400°C (cf. Berna et al. 2007; Forget et al. 2015). Calcite grinding curves

indicated higher disorder in most archaeological samples over local sediment and limestone

controls, especially those from Room 17/NIZ/11 and Courtyard 17/NIZ/10 (Fig. 14). The

samples with the highest calcite disorder were consistently from the courtyard, likely reflecting the high concentrations of dung spherulites in these samples (detailed below) (Dunseth and

Shahack-Gross 2018).

171 540 Carob ash

500 NIZ-C1 alluvial sediment 460 Local Limestone (NIZ-C8) Plaster 420 Spar 380

340

v2 normalized height Controls 300 Courtyard 17/NIZ/10 sediments Open area 17/NIZ/9 sediments 260 Structure 17/NIZ/11 sediments Grey sediments from L. 17/NIZ/11

220 20 40 60 80 100 120 140 160 180 200 220 240 v4 normalized height

Figure 14: Calcite grinding curves for archaeological and control sediments from Nahal Nizzana 332.1. Local sediment controls are plotted (NIZ-C1, alluvial sediment from north of site) and limestone (NIZ-C8). Note all archaeological sediments plot between alluvial controls, ash and plaster. Courtyard sediments plot closer to the plaster trendline, due to elevated concentrations of dung spherulites in the sediments (see below, and cf. Dunseth and Shahack-Gross 2018). Reference trendlines are from Regev et al. 2010 and Dunseth and Shahack-Gross 2018.

172 7.4.5.2 Dung spherulites

Dung spherulites were essentially absent in the control sediments, averaging 0.01±0.02 million/g of sediment. Surface sediments (n = 16, sampled 5 cm below the surface) in the three units showed slight enrichment over controls, averaging 0.15±0.11 million/g sediment in Open

Area 17/NIZ/9, 0.3±0.3 million/g of sediment in Room 17/NIZ/11, and significantly higher concentrations in Courtyard 17/NIZ/10, averaging 2±2 million/g of sediment (Fig. 15A).

In the lower archaeological deposits, Open Area 17/NIZ/9 averaged 0.3±0.2 million/g of sediment, approximately the same or slightly elevated over the surface sediments.

Concentrations of this sort suggest essentially a very minor contribution of dung to the sedimentary context (Fig. 15B).

In comparison, sediments in Room 17/NIZ/11 show a gradient of dung spherulite concentrations, decreasing from the east to the west (Fig. 15B and 16A). Along the eastern half in Squares J15–16 the samples are enriched, averaging 3.5±2.9 million/g of sediment; in the western G and H squares, the average is only 0.4±0.3 million/g of sediment. This can be interpreted in a few ways. One, small or young animals were sheltered inside the room, as done even recently in Bedouin encampments (Palmer et al. 2007: 379). However, researchers have reported that very young (< 2 months old) ovicaprines tend not to produce dung spherulites

(Brochier et al. 1992: 55). It is possible that this reflects dung brought into the structure from the outside by foot traffic, although entryways were not identified during excavation. It is possible that dung was kept inside the structure for use as fuel. If dung was used as a fuel source, it might help explain the low amounts of charcoal and the almost complete lack of ash pseudomorphs at the site (below).

173 2.50 60 A Surface Sediments

50 Dung Spherulite Concentration 2.00 10

40 6 / 1 g of sediment

1.50

30

1.00 / 1 g of sediment

6 20 10 Phytolith Concentration

0.50 10

0 0

Controls Outside Inside Courtyard (n = 6) Structure Structure (n = 6) (n = 4) (n = 6)

2.50 60 B Accumulation on bedrock

50 Dung Spherulite Concentration 2.00 10

40 6 / 1 g of sediment

1.50

30

1.00 / 1 g of sediment

6 20 10 Phytolith Concentration

0.50 10

0 0

Controls Outside Inside Courtyard (n = 6) Structure Structure (n = 21) (n = 20) (n = 20)

Figure 15: A) Boxplots of phytolith and dung spherulite concentrations from surface sediments and controls. Minor enrichment of phytolith concentrations can be observed in all archaeological contexts. Enrichment in dung spherulites is apparent only in the surface sediments of the courtyard. B) Boxplots of phytolith and dung spherulite concentrations from the archaeological accumulation on bedrock. Phytolith concentrations are essentially the same inside and outside the structures (within error) in comparison to the surface sediments, while concentrations in the courtyard are enriched. Dung spherulite concentrations are markedly enriched in the court- yard and in a few samples from inside the structure. Note different scales for phytolith and dung spherulite concentrations.

174 A Dung spherulite B concentrations Phytolith Log value concentrations 0.000105804 - 0.056652863 Log value 0.056652863 - 0.105119819 0.0242 - 0.07121 0.105119819 - 0.14666125 0.07121 - 0.11822 0.14666125 - 0.195128205 0.11822 - 0.16523 0.195128205 - 0.251675264 0.16523 - 0.21224 0.251675264 - 0.317649491 0.21224 - 0.25925 0.317649491 - 0.394622523 0.25925 - 0.30626 0.394622523 - 0.484428012 0.30626 - 0.35327 0.484428012 - 0.589205303 0.35327 - 0.40028 0.589205303 - 0.711450401 0.40028 - 0.44729 0.711450401 - 0.854075427 0.44729 - 0.4943 0.854075427 - 1.020477995 1.020477996 - 1.214622145 1.214622146 - 1.441132779 1.44113278 - 1.705405831

Figure 16: Inverse distance weighted interpolation of logarithmically normalized microremain concentrations. Point locations are marked as white circles. A) Dung spherulite concentration interpolation. Clear enrichment in dung spherulites is interpolated inside Courtyard 17/NIZ/10, and to a lesser extent in the eastern part of Structure 17/NIZ/11. B) Phytolith concentration interpolation. Enriched phytolith concentrations are only clearly interpolated in the northern part of Courtyard 17/NIZ/10. Note in general, higher values indicate higher microremain concentrations.

175 Dung spherulite concentrations were the highest in the central courtyard, averaging

approximately 20±15 million/g of sediment in the grey-brown accumulation on bedrock, an

order of magnitude higher than the surface sediments (Figs. 15, 16A). These concentrations, in

tandem with the elevated phytolith concentrations (below), show these sediments are degraded

dung deposits.

7.4.5.3 Ash pseudomorphs

Ash pseudomorphs were identified rarely (n = 14/72) and only in negligible concentrations,

0.04–0.18 million/g of sediment. In general, there is a weak correlation with dung spherulite concentration (r2 = 0.26). Spatially, ash pseudomorphs were only identified sporadically in the courtyard and western structure. Preservation bias does not appear to be an issue, given the good preservation of dung spherulites (above), shown experimentally to dissolve more readily than ash pseudomorphs (Gur-Arieh et al. 2013).

7.4.5.4 Phytoliths

Phytoliths were essentially absent in controls, averaging only 0.05±0.06 million/g of sediment.

This is only approximately 1–2 phytoliths counted per 16 fields of view per sample (Fig. 16A).

In general, phytolith concentrations were relatively low in surface sediments at the site. In the surface sediments, the three units all showed very minor enrichment over controls: in Open

Area 17/NIZ/9 0.20±0.20 million/g of sediment; in Room 17/NIZ/11 0.19±0.05 million/g of

sediment; in Courtyard 17/NIZ/10, an average of 0.22±0.14 million/g of sediment (Fig. 15A).

The grey deposits of Open Area 17/NIZ/9 and Room 17/NIZ/11 accumulated on bedrock have even lower phytolith concentrations than the surface sediments, averaging only 0.17±0.19 and

0.18±0.10 million/g of sediment. In Room 17/NIZ/11,the values range from 0.22±0.10 to

0.14±0.09 million/g of sediment east to west. The co-occurrence of high concentrations of dung

176 spherulites and phytoliths in this room (at least in the eastern part of this room) indicates a

similar origin, i.e., dung (Figs. 15B, 16).

Courtyard 17/NIZ/10 is characterized by the highest phytolith concentrations, averaging

0.49±0.55 million/g of sediment, approximately double the values analyzed in surface sediments. These values are in the range of other degraded dung sediments at the Iron Age sites

of Nahal Boqer, Atar Haroa (Shahack-Gross and Finkelstein 2008: Table 2; Shahack-Gross et al. 2014: Table 1), Mashabe Sade peripheral site (Dunseth et al. 2016: Table 4), and lower than similar deposits at third millennium BCE Nahal Boqer 66 (Dunseth et al. 2018: Fig. 5).

7.4.5.5 Phytolith morphologies from dung contexts

Phytolith morphologies were analyzed from two samples, one from the central courtyard having the highest phytolith concentrations (sample L11.2) and one from the open area outside

(M19.2) (Fig. 17). Phytoliths in anatomical connection (multicellular phytoliths), generally a good indicator of preservation, were rare (only 2% in both samples). Obvious weathering (e.g., pitting, etc.) of phytoliths was not readily apparent (only in 1–2% of the phytolith assemblage), and no melted phytoliths were identified in the assemblage. In general, the phytoliths can be considered moderately well-preserved, and reliable enough for cautious reconstruction.

Both samples were similar, dominated by morphologies indicative of dicot wood/bark (e.g., ellipsoid, discoid and irregular morphologies) at 48% and 59% respectively, and indeterminate

phytolith morphologies (e.g., mainly fibers, common but not exclusive to dicots, 33 and 28%

respectively). Phytoliths from dicot leaves made up relatively small proportion of 8 and 5%

respectively. Phytoliths indicative of grasses were only represented by morphotypes common

to leaf/stem and inflorescence (e.g., short cells, and parallelepiped thin psilate). Panicoid +

chloridoid to festucoid short cell ratios (cf. Regev et al. 2015: Supplement 3) of 0.5–1 suggest

some dominance of C3 plants in both samples. No phytolith morphologies attributable to grass

177 inflorescence specifically (e.g., echinate, dendritic, verrucate long cells or papillae morphologies) were identified in these samples, or in any other sample analyzed. Sedge (hat- shaped) and palm phytoliths (spherical echinate) were also absent.

Phytolith Morphologies

100% 2%

90% 28% 80% 33%

70%

60%

50%

48% 59% 40%

30%

20% 8% 10% 5% 6% 7% 4% 0% 1% NIZ-L11.2 NIZ-M19.2 Dicot Wood/Bark Grass Leaves/Stems or Inflor Dicot Leaves Grass Leaves/Stems Sedges Grass Inflor Indeterminate Weathered

Figure 17: Phytolith morphologies of samples from the courtyard (NIZ-L11.2) and the open area outside the structure (NIZ-M19.2) at Nahal Nizzana 332.1. Note the dominance of dicotyledonous morphotypes in both samples.

Based on the microremain concentrations and phytolith morphology data from the degraded dung deposits in the central courtyard, animal diet can be reconstructed as free-grazing primarily on woody dicotyledonous vegetation. There is no evidence of foddering with cereals or cereal byproducts. Based on the very low concentrations of phytoliths in general, very low grass phytoliths in particular, very high proportions of phytoliths from dicot wood/bark and low percentages of dicot leaves, late summer or autumn grazing might be tentatively suggested

178 (Shahack-Gross and Finkelstein 2008, but note reservations on identifying seasonality based solely on phytoliths in Dunseth et al. 2019).

7.4.6 Summary of Nahal Nizzana

The macro- and microscopic evidence from Nahal Nizzana follows the same general pastoral pattern found at Nahal Boqer:

1. Radiocarbon dates for the IBA, concentrated only in the first half of the IBA (c. 2400–

2200 BCE).

2. Microarchaeological data indicating accumulations of degraded dung material, mainly

concentrated in the large central courtyard, is clear evidence of herding. Degraded dung

shows a phytolith assemblage characterized by low concentrations and morphologies

consistent with local grazing of woody dicots. Absence of inflorescence phytoliths—

and especially dendritic phytoliths common in domestic cereals in particular—act as

unequivocal evidence for the lack of foddering by agricultural products, and by

extension, lack of opportunistic farming practiced by the inhabitants of this site.

3. Limited faunal and botanical remains, common to many small sites, are likely due to

limited consumption of meat, and/or scavenging of discarded bones by desert fauna

before and after abandonment. Paucity of charred botanical remains might be explained

by the limited exposure of refuse deposits, deflation of archaeological sediments, or use

of dung as fuel.

4. The ceramic assemblage follows a clear pattern common to small sites, limited in

quantity and forms, dominated by cooking and open vessels over closed storage vessels.

This is in direct contrast to the large sites, where the assemblages are more varied and

contain large numbers of closed storage vessels (ranging in size from small

amphoriskoi, medium storage jars, and large pithoi).

179 5. In addition, the lithic assemblage is similar to ad hoc assemblages from a variety of

both small and large sites in the Negev Highlands including the EB sites from the Camel

Site and the Western Negev, as well as IBA sites at Rekhes Nafha 396, Rogem

Be’erotayim, Ein Ziq and Be’er Resisim. In general this is characterized by an ad hoc

flake industry, with significant portions of awls/borers and drilling tools. This lithic

continuity, labeled the ‘Timnian’ by Rosen (2011b) and others, appears to reflect the

indigenous desert populations involved in all aspects of IBA settlement.

180 8 Discussion

8.1 Methodological contributions

With studies of dung to explore ancient subsistence practices, rapidly identifying animal dung is essential for efficient screening of samples before lengthy analysis. For example, because degraded dung remains at sites can appear macroscopically similar to ash, dung accumulations have erroneously been interpreted as ash or even destruction layers (e.g., Shahack-Gross and

Finkelstein 2008 contra Cohen 1970).

The grinding-curve method proposed in Dunseth and Shahack-Gross (2018) based on data from

Nahal Boqer 66 and modern reference material, was applied here on the Nahal Nizzana 332 deposits. The patterns repeated: degraded dung sediments from the central courtyard were easily distinguished from the other sediment samples. As such, it proves to be an effective technique for rapidly evaluating archaeological sediment samples from open air sites in calcareous environments such as the Negev Highlands.

However, because of the overlap of dung and lime plaster grinding curves, it also acts as a cautionary case study against blindly comparing sediments to established standards. Calcite grinding curves must be tailored to each site and should be verified by other means such as optical microscopy or XRD (already, Xu et al. 2015).

A methodological issue raised by other researchers in the Negev (e.g., Bruins and van der Plicht

2017, above), considered whether small-scale microarchaeological investigations can be representative of subsistence practices either of an entire site or regional scale. At a site scale, the detailed excavation of approximately 30% of the site and sampling strategy at Nahal

Nizzana 332 suggests that activity areas and especially the penning of animals can be clearly differentiated even in small probes if room/areas were used for specific purposes. Similar

181 results regarding subsistence strategies would be determined even in smaller probes (i.e., a few square meters) in the central courtyard, such as those reported from Nahal Boqer 66 and other sites (cf. Shahack-Gross and Finkelstein 2008; Shahack-Gross et al. 2014; Dunseth et al. 2016,

2018).

On a larger scale, the consistent and repeated patterns in the macro- and microarchaeological assemblages at different types of sites indicate that small-scale exposures of approximately 25–

75 m2 at multiple sites are sufficient for producing enough comparable data to evaluate subsistence practices at a regional scale. It should be cautioned that a single site cannot be reliably used to determine regional subsistence strategies.

A final methodological contribution of this research is that this is the first—and to date only— microarchaeological investigation into subsistence practices of the IBA in the Negev

Highlands. Most importantly, this research has been able to evaluate older theories regarding subsistence and chronology directly.

8.2 Subsistence strategies at IBA Negev sites: updated model

Previous to this dissertation, Haiman’s (1996) prevailing model for the IBA in the Negev

Highlands argued for two populations: one at small sites subsisting on herding and some opportunistic cereal cultivation, and another dwelling permanently in large sites. At large sites, it was suggested that inhabitants were engaged in copper processing and trade; how they procured subsistence was less clear, and it was suggested they were supported by the smaller sites. Major confirmations and deviations from this model emerge from the research presented here.

182 8.2.1 Small sites: Microarchaeological data as direct indicators of pastoral nomadism

Based on the data presented above, it is clear that the small sites of Nahal Boqer 66 and Nahal

Nizzana 332 in this study were primarily engaged in livestock management. The well-defined

microremain assemblages enriched in concentrations of both dung spherulites and phytoliths

at these sites are clear indicators of abundant dung remains. Based on detailed morphological study of phytoliths in the dung assemblages shows an animal diet based mainly on woody and herbaceous dicotyledonous plants, which is interpreted as free-grazing on the local environment. Lower concentrations of phytoliths, low amounts of grass phytoliths and no dendritic phytoliths most common in cereal inflorescence are overwhelming evidence that livestock were not fed agricultural byproducts, and by extension, that opportunistic cereal agriculture was not practiced by the inhabitants of these small sites. These data follow similar ethnoarchaeological results from pastoral premodern Bedouin sites (Nahal Yatir, Shahack-

Gross and Finkelstein 2008; Umm Sarbut, Shahack-Gross et al. 2014), and at archaeological sites dating to the Iron IIA (Atar Haroa: Shahack-Gross and Finkelstein 2008, Shahack-Gross et al. 2014; Nahal Boqer [Iron Age]: Shahack-Gross et al. 2014; Mashabe Sade peripheral site

[Iron Age]: Dunseth 2013; Dunseth et al. 2016). This phenomenon is in clear contrast to later

Byzantine/Early Islamic sites, large and small (e.g., Wadi el-Mustayer: Shahack-Gross et al.

2014; Shivta: Dunseth et al. 2019), where the substantial microarchaeological data for mixed agropastoralism and foddering with cereal byproducts is corroborated by textual records (e.g.

Kraemer 1958; Mayerson 1965), and detailed OSL investigations of agricultural terraces across the Negev (Avni et al. 2013, 2019).

8.2.2 Large sites: Microarchaeological data as direct indicators of fire activities

In contrast, the microarchaeological assemblages of the large sites of Mashabe Sade and Ein

Ziq showed neither evidence for accumulations of animal dung, nor evidence for the cultivation

183 of domestic cereals in built structures or in open areas. Only refuse deposits and fire features

(hearths, secondary ash deposits) show high concentrations of microremains, dominated by phytoliths, and to a lesser extent, ash pseudomorphs.

Phytolith assemblages from fire features at Ein Ziq were characterized by morphologies similar to dung deposits at the small sites, providing evidence for the exploitation of the local woody dicots. This is confirmed by the limited macrobotanical analysis of charcoals from Ein Ziq, including Anabasis sp., Tamarix sp., Retama raetam, Acacia sp. (Dunseth et al. 2017: Table 1;

V. Caracuta, personal comm.; for earlier studies see also Baruch 1999; Segal and Carmi 2004:

145; and compare to Be’er Resisim: Warnock 1991, 2014). At Ein Ziq, grass phytoliths made up to 30% of the assemblages, including significant portions of inflorescence phytoliths, likely from whole grasses used as kindling. Dendritic phytoliths—most common in domestic cereal inflorescence—specifically were rare. Intriguingly, the use of dung as fuel was not apparent in any context at Ein Ziq. This is relatively uncommon in the history of the Bronze and Iron Ages

(see review in Shahack-Gross 2019: Table 5.1), possibly providing an indirect indicator for the lack of livestock at large sites.

8.2.3 Pottery assemblages of nomadic pastoralists

In addition to the microarchaeological remains, the scant pottery assemblages at Nahal Boqer and Nahal Nizzana follow a distinct pattern common to small sites in the Negev Highlands.

Cribb (1991) suggested that nomadic pastoralists have a narrower range of sizes and types of ceramic vessels due to their mobility. This was shown to be the case at small EB, IBA and

Byzantine–Early Islamic Negev sites by Saidel (2002a, 2002b, 2004), who noted that, in general, ancient small sites are characterized by a paucity of ceramic remains, and a limited number of typological groups. Specifically, the third millennium BCE sites are dominated by

184 holemouth vessels (cooking mainly) and miscellaneous open vessels over storage containers

(cf. Saidel 2002b, Saidel et al. 2006; Saidel and Haiman 2014).

The ceramic assemblages at Nahal Boqer and especially the larger assemblage at Nahal

Nizzana exhibited the same pattern. In general, this appears to be common to many small sites

throughout the history of the Negev Highlands and Sinai (see also scant remains from Gebel

Maghara, Clamer and Sass 1977; and Gebel Gunna, Bar-Yosef et al. 1986), as well as later

pastoral nomadic encampments from the Byzantine/Early Islamic period (Rosen and Avni

1997; Saidel 2004).

8.2.4 Pottery assemblages of large sites as indicators of trade

In contrast to the small sites, the pottery assemblages of Ein Ziq and Mashabe Sade are dominated by closed storage vessels (64% and 81%, respectively) over cooking (18% and 12%) and open forms (15% and 2%) (Dunseth et al. 2018, and unpublished data). At the much larger copper production site of Khirbet Hamra Ifdan (n = 19,486 diagnostic sherds) the assemblage is somewhat similar to Ein Ziq: 40% storage vessels, and 15% cooking; although open serving vessels (40%) make up a far larger percent at Khirbet Hamra Ifdan than at the Negev sites

(Gidding 2016: Table 6.8). These data are comparable to larger 3rd millennium BCE urban sites such as Arad (at least Area K) and ‘Aradian’ type small settlements such as Sheikh Muhsen in south Sinai interpreted as trading outposts (Saidel 2002b: Table 4) and even Byzantine–Early

Islamic farmsteads (Saidel 2004). This dominance is a determinative indicator of the importance of exchange networks to the settlement of large central sites.

8.2.5 Site layout, activity areas and use of space at small and large sites

Based on their architectural layout, nearly all small sites characteristic to various periods have been interpreted as short-term encampments attached to animal pens, that is, as representing

185 (agro)pastoral nomadism. This has been suggested by many Negev researchers based on (stated

or unstated) 19th and 20th century ethnographic analogies to architectural layout and use of space (e.g., Woolley and Lawrence 1914: 22; Glueck 1955; Dever 1985; Haiman 1992, 1996;

Cohen 1999; Finkelstein 1995; Gittlen 2006; see also discussion in Dunseth 2013 and Dunseth et al. 2016).

Utilizing geoarchaeological methods, this dissertation empirically confirms that—at least at

the third millennium BCE sites of Nahal Boqer and Nahal Nizzana—small sites were primarily

engaged in animal husbandry, with large open courtyards presenting unambiguous evidence

for stabling of animals. At Nahal Nizzana, the high-resolution sampling also shows this is

particularly limited to the courtyard, and that animal penning areas were maintained clear of debris and refuse (possibly to protect livestock from injury). At Nahal Boqer, the distribution and concentration of dung is more extensive, which is possibly a reflection of its significantly longer use (based on radiocarbon data—over a millennium in comparison to only brief activity during the IBA at Nahal Nizzana); alternatively, it may also be explained by Nahal Boqer’s relative size (c. 2000 m2 vs c. 150 m2), which could have held larger herds (and thus increased deposition of dung). It is notable that the dung-rich areas excavated across Nahal Boqer did not contain many macroscopic remains, possibly reflecting similar maintenance and cleaning activities in the past as indicated at Nahal Nizzana, or more modern Bedouin sites (see especially, Palmer et al. 2007).

Encircling the courtyard are smaller attached rooms showing evidence for fire activity (in the case of Nahal Boqer) as well as cleaning and maintenance activities (both Nahal Boqer 66 and

Nahal Nizzana 332). Based on the high-resolution sampling and excavation of Nahal Nizzana, domestic activity appears to be focused inside structures, with lithic production and domestic activities (cooking) inside structures, which is in contrast to what was observed at the larger

186 sites. This appears to be similar at Rekhes Nafha 396 (Saidel 2002a: 45, Fig. 3), and the Camel

Site (Rosen 2011a), where small hearths were found inside rooms surrounding the central

courtyards. At the larger site of Ein Ziq (Dunseth et al. 2018, above) the majority of activity

remains (and numerous fire features) were found both inside and outside structures. Based on

the hidden caches, and concentrations of storage vessels at Ein Ziq, it is likely that many of the

small roofed structures common to large central sites acted as storage facilities (see already

Finkelstein 1989: 134).

Although the material culture (e.g., ceramics, lithics and copper items) at small and large sites

is similar and recent radiocarbon data (Dunseth et al. 2017, above) shows that during the IBA

the sites were inhabited contemporaneously, it is still unclear whether (or how) the large and

small settlements interacted with each other. Assuming that small sites were inhabited by

pastoral desert groups and that the large sites were visited by (local or non-local) traders en

route, the presence of copper items and Red Sea shells at small sites may reflect exchange

relations between these groups. Furthermore, zooarchaeological evidence, though scarce,

indicates consumption of meat-rich portions at large sites (Hakker-Orion 1999). It is therefore

possible that pastoral nomadic groups living at small sites supplied inhabitants of large sites

with meat, milk and other herd products (already Haiman 1996: 24–25).

8.2.6 Trade: Copper and desert goods

Based on site locations, layout and especially the presence of copper ingots, numerous grinding stones and hammerstones, Haiman (1996) suggested that the large sites were permanent sites engaged primarily in trade and copper production/processing. Previous studies on the lithic assemblages at Ein Ziq and Be’er Resisim (Vardi 2005; Vardi et al. 2007) questioned the plausibility of copper production based on the lack of associated tools and installations (e.g., anvils, crucibles, ingot molds, etc.) found in abundance at copper production sites in Faynan

187 (cf. Adams 2000; Levy et al. 2002; Gidding 2016). Regardless, copper processing was still assumed to be a supporting industry at large sites, based on the presence of slags and the aforementioned hammer- and groundstone assemblages (e.g., Vardi et al. 2008).

Although numerous pieces of copper scrap and finished products were noted in both our and previous excavations at the large sites (cf. Cohen 1999: 137–188; Dunseth et al. 2018), an exhaustive XRF investigation of archaeological sediments found no evidence for any copper processing activities at Ein Ziq (Dunseth et al. 2018: Appendix 4). This paralleled earlier data from Mashabe Sade—admittedly a smaller dataset—which also showed no evidence for copper processing activities in refuse deposits (Dunseth et al. 2016). It is unlikely that copper processing occurred at large central sites, at least at the scale proposed by Haiman, and certainly not as a coherent industry or subsistence base for the site’s population.

Regardless, evidence for long-distance desert trade at large central sites is overwhelming, and includes the ingot hoards and finished copper objects from Faynan (Segal et al. 1996–1997;

Hauptmann et al. 2015), bitumen from the Dead Sea (Nissenbaum et al. 1999), marine shells from the Red Sea (Bar-Yosef 1999; Dunseth et al. 2016), and the dominance of storage vessels over other types of pottery (Dunseth et al. 2018). Provenance analysis of ceramic vessels also showed that most material from large Negev Highlands sites was imported from locations in the surrounding regions including the Northern Negev/Shephelah, the Judean and Samarian mountains and Transjordan (Goren 1996).

Given this evidence for long-distance trade, the complete absence of macro- or microarchaeological signals for pastoral activity or food production, as well as their location in defensible localities or near water sources (Haiman 1996), large central sites can best be understood as functioning as trading hubs and/or waystations along the copper exchange network.

188 Trade activity is also attested at small sites, although not very clearly in the excavations

presented here. Previous petrographic analysis of sherds from small sites showed that only a

small portion of the pottery found at small sites can be sourced from the central Negev. The

majority of analyzed sherds (n < 100 total from published excavations) were produced in the

surrounding regions, even as far away as the Judean and Samarian mountains, Transjordan,

and possibly Egypt and Sinai (e.g., Saidel 2002a: 54–57, Figs. 15–16 [petrography by Y.

Goren]; Rosen 2011a: 72–75, Table 5.2 [petrography by Y. Goren]; Cohen-Weinberger and

Saidel 2014). Preliminary analysis of holemouth cooking pots from both Nahal Boqer and

Nahal Nizzana identified weathered granitic inclusions, which can be traced to either south

Sinai or Wadi Arabah (M.A.S. Martin, personal comm., also personal observations, cf. Porat

1984: 59–61, 2003: 266; Goren 1996: 47; Cohen-Weinberger and Saidel 2014: 163). However, a more detailed study is needed in order to reconstruct the trade networks of small IBA sites in the Negev Highlands.

Numerous ethnographic and historic studies show that mobile desert communities are never isolated from the sedentary communities in surrounding regions (e.g., Barth 1961; Marx 1967,

1992; Khazanov 1984; for a thorough recent review, see Hammer and Arbuckle 2019; for a few local historical examples of interaction, for the medieval period, Hütteroth and Abdulfattah

1977; for Ottoman and Mandate periods, Frantzman and Kark 2011). While the lithic assemblages of all sites are likely produced locally (see Rosen 1997 for a broad discussion of the differences in lithic material between the Negev and further north), it is likely that small

sites traded animals and animal byproducts with large central sites for copper (already Haiman

1996), and communities in more fertile regions for grain and other goods.

189 8.2.7 A note on copper exchange vectors

Haiman (1996) traced the movement of copper from east to west, mined in Faynan, heading

northwest through the central sites in the Negev Highlands, and then further northwest along a

string of smaller sites in north Sinai towards Egypt. Originally attributed to the First

Intermediate Period, after the new radiocarbon chronology of Regev et al. (2012), this has since been convincingly linked to prosperity in the Old Kingdom, continuing from the EB (e.g.,

Dunseth 2013; Gidding 2016; Ben-Yosef et al. 2016; Dunseth et al. 2016, 2017, 2018;

Finkelstein et al. 2018). Lead isotope studies confirmed that the Negev and Hebron Hill ingots were from copper ores in Faynan (Segal and Halicz 2005; Hauptmann et al. 2015).

Since the above publications, two important papers have analyzed lead isotopes in Egyptian copper artifacts from museum collections (of the Old Kingdom specimens, 1st – 6th Dynasties,

Kmošek et al. 2018; Rademakers et al. 2018). These papers show varied ores for the copper

objects, including possible sources in the Eastern Desert, Sinai, Arabia and Anatolia. Faynan,

previously assumed to be the main source of copper, was only clearly the source of a few

objects from the 5th Dynasty tombs at el Mahâsna (Rademakers et al. 2018: 184). Although

some of the conclusions in these studies—specifically, provenancing items to specific mines

and underexplored areas, issues of mixing, etc.—may be have some methodological issues

(Ben-Yosef 2018), these papers importantly suggest that the copper is not only from Faynan

and Sinai, but from across the ancient Near East.

This revelation should also force some reconsideration not only of the demands of the Egyptian

copper market, but also raise the possibility that copper mined in Faynan was traded elsewhere.

If sites with where IBA ingots outside the Negev are plotted on a map, they indicate a northern direction: Hebron Hills (although the findspot is unclear, Dever and Tadmor 1976), Tel Hebron

(Hauptmann et al. 2015: Table 2), Lachish (Tufnell 1958) and Hazor (Yahalom-Mack et al.

190 2014). Hence, it is not unreasonable to assume that copper is moving through the Negev towards the coast (cf. IBA settlement in the coastal plain, Gophna and Portugali 1988) and north along either maritime or land routes towards the EB IV urban sites in Syria (e.g.,

Akkermans and Schwartz 2003; Cooper 2006). Although the link between the Faynan copper industry and the Egyptian market cannot be denied, especially as the trajectory of the southern

EB–IBA copper system parallels the rise and fall of the Egyptian Old Kingdom, a possible

Syrian copper connection should be studied further.

8.2.8 Lithic assemblages as indirect indicators of subsistence

Indirect indicators of farming—namely sickle blades—are largely absent from all EB and IBA assemblages from the Negev Highlands. Although only one lithic assemblage referred-to in this dissertation was analyzed in detail, it is notable that not a single sickle blade was identified in any of the excavations presented above. This follows the general pattern of small and large

Negev sites during the EB and IBA, as observed in more detailed lithic analyses. Researchers suggest that the paucity of sickle blade from all assemblages of published Negev Highland sites reflects the absence of (or extremely limited) farming practiced during the EB and IBA (cf.

Saidel 2002a; Saidel et al. 2006; Vardi et al. 2007; Vardi 2014; Hermon et al. 2011; Rosen and

Vardi 2014). This is in direct contrast to the Uvda Valley (Avner 1990; Rosen 1997), which has a microenvironment better-suited to seasonal agriculture, and especially to northern EB tell sites such as Bet Yerah, Yarmouth, and Hartuv where specialized Canaanean sickle blades made up sometimes over 32% of the total tool assemblages (Rosen 1997: Fig. 3.17).

It should be emphasized that the suggestions about subsistence derived by lithic analyses at

Ein Ziq (Vardi 2005, Vardi et al. 2007) and Nahal Nizzana (this study) are corroborated and directly confirmed by the microarchaeological assemblages (phytolith, dung spherulite, mineralogical and elemental). Thus, in the absence of other assemblages, the

191 microarchaeological data alone can be used to determine the presence/absence of cereal agriculture (see also the models presented in Harvey and Fuller 2005).

8.3 Chronology and settlement history of the Negev Highlands

The data presented in this dissertation suggests that all IBA sites, large central trading hubs and small pastoral encampments, coexisted only during the first half of the IBA, c. 2500–2200.

This parallels the Egyptian Old Kingdom (Ben-Yosef et al. 2016; Dunseth et al. 2017;

Finkelstein et al. 2018; above). It is reasonable to propose that prosperity in the south ended with the decline of Old Kingdom Egypt c. 2200 BCE and the resulting decay of demand for

Arabah copper (Ben-Yosef et al. 2016: 80; Dunseth et al. 2016; Finkelstein et al. 2018).

Small sites appear to be a long-term phenomenon, beginning in the EB I at the latest and continuing (and likely increasing) through the first half of the IBA uninterrupted. This argues that the southern settlement system, often reacting to florescence of settlement further north

(cf. Finkelstein 1988, 1995b; Rosen 2011b), is actually somewhat insulated from the urban system to the north during the third millennium BCE. Turning to the broader picture, during the preceding EB II–III, the copper system was arguably controlled by the gateway community at Arad (Ilan and Sebbane 1989; Finkelstein 1991, 1995b; Gidding 2016; Finkelstein et al.

2018), which administered the distribution of copper towards destinations in Egypt and northern Canaan. The collapse of urban sites to the north at the end of the EB III c. 2500 BCE brought about drastic change.

Arad, which appears to have continued into the EB III (see Gidding 2016 and especially

Finkelstein et al. 2018 for a detailed discussion), was abandoned. The copper exchange networks were taken over (or rerouted) by indigenous desert groups during the IBA. To differ from the Beersheva–Arad Valley in the EB, the desert locations further south was preferred as pathways to the Mediterranean coast (cf. Haiman 1996) and eventually to Egypt (and possibly

192 north, see above). At important stops, either defensible locations or near perennial water sources, trading outposts developed, likely growing organically over time. The production and transportation of copper intensified because of increased demand during the peak period of the

Old Kingdom during the 5th and 6th Dynasties (and possibly, also the growth of urban Syrian sites further north).

With the collapse of the Old Kingdom around 2200 BCE, the core market for the IBA copper system fell apart. The abandonment of both large sites and small sites simultaneously suggest they too were in some ways mutually dependent. After a short lacuna, small-scale opportunistic copper production at a much smaller scale resumed at a handful of copper production sites – e.g., Ein Yahav, Hamra Ifdan (Gidding 2016). The resumption of copper trade likely also explains the single radiocarbon date at Ein Ziq dated to the second half of the IBA.

Interestingly, while copper production resumed there is no evidence for either the establishment of new or re-habitation of older small pastoral sites.

193 9 Conclusion

To summarize:

• This dissertation presents a new method to rapidly identify ancient dung at

archaeological sites using FTIR spectroscopy. This has the potential to impact routine

in lab- or on-site field analysis, and has ramifications on interpretation of archaeological

sediments. In addition, the method suggests that the mineralogical signature of ancient

animal dung is resistant to diagenetic change in arid environments, and may become an

efficient proxy in the future to assess the state of preservation of dung deposits used for

isotopic studies, paleoclimate reconstructions and radiocarbon dating.

• This dissertation presents new macro- and microarchaeological investigations at the

large IBA central sites of Ein Ziq and Mashabe Sade, as well as the two small sites of

Nahal Boqer 66 and Nahal Nizzana 332 to explore issues of subsistence, chronology

and settlement history. These are first high resolution microarchaeological studies

conducted at IBA sites anywhere in the Southern Levant.

• The microarchaeological data indicates that small IBA Negev sites were supported by

livestock management. There is no evidence for cereal agriculture at these sites. Scant

pottery assemblages suggest a limited ceramic repertoire dominated by cooking and

serving vessels. Distribution of archaeological remains at Nahal Nizzana 332 in

particular follow ethnographic and ethnoarchaeological patterns observed at ephemeral

sites across the world.

• In contrast, the microarchaeological data from large sites shows no evidence for food

production – neither livestock management nor cereal agriculture. Extensive XRF

analysis of sediments showed that copper, ore or slag were not processed at these sites.

The lack of evidence for food production and copper processing, the dominance of non-

194 local storage vessels and the presence of desert trade goods including copper (finished

items and scrap), bitumen and Red Sea shells suggests that trade was the economic base

for large central sites.

• Regarding chronology and settlement history, based on the radiocarbon investigations

presented here and a review of previous evidence, large central and small site types

were contemporary. Small sites have a long settlement history in the Negev, existing at

least from the EB I through the IBA (c. 3300–2300 BCE). Large central sites are unique

to the first half of the IBA (c. 2500–2200 BCE). Both site types were abandoned after

c. 2200 BCE. This parallels the trajectory of the Egyptian Old Kingdom, and matches

the important data coming out of Jordan and the copper producing areas (Gidding 2016;

Ben-Yosef et al. 2016).

195 10 Bibliography

Note that the bibliography includes only sources referred to in the main text of the dissertation, not the published articles.

Adams, R.B. 2000. The Early Bronze Age III–IV transition in Southern Jordan: Evidence from Khirbet Hamra Ifdan. In: Philip, G. and Baird, D. eds. Ceramics and Change in the Early Bronze Age of the Southern Levant. Sheffield: 397–401.

Adams, R. and Genz, H. 1995. Excavations at Wadi Fidan 4: A Chalcolithic village complex in the copper ore district of Feinan, Southern Jordan. Palestine Exploration Quarterly 127: 8–20.

Aharoni, Y., Evenari, M., Shanan, L., and Tadmor, N.H. 1960. The ancient desert agriculture in the Negev V: An Israelite agricultural settlement at Ramat Matred. Israel Exploration Journal 10: 23–36, 97–111.

Akkermans, P.M.M.G. and Schwartz, G.M. 2003. The Archaeology of Syria: From Complex Hunter-gatherers to Early Urban Societies (c. 16,000–300 BC). Cambridge.

Albert, R.M., and Weiner, S. 2001. Study of phytoliths in prehistoric ash layers from Kebara and Tabun Caves using a quantitative approach. In: Meunier, J.D. and Colins, F., eds., Phytoliths: Applications in Earth Sciences and Human History. Lisse: 251–266.

Albert, R.M., Shahack-Gross, R., Cabanes, D., Gilboa, A., Lev-Yadun, S., Portillo, M., Sharon, I., Boaretto, E., and Weiner, S. 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: 57–75.

Anderson, P.C. 1999. Experimental cultivation, harvest and threshing of wild cereals. In: Anderson, P.C., ed. Prehistory of Agriculture: New Experimental and Ethnographic Approaches (UCLA Monograph 40). Los Angeles: 118–144.

Arbuckle, B.S. and Hammer, E.L. 2019. The rise of pastoralism in the Ancient Near East. Journal of Archaeological Research. https://doi.org/10.1007/s10814-018-9124-8 (online ahead of print).

Arnot, H.J. and Pautard, F.G.E. 1970. Calcification in plants. In: Schraer, H., ed., Biological Calcification: Cellular and Molecular Aspects. New York: 375–446.

196 Avner, U. 1990. Ancient agricultural settlement and religion in ‘Uvda Valley, Southern Israel. Biblical Archaeologist 53: 125–141.

Avner, U. 2006. Settlement patterns in the Wadi Arabah and the adjacent desert areas: A view from the Eilat region. In: Bienkowski, P., and Galor, K., eds., Crossing the Rift: Resources, Settlement Patterns and Interaction in the Wadi Arabah. Oxford: 51–74.

Avner, U. and Carmi, I. 2001. Settlement patterns in the Southern Levant deserts during the 6th–3rd millennia BC: A revision cased on 14C dating. Radiocarbon 43: 1203–1216.

Avner, U., Carmi, I. and Segal, D. 1994. Neolithic to Bronze Age settlement of the Negev and Sinai in light of radiocarbon dating: A view from the Southern Negev. In: Bar-Yosef, O. and Kra, R., eds., Late Quaternary Chronology and Paleoclimates of the Eastern Mediterranean. Tucson: 265–300.

Avni, G., Porat, N., and Avni, Y. 2013. Byzantine–Early Islamic agricultural systems in the Negev Highlands: Stages of development as interpreted through OSL dating. Journal of Field Archaeology 38: 332–346.

Avni, Y. 2005. Gully incision as a key factor in desertification in an arid environment, the Negev Highlands, Israel. Catena 63: 185–220.

Avni, Y. and Weiler, N. 2013. Geological Map of Israel: Sede Boqer (Sheet 18-IV) (1:50,000). Geological Survey of Israel, Jerusalem.

Avni, Y., Avni, G. and Porat, N. 2019. A review of the rise and fall of ancient runoff agriculture in the Negev Highlands – A model for the southern Levant deserts. Journal of Arid Environments 163: 127–137.

Avni, Y., Porat, N., and Avni, G. 2012. Pre-farming environment and OSL chronology in the Negev Highlands, Israel. Journal of Arid Environments 86: 12–27.

Avni, Y., Porat, N., Plakht, J. and Avni, G. 2006. Geomorphologic changes leading to natural desertification processes versus anthropogenic land conservation in an arid environment, the Negev Highlands, Israel. Geomorphology 82: 177–200.

Babenko, A.N., Kiseleva, N.K., Plakht, I., Rosen, S., Savinetskii, A.B., and Khasanov, B.F. 2007. Reconstruction of the Holocene vegetation in the Central Negev Desert, Israel, on the basis of palynological data on the Atzmaut Zoogenic Deposit. Russian Journal of Ecology 38: 417–426.

197 Baetan, J., Mees, F., Marinova, E., De Dapper, M., De Vos, D., Huyge, D., Van Strydonck, M., Vandenberghe, D. and Linseele, V. 2018. Late Pleistocene coprolites from Qurta (Egypt and the potential of interdisciplinary research involving micromorphology, plant macrofossil and biomarker analyses. Review of Palaeobotany and Palynology 259: 93–111.

Ball, T., Chandler-Ezell, K., Dickau, R., Duncan, N., Hart, T.C., Iriarte, J., Lentfer, C., Logan, A., Lu, H., Madella, M., Pearsall, D.M., Piperno, D.R., Rosen, A.M., Vrydaghs, L., Weisskopf, A. and Zhang, J. 2016. Phytoliths as a tool for investigations of agricultural origins and dispersals around the world. Journal of Archaeological Science 68: 32–45.

Banning, E.B. and Köhler-Rollefson, I. 1992. Ethnographic lessons from the pastoral past: Camp locations and material remains near Beidha, Southern Jordan. In: Bar-Yosef, O. and Khazanov, A., eds., Pastoralism in the Levant: Archaeological Materials in Anthropological Perspective (Monographs in World Archaeology 10). Madison: 181– 204.

Bar-Matthews, M., Ayalon, A., Gilmour, M., Matthews, A. and Hawksworth, C.J. 2003. Sea- land oxygen isotopic relationships from plantonic foraminifera and speleothems in the Eastern Mediterranean region and their implications for paleorainfall during interglacial intervals. Geochemica et Cosmochimica Acta 67: 3181–3199.

Bar-Yosef, D. 1999. Shells from three Middle Bronze Age I sites in the Negev Highlands. In: Cohen, R. Ancient Settlement of the Central Negev, Vol. I: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports No. 6). Jerusalem: 322–326.

Bar-Yosef, O. and Phillips, J.L., eds. 1977. Prehistoric Investigations in Gebel Maghara Northern Sinai (Qedem 7). Jerusalem.

Bar-Yosef, O., Belfer-Cohen, A., Goren, A., Herskovitz, I., Mienis, H., Sass, B. and Ilan, O. 1986. Nawamis and habitation sites near Gebel Gunna, Southern Sinai. Israel Exploration Journal 27: 65–88.

Baron, A.G. 1978. The Glueck Survey: Issues and Problems in the Archaeology of the Negev. PhD. dissertation, University of Michigan, Ann Arbor. Ann Arbor.

Barth, F. 1961. Nomads of South Persia: The Basseri Tribe of the Khamseh Confederacy. New York.

198 Baruch, U. 1999. Identification of wood remains from Horbat ‘En Ziq. In: Cohen, R., Ancient Settlement of the Central Negev, Vol. I: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports No. 6). Jerusalem: 7*–11*.

Baumgarten, Y. 1993. Ro’i 2. In: Stern, E., ed. New Encyclopedia of Archaeological Excavations in the Holy Land. Jerusalem: 1288.

Beit-Arieh, I. 1977. South Sinai in the Early Bronze Age. PhD. dissertation, Tel Aviv University. Tel Aviv.

Beit-Arieh, I. 2003. Archaeology of Sinai: The Ophir Expedition. Tel Aviv.

Bell, B. 1971. The Dark Ages in ancient history I. The first Dark Age in Egypt. American Journal of Archaeology 75: 1–25.

Ben-Yosef, E. 2018. Provenancing Egyptian metals: A methodological comment. Journal of Archaeological Science 96: 208–215.

Ben-Yosef, E., Gidding, A., Tauxe, L., Davidovich, U., Najjar, M. and Levy, T.E. 2016. Early Bronze Age copper production systems in the Northern Arabah Valley: New insights from archaeomagnetic study of slag deposits in Jordan and Israel. Journal of Archaeological Science 72: 71–84.

Berna, F., Behar, A., Shahack-Gross, R., Berg, J., Boaretto, E., Gilboa, A., Sharon, I., Shalev, S., Shilstein, S., Yahalom-Mack, N., Zorn, J.R. and Weiner, S. 2007. Sediments exposed to high temperatures: reconstructing pyrotechnological processes in Late Bronze and Iron Age strata at Tel Dor (Israel). Journal of Archaeological Science 34: 358–373.

Brochier, J.E. and Thinon, M. 2003. Calcite crystals, starch grains aggregates or…POCC? Comment on ‘calcite crystals inside archaeological plant tissues’. Journal of Archaeological Science 30: 1211–1214.

Brochier, J.E., Villa, P., and Giacomarra, M. 1992. Shepherds and sediments: Geo- ethnoarchaeology of pastoral sites. Journal of Anthropological Archaeology 11: 47– 102.

Bronk Ramsey, C., Dee, M.W., Rowland, J.M., Higham, T.F.G., Harris, S.A., Brock, F., Quiles, A., Wild, E.M., Marcus, E.S., and Shortland, A.J. 2010. Radiocarbon-based chronology for Dynastic Egypt. Science 328: 1554–1557.

199 Bruins, H.J. and van der Plicht, J. 2017. Iron Age agriculture – A critical rejoinder to “Settlement oscillations in the Negev Highlands: the impact of microarchaeological methods”. Radiocarbon 59: 1–16.

Butler, D.H., Lopez-Forment, A. and Dawson, P.C. 2018. Multi-element and biomolecular analyses of soils as a means of sustainable site structure research on hunter-gatherer sites: A case study from the Canadian Artic. Journal of Archaeological Science: Reports 17: 973–991.

Cabanes, D., Mallol, C., Expósito, I. and Baena, J. 2010. Phytolith evidence for hearths and beds in the late Mousterian occupations of Esquilleu cave (Cantabria, ). Journal of Archaeological Science 37: 2947–2957.

Cameron, C.M. and Tomka, S.A. 1993. Abandonment of Settlements and Regions: Ethnoarchaeological and Archaeological Approaches. Cambridge.

Canti, M.G. 1997. An investigation into microscopic calcareous spherulites from herbivore dungs. Journal of Archaeological Science 23: 219–231.

Canti, M.G. 1998. The micromorphological identification of faecal spherulites from archaeological and modern materials. Journal of Archaeological Science 25: 435–444.

Canti, M.G. 1999. The production and preservation of faecal spherulites: Animals, environment and taphonomy. Journal of Archaeological Science 26: 251–258.

Canti, M.G. 2003. Aspects of the chemical and microscopic characteristics of plant ashes found in archaeological soils. Catena 54: 339–361.

Canti, M.G. and Brochier, J.E. 2017. Faecal spherulites. In: Nicosia, C. and Stoops, G.E., eds., Archaeological Soil and Sediment Micromorphology. Hoboken: 51–54.

Carmi, I. and Segal, D. 1992. Rehovot Radiocarbon Measurements IV. Radiocarbon 34: 115– 132.

Chen, F.-W. and Liu, C.-W. 2012. Estimation of spatial rainfall distribution using inverse distance weighting (IDW) in the middle of Taiwan. Paddy and Water Environment 10: 209–222.

Clamer, C. and Sass, B. 1977. Middle Bronze I. In: Bar-Yosef, O. and Philips, J.L., eds., Prehistoric Investigations in Gebel Maghara Northern Sinai (Qedem 7). Jerusalem: 245–254.

200 Cohen-Weinberger, A. and Saidel, B.A. 2014. Ceramic petrography. In: Saidel, B.A. and Haiman, M. Excavations in the Western Negev Highlands: Results of the Negev Emergency Survey 1978–89 (BAR International Series 2684). Oxford: 230–244.

Cohen, R. 1970. Atar Haroa. ‘Atiqot 6: 6–24. (Hebrew).

Cohen, R. 1979. The Iron Age fortresses in the Central Negev. Bulletin of the American Schools of Oriental Research 236: 61–79.

Cohen, R. 1983. The mysterious MBI people. Biblical Archaeology Review 9: 16–29.

Cohen, R. 1985. Archaeological Survey of Israel – Map of Sede Boqer-West (167). Jerusalem.

Cohen, R. 1986. The Settlement of the Central Negev: In the Light of Archaeology and Literary Sources During the 4th–1st Millennium B.C.E. PhD dissertation, Hebrew University of Jerusalem. Jerusalem. (Hebrew with English Summary).

Cohen, R. 1992. The nomadic or semi-nomadic Middle Bronze Age I settlements in the Central Negev. In: Bar-Yosef, O. and Khazanov, A. eds. Pastoralism in the Levant: Archaeological Materials in Anthropological Perspectives (Monographs in World Archaeology 10). Madison: 105–131.

Cohen, R. 1999. Ancient Settlement of the Central Negev, Vol. 1: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports, No. 6). Jerusalem.

Cohen, R. and Dever, W.G. 1978. Preliminary report of the pilot season of the “Central Negev Highlands Project.” Bulletin of the American Schools of Oriental Research 232: 29–45.

Cohen, R. and Dever, W.G. 1979. Preliminary report of the second season of the “Central Negev Highlands Project.” Bulletin of the American Schools of Oriental Research 236: 41–60.

Cohen, R. and Dever, W.G. 1981. Preliminary report of the third and final season of the “Central Negev Highlands Project.” Bulletin of the American Schools of Oriental Research 243: 55–77.

Cooper, L. 2006. The demise and regeneration of Bronze Age urban centers in the Euphrates Valley of Syria. In: Schwartz, G.M. and Nichols, J.J., eds., After Collapse: The Regeneration of Complex Societies. Tucson: 18–37.

Cribb, R. 1991. Nomads in Archaeology. Cambridge.

201 Crouvi, O., Amit, R., Enzel, Y., Porat, N. and Sandler, A. 2008. Sand dunes as a major proximal dust source for Late Pleistocene loess in the Negev Desert, Israel. Quaternary Research 70: 275–282.

Crouvi, O., Amit, R., Porat, N., Gillespie, A.R., McDonald, E.V. and Enzel, Y. 2009. Significance of primary hilltop loess in reconstructing dust chronology, accretion rates, and sources: An example from the Negev Desert, Israel. Journal of Geophysical Research 114: F0127 https://doi.org/10.1029/2008JF001083.

D’Andrea, M. 2012. The Early Bronze IV Period in South-Central Transjordan: Reconsidering chronology through ceramic technology. Levant 44: 17–50.

Danin, A. 1983. Desert Vegetation of Israel and Sinai. Jerusalem.

Danin, A. and Plitman, U. 1987. Revision of the plant geographical territories of Israel and Sinai. Plant Systematics and Evolution 156: 43–53.

Dever, W.G. 1970. The ‘Middle Bronze I’ Period in Syria and Palestine. In: Sanders, J.A., ed., Near Eastern Archaeology in the Twentieth Century: Essays in Honor of Nelson Glueck. Garden City: 132–163.

Dever, W.G. 1971. The peoples of Palestine in the Middle Bronze I Period. The Harvard Theological Review 64: 197–226.

Dever, W.G. 1973. The EB IV–MB I horizon in Transjordan and Southern Palestine. Bulletin of the American Schools of Oriental Research 210: 37–63.

Dever, W.G. 1980. New vistas on the EB IV (“MB I”) horizon in Syria-Palestine. Bulletin of the American Schools of Oriental Research 237: 35–64.

Dever, W.G. 1985. Village planning at Be’er Resism and socio-economic structure in Early Bronze Age IV Palestine. Eretz Israel 18: 18*–28*.

Dever, W.G. 1995. Social structure in the Early Bronze IV Period in Palestine. In: Levy, T.E., ed., The Archaeology of Society in the Holy Land. London: 282–296.

Dever, W.G., and Tadmor, M. 1976. A copper hoard of the Middle Bronze Age I. Israel Exploration Journal 26: 163–169.

Dever. W.G. 2014. Excavations at the Early Bronze IV Sites of Jebel Qa’aqir and Be’er Resisim. Winona Lake.

202 Dunseth, Z.C. 2013. Subsistence Practices in the Negev Highlands during the Intermediate Bronze Age: A Microarchaeological Investigation at Mashabe Sade. M.A. thesis, Tel Aviv University. Tel Aviv.

Dunseth, Z.C. and Shahack-Gross, R. 2018. Calcitic dung spherulites and the potential for rapid identification of degraded animal dung at archaeological sites using FTIR spectroscopy. Journal of Archaeological Science 97: 118–124.

Dunseth, Z.C., Finkelstein, I. and Shahack-Gross, R. 2018. Intermediate Bronze Age subsistence practices in the Negev Highlands, Israel: Macro- and microarchaeological results from the sites of Ein Ziq and Nahal Boqer 66. Journal of Archaeological Science: Reports 19: 712–726.

Dunseth, Z.C., Fuks, D., Langgut, D., Weiss, E., Melamed, Y., Butler, D.H., Yan, X., Boaretto, E., Tepper, Y., Bar-Oz, G. and Shahack-Gross, R. 2019. Archaeobotanical proxies and archaeological interpretation: A comparative study of phytoliths, pollen and seeds in dung pellets and refuse deposits at Early Islamic Shivta, Negev, Israel. Quaternary Science Reviews 211: 166–185.

Dunseth, Z.C., Junge, A., Fuchs, M., Finkelstein, I. and Shahack-Gross, R. 2016. Geoarchaeological investigation in the Intermediate Bronze Age site of Mashabe Sade. Tel Aviv 43: 43–75.

Dunseth, Z.C., Junge, A., Lomax, J., Boaretto, E., Finkelstein, I., Fuchs, M. and Shahack- Gross, R. 2017. Dating archaeological sites in an arid environment: A multi-method case study in the Negev Highlands, Israel. Journal of Arid Environments 144: 156–169.

Eddy, F.W. and Wendorf, F. 1999. An Archaeological Investigation of the Central Sinai, Egypt. Boulder.

Eliyahu-Behar, A., Shilstein, S., Raban-Gerstel, N., Goren, Y., Gilboa, A., Sharon, I., and Weiner, S. 2008. An integrated approach to reconstructing primary activities from pit deposits: Iron smithing and other activities at Tel Dor under Neo-Assyrian domination. Journal of Archaeological Science 35: 2895–2908.

Elliot, S., Bendrey, R., Whitlam, J., Aziz, K.R., and Evans, J. 2015. Preliminary ethnoarchaeological research on modern animal husbandry in Bestansur, Iraqi Kurdistan: Integrating animal, plant and environmental data. Environmental Archaeology 20: 283–303.

203 Enzel, Y., Ken-Tor, R., Sharon, D., Gvirtzman, H., Dayan, U., Ziv, B., Stein, M., 2003. Late Holocene climates of the Near East deduced from Dead Sea level variations and regional winter rainfall. Quaternary Research 60: 263–273.

Esse, D.L. 1991. Subsistence, Trade, and Social Change in Early Bronze Age Palestine. Chicago.

Evenari, M., Aharoni, Y., Shanan, L. and Tadmor, N. 1958. The ancient desert agriculture in the Negev III. Early beginnings. Israel Exploration Journal 8: 231–268.

Evenari, M., Shanan, L. and Tadmor, N. 1982. The Negev: The Challenge of a Desert (2nd Edition). Cambridge.

Falconer, S.E. and Magness-Gardiner, B. 1984. Preliminary report of the first season of the Tell el-Hayyat Project. Bulletin of the American Schools of Oriental Research 255: 49– 74.

Fantalkin, A. and Finkelstein, I. 2006. The Sheshonq I campaign and the 8th-century-BCE earthquake – More on the archaeology and history of the south in the Iron I–IIA. Tel Aviv 33: 18–42.

Finkelstein, I. 1988. Arabian trade and socio-political conditions in the Negev in the twelfth– eleventh centuries B.C.E. Journal of Near Eastern Studies 47: 241–252.

Finkelstein, I. 1989. Further observations on the socio-demographic structure of the Intermediate Bronze Age. Levant 21: 129–140.

Finkelstein, I. 1991. Early Arad: Urbanism of the nomads. Zeitschrift des Deutschen Palästina- Vereins 106: 34–50.

Finkelstein, I. 1995a. Living on the Fringe: the Archaeology and History of the Negev, Sinai and Neighboring Regions in the Bronze and Iron Ages (Monographs in Mediterranean Archaeology 6). Sheffield.

Finkelstein, I. 1995b. Two notes on Early Bronze urbanization and urbanism. Tel Aviv 22: 47– 69.

Finkelstein, I., Adams, M.J., Dunseth, Z.C. and Shahack-Gross, R. 2018. The archaeology and history of the Negev and neighboring areas in the third millennium BCE: A new paradigm. Tel Aviv 45: 63–88.

204 Forget, M.C.L., Regev, L., Friesem, D.E. and Shahack-Gross, R. 2015. Physical and mineralogical properties of experimentally heated chaff-tempered mud bricks: implications for reconstruction of environmental factors influencing the appearance of mud bricks in conflagration events. Journal of Archaeological Science: Reports 2: 80– 93.

Frantzman, S.J. and Kark, R. 2011. Bedouin settlement in Late Ottoman and British Mandate Palestine: Influence on the cultural and environmental landscape, 1870–1948. New Middle Eastern Studies 1: 1–24.

Friesem, D.E. 2016. Geo-ethnoarchaeology in action. Journal of Archaeological Science 70: 145–157.

Friesem, D.E., Tsartsidou, G., Karkanas, P. and Shahack-Gross, R. 2014a. Where are the roofs? A geo-ethnoarchaeological study of mud brick structures and their collapse processes, focusing on the identification of roofs. Archaeological and Anthropological Sciences 6: 73–92.

Friesem, D.E., Zaidner, Y. and Shahack-Gross, R. 2014b. Formation processes and combustion features at the lower layers of the Middle Palaeolithic open-air site of Nesher Ramla, Israel. Quaternary International 331: 128–138.

Fritz, V. 1994. Vorbericht über die Grabungen in Barqā el-Hetīye im Gebiet von Fēnān, Wādī el-‘Araba (Jordanien) 1990. Zeitschrift des Deutschen Palästina-Vereins 110: 125–150.

Frumkin, A. 2009. Stable isotopes of a subfossil Tamarix tree from the Dead Sea region, Israel, and their implications for the Intermediate Bronze Age environment. Quaternary Research 71: 319–328.

Gidding, A. 2016. Approaches to Production and Distribution in Anthropological Archaeology: Views from the Early Bronze Age of Jordan and Israel. University of California, San Diego, PhD. dissertation. San Diego.

Ginguld, M., Perevolotsky, A. and Ungar, E.D. 1997. Living on the margins: Livelihood strategies of Bedouin herd-owners in the Northern Negev, Israel. Human Ecology 25: 567–591.

Gittlen, B.M. 2006. Har Resisim 126: An ephemeral Early Bronze Age site. In: Gitin, S, Wright, J.E., Dessel, J.P., eds., Confronting the Past: Archaeological and Historical Essays on Ancient Israel in Honor of William G. Dever. Winona Lake: 41–59.

205 Glueck, N. 1942. Further explorations in Eastern Palestine. Bulletin of the American Schools of Oriental Research 86: 14–24.

Glueck, N. 1953. Explorations in Western Palestine. Bulletin of the American Schools of Oriental Research 131: 6–15.

Glueck, N. 1955. Further explorations in the Negeb. Bulletin of the American Schools of Oriental Research 137: 10–22.

Glueck, N. 1956. The fourth season of exploration in the Negeb. Bulletin of the American Schools of Oriental Research 142: 17–35.

Glueck, N. 1957. The fifth season of exploration in the Negeb. Bulletin of the American Schools of Oriental Research 145: 11–25.

Glueck, N. 1958a. The sixth season of archaeological exploration in the Negeb. Bulletin of the American Schools of Oriental Research 149: 8–17.

Glueck, N. 1958b. The seventh season of archaeological exploration in the Negeb. Bulletin of the American Schools of Oriental Research 152: 18–38.

Glueck, N. 1960. Archaeological exploration of the Negev in 1959. Bulletin of the American Schools of Oriental Research 159: 3–14.

Glueck, N. 1965. Further explorations in the Negev. Bulletin of the American Schools of Oriental Research 179: 6–29.

Goodfriend, G.A. 1988. Mid-Holocene rainfall in the Negev Desert from 13C of land snail shell organic matter. Nature 333: 757–760.

Goodfriend, G.A. 1990. Rainfall in the Negev Desert during the middle Holocene, based on 13C of organic matter in land snail shells. Quaternary Research 34: 186 – 197.

Goodfriend, G.A. 1991. Holocene trends in 18O in land snail shells from the Negev Desert and their implications for changes in rainfall source areas. Quaternary Research 35: 417– 426.

Goodfriend, G.A. 1999. Stable isotope records of Late Quaternary paleoclimates in the Eastern Mediterranean Region. Quaternary Science Reviews 18: 501–513.

Gophna, R. and Portugali, J. 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.

206 Goren, Y. 1996. The Southern Levant in the Early Bronze Age IV: The petrographic perspective. Bulletin of the American Schools of Oriental Research 303: 33–72.

Goren, Y. 1999. On determining use of pastoral cave sites: A critical assessment of spherulites in archaeology. Journal of Israel Prehistoric Society 29: 123–129.

Goring-Morris, N. 1993. From foraging to herding in the Negev and Sinai: the Early to Late Neolithic transition. Paléorient 19: 65–89.

Govrin, Y. 2002. Archaeological Survey of Israel – Map of Nahal Yatir (139). Jerusalem.

Govrin, Y. 2016. Archaeological Survey of Israel – Map of Qasif (140). https://survey.antiquities.org.il/index_Eng.htm#/MapSurvey/2159

Gur-Arieh, S., Madella, M., Lavi, N. and Friesem, D.E. 2018. Potentials and limitations for the identification of outdoor dung plasters in humid tropical environment: a geo- ethnoarchaeological case study from South India. Archaeological and Anthropological Sciences. https://doi.org/10.1007/s12520-018-0682-y (online ahead of print).

Gur-Arieh, S., Mintz, E., Boaretto, E. and Shahack-Gross, R. 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: 4331–4347.

Gur-Arieh, S., Shahack-Gross, R., Maeir, A.M., Lehmann, G., Hitchcock, L.A., Boaretto, E. 2014. The taphonomy and preservation of wood and dung ashes found in archaeological cooking installations: case studies from Iron Age Israel. Journal of Archaeological Science 46: 50–67.

Haiman, M. 1989. Preliminary report of the Western Negev Highlands Emergency Survey. Israel Exploration Journal 39: 173–191.

Haiman, M. 1991. Archaeological Survey of Israel – Map of Mizpe Ramon Southwest (200). Jerusalem.

Haiman, M. 1994. The Iron Age II sites of the Western Negev Highlands. Israel Exploration Journal 44: 36–61.

Haiman, M. 1996. Early Bronze Age IV settlement pattern of the Negev and Sinai Deserts: View from small marginal temporary sites. Bulletin of the American Schools of Oriental Research 303: 1–32.

207 Hakker-Orion, D. 1999. Faunal remains from Middle Bronze Age I sites in the Negev Highlands. In: Cohen, R. Ancient Settlement of the Central Negev, Vol. I: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (IAA Reports No. 6). Jerusalem: 327–335.

Hammer, Ø., Harper, D.A.T. and Ryan, P.D. 2001. PAST: Paleontological statistics software package for education and data analysis. Palaeontologia Electronica 4. https://palaeo- electronica.org/2001_1/past/issue1_01.htm

Harvey, E.L. and Fuller, D.Q. 2005. Investigating crop processing using phytolith analysis: the example of rice and millets. Journal of Archaeological Science 32: 739–752.

Hauptmann, A. 2007. The Archaeometallurgy of Copper. Berlin.

Hauptmann, A., Schmitt-Strecker, S., Levy, T.E. and Begemann, F. 2015. On Early Bronze Age copper bar ingots from the Southern Levant. Bulletin of the American Schools of Oriental Research 373: 1–24.

Hayden, B. and Cannon, A. 1983. Where the garbage goes: Refuse disposal in the Maya Highlands. Journal of Anthropological Archaeology 2: 117–163.

Hermon, S., Vardi, J. and Rosen, S.A. 2011. The lithic assemblage from the Camel Site. In: Rosen, S.A. An Investigation into Early Desert Pastoralism: Excavations at the Camel Site, Negev.

Hillel, D. and Tadmor, N. 1962. Water regime and vegetation in the Central Negev Highlands of Israel. Ecology 43: 33–41.

Holdorf, P.S. 2010. Comparison of EB IV radiocarbon results from Khirbat Iskandar and Bab adh-Dhra. In: Richard, S., Long, J.C., Holdorf, P.S. and Peterman, G., eds. Khirbat Iskandar: Final Report on the Early Bronze Age IV Area C “Gateway” and Cemeteries (ASOR Reports 14). Boston: 267–270.

Holzer, A., Avner, U., Porat, N. and Kolska Horwitz, L. 2010. Desert kites in the Negev desert and northeast Sinai: Their function, chronology and ecology. Journal of Arid Environments 74: 806–817.

Hölfmayer, F., Dee, M.W., Genz, H. and Riehl, S. 2014. Radiocarbon evidence fro the Early Bronze Age Levant: the site of Tell Fadous-Kfarabida ()) and the end of the Early Bronze III Period. Radiocarbon 56: 529–542.

208 Hütteroth, W.D. and Abdulfattah, K. 1977. Historical Geography of Palestine, Transjordan and Southern Syria in the Late 16th (Sixteenth) Century (Vol. 5). Erlangen.

Ilan, O. and Sebanne, M. 1989. Copper metallurgy, trade and urbanization of southern Canaan in the Chalcolithic and Early Bronze Age. In: de Miroschedji, P., ed., L’urbanisation de la Palestine à l’âge du Bronze ancien (BAR International Series 527). Oxford: 139– 162.

Junge, A., Lomax, J., Shahack-Gross, R., Finkelstein, I. and Fuchs, M. 2018. Chronology of an ancient water reservoir and the history of human activity in the Negev Highlands, Israel. Geoarchaeology 33: 695–707.

Kafle, H.K. and Bruins, H.J. 2009. Climatic trends in Israel 1970–2002: Warmer and increasing aridity inland. Climatic Change 96: 63–77.

Kagan, E.J., Langgut, D., Boaretto, E., Neumann, F.H. and Stein, M. 2015. Dead Sea levels during the Bronze and Iron Ages. Radiocarbon 57: 237–252.

Karkanas P. 2006. Late Neolithic household activities in marginal areas: the micromorphological evidence from the Kouveleiki caes, Peloponnese, Greece. Journal of Archaeological Science 33: 1628–1641.

Karkanas, P. and Goldberg, P. 2019. Reconstructing Archaeological Sites: Understanding the Geoarchaeological Matrix. Oxford.

Karkanas, P., Shahack-Gross, R., Ayalon, A., Bar-Matthews, M., Barkai, R., Frumkin, A., Gopher, A. and Stiner, M. 2007. Evidence for habitual use of fire at the end of the Lower Paleolithic: Site-formation processes at Qesem Cave, Israel. Journal of Human Evolution 53: 197–212.

Katz, O., Cabanes, D., Weiner, S., Maeir, A.M., Boaretto, E. and Shahack-Gross, R. 2010. Rapid phytolith extraction for analysis of phytolith concentrations and assemblages during an excavation: an application at Tell es-Safi/Gath, Israel. Journal of Archaeological Science 37: 1557–1563.

Kennedy, M.A. 2016. The End of the 3rd Millennium BC in the Levant: New Perspectives and Old Ideas. Levant 48: 1–32.

Kenyon, K. 1951. Some Notes on the History of Jericho in the Second Millennium B.C. Palestine Exploration Quaterly 83: 101–138.

209 Kenyon, K. 1956. Tombs of the Intermediate Early Bronze-Middle Bronze at Tell el-Ajjul. Annual of the Department of Antiquities of Jordan 3: 41–55.

Kenyon, K. 1966. Amorites and Canaanites: The Schweich Lectures of the British Academy 1963

Khazanov, A.M. 1984. Nomads and the Outside World. Cambridge.

Kidron, G.J. 1999. Altitude dependent dew and fog in the Negev Desert. Agricultural and Forest Meteorology 96: 1–8.

Kmošek, J., Odler, M., Fikrie, M. and Kochergina, Y.V. 2018. Invisible connections. Early Dynastic and Old Kingdom Egyptian metalwork in the Egyptian Museum of Leipzig University. Journal of Archaeological Science 96: 191–207.

Kochavi, M. 1967. The Settlement of the Negev in the Middle Bronze (Canaanite) I Age. PhD dissertation, Hebrew University of Jerusalem. Jerusalem (Hebrew with English abstract).

Kochavi, M. 2009. The Intermediate Bronze Age (IBA) in the Negev, 40 years later. In: Parr, P.J., ed., Levant in Transition: Proceedings of a Conference Held at the British Museum on 20–24 April, 2004. London: 43–48.

Kociba, K.J. and Gallagher, P.K. 1996. A study of calcium oxalate monohydrate using dynamic differential scanning calorimetry and other thermoanalytical techniques. Thermochimica Acta 282–283: 277–296.

Korstanje, M.A. 2005. Microfossils in Camelid dung: Taphonomic considerations for the archaeological study of agriculture and pastoralism. In: O’Connor, T., ed., Biosphere to Lithosphere: New studies in vertebrate taphonomy. Proceedings of the 9th International Council of Archaeozoology, Durham, August 2002. Oxford: 69-77.

Kozloff, B. 1974. A brief note on the lithic industries of Sinai. Museum Ha’aretz Yearbook 15/16: 35-49.

Kraemer, C.J. 1958. Non-literary papyri. In: Colt, H.D., ed., Excavations at Nessana (Auja Hafir, Palestine), Vol. III. Princeton.

Langgut, D., Finkelstein, I., Litt, T., Neumann, F.H. and Stein, M. 2015. Vegetation and climate changes during the Bronze and Iron Ages (~3600-600 BCE) in the southern Levant based on palynological records. Radiocarbon 57: 217-235.

210 Langgut, D., Neumann, F.H., Stein, M., Wagner, A., Kagan, E.J., Boaretto, E. and Finkelstein, I. 2014. Dead Sea pollen record and history of human activity in the Judean Highlands (Israel) from the Intermediate Bronze Age into the Iron Ages (~2500-500 BCE). Palynology 38: 280-302.

Laugomer, B., Bar-Matthews, M., Finkelstein, I. and Langgut, D. Forthcoming. The climate of the Bronze and Iron Ages in the Southern Levant: A reconstruction based on Soreq Cave speleothems.

Levy, T.E., Adams, R.B., Hauptmann, A., Prange, M., Schmitt-Strecker, S., and Najjar, M. 2002. Early Bronze Age Metallurgy: A Newly Discovered Copper Manufactory in Southern Jordan. Antiquity 75: 425-437.

Levy, T.E., Najjar, M., Higham, T., Arbel, Y., Muniz, A., Ben-Yosef, E., Smith, N.G., Beherec, M., Gidding, A., Jones, I.W., Frese, D., Smitheram, C. and Robinson, M. 2014. Excavations at Khirbet en-Nahas 2002–2009: An Iron Age copper production center in the lowlands of Edom. In: Levy, T.E., Najjar, M. and Ben-Yosef, E. New Insights into the Iron Age Archaeology of Edom, Southern Jordan, Vol. 1. Los Angeles: 89–245.

Liao, K.-H., Xu, S.-H., Wu, J.-C, Ji, S.-H. and Qing, L. 2011. Cokriging of soil cation exchange capacity using the first principal component derived from soil physio-chemical properties. Agricultural Sciences in China 10: 1246-1253.

Liphschitz, N. 2003. Archaeobotanical remains. In: Beit-Arieh, I. Archaeology of Sinai. Tel Aviv: 257–262.

Litt, T., Ohlwein, C., Neumann, F.H., Hense, A. and Stein, M. 2012. Holocene Climate Variability in the Levant from the Dead Sea Pollen Record. Quaternary Science Reviews 49: 95-105.

Liverani, M. 1970. Per una considerazione storica del problema amorreo. Oriens Antiquus 9: 5-27.

Liverani, M. 1973. The Amorites. In: Wiseman, D.J. Peoples of the Old Testament Times. Oxford: 100-133.

London, G.A. 1985. Decoding Designs: The Late Third Millennium B.C. Pottery from Jebel Qa‘aqir. PhD. dissertation, University of Arizona. Tucson.

MacDonald, B. 1992. The Southern Ghors and Northeast ‘Arabah Archaeological Survey (Sheffield Archaeological Monographs 5). Sheffield.

211 MacDonald, B., ed. 1988. The Wadi Hasa Archaeological Survey (1979-1983), West-Central Jordan. Waterloo.

MacDonald, B., Rollefson, G.O., Banning, E.B., Byrd, B.F., and D’Annibale, C. 1983. The Wadi el Hasa Archaeological Survey 1982: A Preliminary Report. Annual of the Department of Antiquities of Jordan 27: 311-323.

Madella, M., Alexandre, A., and Ball, T. 2005. International Code for Phytolith Nomenclature 1.0. Annals of Botany 96: 253-260.

Malek, J. 2000. The Old Kingdom (c. 2686-2160 BC). In Shaw, I. ed. The Oxford History of Ancient Egypt. Oxford: 83-107.

Mallol, C., Marlowe, F.W., Wood, B.M. and Porter, C.C. 2007. Earth, wind, and fire: Ethnoarchaeological signals of Hadza fires. Journal of Archaelogical Science 34: 2035- 2052.

Marx, E. 1967. Bedouin of the Negev. Manchester.

Marx, E. 1992. Are there pastoral nomads in the Middle East? In: Bar-Yosef, O. and Khazanov, A., eds., Pastoralism in the Levant: Archaeological Materials in Anthropological Perspectives (Monographs in World Archaeology 10). Madison: 255-260.

Matthews, W., French, C.A.I., Lawrence, T., Cutler, D.F., and Jones, M.K. 1997. Microstratigraphic Traces of Site Formation Processes and Human Activities. World Archaeology 29: 281-308.

Mayerson, P. 1962. The Ancient Agricultural Regime of Nessana and the Central Negeb. In: Colt, H.D., ed., Excavations at Nessana (Auja, Hafir, Palestine), Vol I. London: 211- 269.

Migowski, C., Stein, M., Prasad, S., Negendank, J.F.W., and Agnon, A. 2006. Holocene climate variability and cultural evolution in the Near East from the Dead Sea sedimentary record. Quaternary Research 66: 421-431.

Miller, N.F. 1984. The use of dung as fuel: an ethnographic example and an archaeological application. Paléorient 10: 71-79.

Muniz, A.A. 2007. Feeding the Periphery: Modeling Early Bronze Age Economies and the Cultural Landscape of the Faynan District, Southern Jordan. PhD. dissertation, University of California, San Diego. San Diego.

212 Musil, A. 1908. Arabia Petraea III. Vienna.

Musil, A. 1928. The Manners and Customs of the Rwala Bedouins. New York.

Nissenbaum, A., Serban, A. and Connan, J. 1999. Asphalt from Horbat En Ziq. In: Cohen, R., Ancient Settlement of the Central Negev, Vol. I: The Chalcolithic Period, the Early Bronze Age and Middle Bronze Age I (IAA Reports 6). Jerusalem: 12-16.

Noy, T. and Cohen, R. 1974. Nahal Boker: an early Pre-Pottery Neolithic-B site. Mitekufat Haeven: Journal of the Israel Prehistoric Society 13: 15-25.

Oren, E.D. 1973. The overland route between Egypt and Canaan in the Early Bronze Age (preliminary report). Israel Exploration Journal 23: 198-205.

Oren, E.D. 1989. Early Bronze Age settlement in North Sinai: A model for Egypto-Canaanite interconnections. In: de Miroschedji, P., ed., L’urbanisation de la Palestine à l’âge du Bronze ancien (BAR International Series 527). Oxford: 389-405.

Oren, E.D. and Yekutieli, Y. 1990. North Sinai during the MB I Period – Pastoral Nomadism and Sedentary Settlement. Eretz-Israel 21: 6-22 (Hebrew with English summary).

Palmer, C., Smith, H. and Daly, P. 2007. Ethnoarchaeology. In: Barker, G., Gilbertson, D. and Mattingly, D., eds., Archaeology and Desertification: The Wadi Faynan Landscape Survey, Southern Jordan. Oxford: 369-395.

Palmer, E.H. 1871. Desert of the Exodus. Cambridge.

Palumbo, G. 1990. The Early Bronze Age IV in the Southern Levant: Settlement Patterns, Economy, and Material Culture of a ‘Dark Age.’ Rome.

Palumbo, G. 2008. The Early Bronze Age IV. In: Adams, R.B., ed., Jordan: An Archaeological Reader. London: 227-262.

Piperno, D. 2006. Phytoliths: A Comprehensive Guide for Archaeologists and Paleoecologists. Lanham.

Polo-Díaz, A., Alonso Eguíluz, M., Ruiz, M., Pérez, S., Mújika, J., Albert, R.M. and Fernández Eraso, J. 2016. Management of residues and natural resources at San Cristóbal rock- shelter: Contribution to the characterization of chalcolithic agropastoral groups in the Iberian Peninsula. Quaternary International 414: 202-225.

Porat, N. 1984.

213 Porat, N. 1989. Composition of Pottery—Application to the Study of the Interrelations between Canaan and Egypt during the 3rd Millennium B.C. PhD. dissertation, Hebrew University of Jerusalem. Jerusalem.

Portillo, M., Belarte, M.C., Ramon, J., Kallala, N., Sanmartí, J. and Albert, R.M. 2017. An ethnoarchaeological study of livestock dung fuels from cooking installations in northern Tunisia. Quaternary International 431: 131-144.

Portillo, M., Kadowski, S., Nishiaki, Y. and Albert, R.M. 2014. Early Neolithic household behavior at Tell Seke al-Aheimar (Upper Khabur, Syria): a comparison to ethnoarchaeological study of phytoliths and dung spherulites. Journal of Archaeological Science 42: 107-118.

Portillo, M., Valenzuela, S. and Albert, R.M. 2012. Domestic patterns in the Numidian site of Althiburos (northern Tunisia): the results from the combined study of animal bones, dung and plant remains. Quaternary International 275: 85-96.

Prag, K. 1985. Ancient and Modern Pastoral Migration in the Levant. Levant 17: 81-88.

Prins, A.B. 2016. 3D modelling for archaeological documentation: using the JVRP method to record archaeological excavations with millimeter-accuracy (version 3.0). JVRP White Papers in Archaeological Technology. https://www.jezreelvalleyregionalproject.com/3d-modeling (accessed 10 March 2018).

Prost, K., Birk, J.J., Lehndorff, E., Gerlach, R. and Amelung, W. 2017. Steroid biomarkers revisited – Improved source identification of faecal remains in archaeological soil material. PLoSONE 12: e0164882. https://doi.org/10.1371/journal.pone.0164882

Rademakers, F.W., Verly, G., Delvaux, L. and Degryse, P. Copper for the afterlife in Predynastic to Old Kingdom Egypt: Provenance characterization by chemical and lead isotope analysis (RMAH collection, Belgium). Journal of Archaeological Science 96: 175-190.

Rapp, G. and Mulholland, S.C. 1992. Phytolith Systematics: Emerging Issues. New York.

Rast, W.E. and Schaub, R.T. 2003. Bâb edh-Dhrâc: Excavations at the Town Site (1975-1981) (Reports of the Expedition to the Dead Sea Plain, Jordan Volume 2). Winona Lake.

Reddy, S.N. 1999. Fueling the hearths in India: The role of dung in paleoethnobotanical interpretation. Paléorient 24: 61-70.

214 Regev, J., Miroschedji, de P., Greenberg, R., Braun, E., Greenhut, Z., Boaretto, E. 2012. Chronology of the Early Bronze Age in the Southern Levant: New Analysis for a High Chronology. Radiocarbon 54: 525-566.

Regev, L., Cabanes, D., Homsher, R., Kleiman, A., Weiner, S., Finkelstein, I. and Shahack- Gross, R. 2015. Geoarchaeological investigation in a domestic Iron Age quarter, Tel Megiddo, Israel. Bulletin of the American Schools of Oriental Research 374: 135-157.

Regev, L., Poduska, K.M., Addadi, L., Weiner, S., and Boaretto, E. 2010. Distinguishing between calcites formed by different mechanisms using infrared spectrometry: archaeological applications. Journal of Archaeological Science 37: 3022-3029.

Richard, S. 1980. Towards a Consensus of Opinion on the End of the Early Bronze Age in Palestine-Transjordan. Bulletin of the American Schools of Oriental Research 237: 5- 34.

Richard, S. and Boraas, R. 1988. The Early Bronze IV fortified site of Khirbet Iskander, Jordan: third preliminary report, 1984 season. Bulletin of the American Schools of Oriental Research Supplementary Studies 25: 107-130.

Richard, S., Long, J.C., Holdorf, P.S., and Peterman, G. 2010. Khirbet Iskander: Final Report on the Early Bronze IV Area C “Gateway” and Cemeteries. Boston.

Rodriguez-Navarro, C., Jimenez-Lopez, C., Rodriguez-Navarro, A., Gonzalez-Munoz, M.;T., and Roiriguez-Gallego, M. 2007. Bacterially mediated mineralization of vaterite. Geochemica et Comochimica Acta 71: 1197-1213.

Rondelli, B., Lancelotti, C., Madella, M., Pecci, A., Balbo, A., Ruiz Pérez, J., Inserra, F., Gadekar, C., Cau Ontiveros, M.A. and Ajithprasad, P. 2014. Anthropic activity markers and spatial variability: an ethnoarchaeological experiment in a domestic unit of Northern Gujarat (India). Journal of Archaeological Science 41: 482-592.

Rosen, A.M. 1989. Environmental Change at the End of Early Bronze Age Palestine. In: de Miroschedji, P. ed L’urbanisation de la Palestine à l’âge du Bronze ancient: Part ii (BAR International Series 527), Oxford: 247-255.

Rosen, A.M. 1995. The Social Response to Environmental Change in Early Bronze Age Canaan. Journal of Anthropological Archaeology 14: 26-44.

Rosen, A.M. 1999. Past environments and farming potential at Horbat ‘En Ziq. In: Cohen, R. Ancient Settlement of the Central Negev, Vol. 1: The Chalcolithic Period, The Early

215 Bronze Age and the Middle Bronze Age I (Israel Antiquities Authority Reports 6). Jerusalem: 17-21.

Rosen, S.A. 1987. Demographic Trends in the Negev Highlands: Preliminary Results from the Emergency Survey. Bulletin of the American Schools of Oriental Research 266: 45-58.

Rosen, S.A. 1992. Nomads in archaeology: a response to Finkelstein and Perevolotsky. Bulletin of the American Schools of Oriental Research 287: 75-85.

Rosen, S.A. 1997. Lithics after the Stone Age: A Handbook of Stone Tools from the Levant. Walnut Creek.

Rosen, S.A. 2011a. The Desert and the Pastoralist: An Archaeological Perspective on Human- Landscape Interaction in the Negev over the Millennia. Annals of Arid Zone 50: 1-15.

Rosen, S.A. 2011b. An Investigation into Early Desert Pastoralism: Excavations at the Camel Site, Negev. Los Angeles.

Rosen, S.A. 2016. Revolutions in the Desert: The Rise of Mobile Pastoralism in the Southern Levant. London.

Rosen, S.A. and Avni, G. 1997. The ‘Oded Sites: Investigations of Two Early Islamic Pastoral Camps South of the Ramon Crater. Jerusalem.

Rosen, S.A. and Gopher, A. 2003. II: Flint tools from the survey. In: Beit-Arieh, I. Archaeology of Sinai: the Ophir Expedition. Tel Aviv: 184-194.

Rosen, S.A. and Vardi, J. 2014. Appendix IIH: The Chipped Stone Assemblage from Be’er Resisim: A Final Report. In: Dever. W.G., Excavations at the Early Bronze IV Sites of Jebel Qa’aqir and Be’er Resisim. Winona Lake: 327-338.

Rosen, S.A., Savinetsky, A.B., Plakht, Y., Kisseleva, N.K., Khassanov, B.F., Pereladov, A.M., and Haiman, M. 2005. Dung in the Desert: Preliminary Results of the Negev Holocene Ecology Project. Current Anthropology 46: 317-326.

Rosen, S.A., Shugar, A. and Vardi, J. 2014. Function and value in sickle segment analysis: Odellian perspectives. In: Shott, M.J., ed., Works in Stone: Contemporary Perspectives on Lithic Analysis. Salt Lake City: 116-130.

Rothenberg, 1970. An archaeological survey of South Sinai. Palestine Exploration Quarterly 102: 4-29.

216 Rothenberg, B. 1999. Archaeo-metallurgical researches in the Southern Arabah 1959-1990, Part I: Late Pottery Neolithic to Early Bronze IV. Palestine Exploration Quarterly 131: 68-89.

Rothenberg, B. and Glass, J. 1992. The beginning and development of early metallurgy and the settlement and chronology of the Western Arabah, from the Chalcolithic Period to Early Bronze Age IV. Levant 24: 141-157.

Rowton, M.B. 1977. Dimorphic structure and the parasocial element. Journal of Near Eastern Studies 36: 181-198.

Saidel, B.A. 2002a. The excavations at Rekhes Nafha 396 in the Negev Highlands, Israel. Bulletin of the American Schools of Oriental Research 325: 37-63.

Saidel, B.A. 2002b. Pot luck? Variation and function in the ceramic assemblages of pre-camel pastoralists in the Negev Highlands, Israel. Mitekufat Haeven: Journal of the Israel Prehistoric Society 32: 175-196.

Saidel, B.A. 2004. Vessel functions in agricultural and pastoral societies of Byzantine and Early Islamic Israel. Journal of Field Archaeology 29: 437-445.

Saidel, B.A. and Haiman, M. eds. 2014. Excavations in the Western Negev Highlands: Results of the Negev Emergency Survey 1979–1989 (BAR International Series 2684). Oxford.

Saidel, B.A., Erickson-Gini, T.,Vardi, J., Rosen, S.A., Maher, E., and Greenfield, H. 2006. Test excavations at Rogem Be’erotayim in Western Negev. Journal of the Israel Prehistoric Society 36: 201-229.

Schiegl, S., Goldberg, P., Bar-Yosef, O. and Weiner, S. 1996. Ash deposits in Hayonim and Kebara Caves, Israel: Macroscopic, microscopic and mineralogical observations and their archaeological implications. Journal of Archaeological Science 23: 763-781.

Sebbane, M., Ilan, O., Avner, U. and Ilan, D. 1993. The dating of Early Bronze Age settlements in the Negev and Sinai. Tel Aviv 20: 41–54.

Segal, D. 1999. Carbon-14 dating from Early and Middle Bronze Ages sites in Israel and the Region. In: Cohen, R. Ancient Settlement of the Central Negev, Vol. 1: The Chalcolithic Period, The Early Bronze Age and the Middle Bronze Age I (Israel Antiquities Authority Reports 6). Jerusalem: 336–339.

Segal, D. and Carmi, I. 1996. Rehovot radiocarbon date list V. ‘Atiqot 29: 79-106.

217 Segal, I. and Carmi, I. 2004. Rehovot radiocarbon date list VI. ‘Atiqot 48: 123-148.

Segal, I. and Halicz, L. 2005. Provenance studies in archaeometallurgy using lead isotopes ratio determination by Q-ICP-MS and MC-ICP-MS. Israel Journal of Earth Science 54: 87-96.

Segal, I., Roman, I., Cohen, R., and Brenner, I.B. 1996–1997. Chemical and metallurgical studies of ‘En Ziq and Be’er Resisim ingots. Arx 2–3: 43–51.

Seligman, N., Tadmor, N., and Raz, Z. 1962. Range survey of the Central Negev. The National and University Institute of Agriculture, Bulletin 67.

Shahack-Gross, R. 2011. Herbivorous livestock dung: formation, taphonomy, methods for identification and archaeological significance. Journal of Archaeological Science 38: 205–218.

Shahack-Gross, R. 2017. Archaeological formation theory and geoarchaeology: State-of-the- art in 2016. Journal of Archaeological Science 79: 36–43.

Shahack-Gross, R. 2019. Fire and society in the Eastern Mediterranean. In: Yassur-Landau, A., Cline, E.H. and Rowan, Y., eds., The Social Archaeology of the Levant. Cambridge: 86–97.

Shahack-Gross, R. and Ayalon, A. 2013. Stable carbon and oxygen isotopic compositions of wood ash: an experimental study with archaeological implications. Journal of Archaeological Science 40: 570–578.

Shahack-Gross, R. and Finkelstein, I. 2008. Subsistence practices in an arid environment: a geoarchaeological investigation in an Iron Age site, the Negev Highlands, Israel. Journal of Archaeological Science 35: 965–982.

Shahack-Gross, R. and Finkelstein, I. 2015. Settlement oscillations in the Negev Highlands revisited: the impact of microarchaeological methods. Radiocarbon 57: 253–264.

Shahack-Gross, R. and Finkelstein, I. 2017. Iron Age agriculture in the Negev Highlands? Methodological and factual comments on Bruins and van der Plicht 2017a (Radiocarbon Vol. 59. NR.1). Radiocarbon 59: 1227–1231.

Shahack-Gross, R., Boaretto, E., Cabanes, D., Katz, O. and Finkelstein, I. 2014. Subsistence economy in the Negev Highlands: The Iron Age and Byzantine/Early Islamic Period. Levant: 46: 98–117.

218 Shahack-Gross, R., Marshall, F., and Weiner, S., 2003. Geo-ethnoarchaeology of pastoral sites: the identification of livestock enclosures in abandoned Maasai settlements. Journal of Archaeological Science 30: 439–459.

Shanan, L. 2000. Runoff, erosion, and the sustainability of ancient irrigation systems in the Central Negev Desert. In: Hassan. M.A., Slaymaker, O., Berkowicz, S.M., eds., The Hydrology–Geomorphology Interface: Rainfall, Floods, Sedimentation, Land Use. Jerusalem: 75–106.

Sharon, D. 1972. The spottiness of rainfall in a desert area. Journal of Hydrology 17: 161–175.

Sharon, D. and Kutiel, H. 1986. The distribution of rainfall intensity in Israel, its regional and seasonal variations and its climatological evaluation. Journal of Climatology 6: 277– 291.

Shillito, L.-M., Bull, I.D., Matthews, W., Almond, M.J., Williams, J.M. and Evershed, R.P. 2011. Biomolecular and micromorphological analysis of suspected faecal deposits at Neolithic Çatalhöyük, Turkey. Journal of Archaeological Science 38: 1869–1877.

Sillar, B. 2000. Dung by preference: The choice of fuel as an example of how andean pottery production is embedded within wider technical, social, and economic practices. Archaeometry 42: 43–60.

Simms, S.R. 1988. The archaeological structure of a Bedouin camp. Journal of Archaeological Science 15: 197–211.

Smith, A., Proctor, L., Hart, T.C. and Stein, G.J. 2018. The burning issue of dung in archaeobotanical samples: a case-study integrating macro-botanical remains, dung spherulites, and phytoliths to assess sample origin and fuel use at Tell Zeidan, Syria. Vegetation History and Archaeobotany https://doi.org/10.1007/s00334-018-0692-9 (online ahead of print).

Starinsky, A., Zilberman, E., Braun, M. and Sneh, A. 2010. Geological Map of Israel: Revivim (Sheet 18-II) (1:50,000). Geological Survey of Israel, Jerusalem.

Stern, E., ed. 2008. The New Encyclopedia of Archaeological Excavations in the Holy Land: Supplementary Volume 5. Jerusalem.

Tsartsidou, G., Lev-Yadun, S., Albert, R.M., Rosen, A.M., Efstratiou, N. and Weiner, S. 2007. The phytolith archaeological record: strengths and weaknesses evaluated based on a quantitative modern reference collection from Greece. Journal of Archaeological

219 Science 34: 1262–1275.

Tufnell, 1958. Lachish IV: The Bronze Age. London.

Twiss, P.C., Suess, E. and Smith, R.M. 1969. Morphological classification of grass phytoliths. Social Science Society of America Proceedings 33: 109–115.

Valamoti, S.M., and Charles, M. 2005. Distinguishing food from fodder through the study of charred plant remains: An experimental approach to dung-derived chaff. Vegetation History and Archaeobotany 14: 528–533.

Vardi, J. 2005. The Analysis of the Lithic Assemblage from Ein Ziq, An Early Bronze IV (2300– 2000 BCE) Site in the Negev Highlands. MA thesis, Ben Gurion University of the Negev. Beersheva. Vardi, J. 2014. The lithic assemblages. In: Saidel, B.A. and Haiman, M., eds., Excavations in the Western Negev Highlands: Results of the Emergency Survey 1978–89. Oxford: 59– 168.

Vardi, J., Rosen, S.A., and Hermon, S. 2007. The economy of the Early Bronze Age IV Period (c. 2200–2000): The lithic evidence. In: Saidel, B. and van der Steen, E.J., eds., On the Fringe of Society: Archaeological and Ethnographic Perspectives on Pastoral and Agricultural Societies (BAR International Series 1657). London: 99–116.

Vardi, J., Shilstein, S., Shalev, S., and Yekutieli, Y. 2008. The Early Bronze Age IV chipped and ground stone assemblages of ‘En Yahav and its relation to copper smelting activities. Journal of the Israel Prehistoric Society 38: 1–20.

Vogel, J.C., Fuls, A., and Danin, A. 1986. Geographical and environmental distribution of C3

and C4 grasses in the Sinai, Negev, and Judean Deserts. Oecologia 70: 258–265.

Warnock, P.J. 1991. Analysis of Wood Charcoal from Be’er Resisim, The “Central Negev Highlands Project,” Israel: Implications for Economic and Cultural Strategies. M.A. thesis, Texas A&M University. College Station.

Warnock, P.J. 2014. Appendix IID: Analysis of Wood Charcoal. In: Dever. W.G., Excavations at the Early Bronze IV Sites of Jebel Qa’aqir and Be’er Resisim. Winona Lake: 301– 310.

Wattez, J. and Courty, M.A. 1987. Morphology of ash of some plant materials. In: Fedoroff, N., Bresson, L.M., and Courty, M.A., eds., Soil Micromorphology. Plaisir: 677–683.

Weiner, S. 2010. Microarchaeology: Beyond the Visible Archaeological Record. Cambridge.

220 Weiner, S., Goldberg, P. and Bar-Yosef, O. 1993. Bone preservation in Kebara Cave, Israel using on-site Fourier-transform infrared spectrometry. Journal of Archaeological Science 20: 613–627.

Weinstein, J.M. 1984. Radiocarbon dating in the Southern Levant. Radiocarbon 26: 297–366.

Weinstein, J.M. 2003. A new set of radiocarbon dates from the town site. In: Rast, W.E. and Schaub, R.T. 2003. Bâb edh-Dhrâc: Excavations at the Town Site (1975-1981) (Reports of the Expedition to the Dead Sea Plain, Jordan Volume 2). Winona Lake: 638–648.

Weiss, H., Courty, M.-A. Wetterstrom, W., Guichard, F., Senior, L., Meadow, R. and Curnow, A. 1993. The genesis and collapse of third millennium North Mesopotamian civilization. Science 261: 995–1004.

Wilkinson, T.J., Philip, G., Bradbury, J., Dunford, R., Donoghue, D., Galiatsatos, N., Lawrence, D., Ricci, A. and Smith, S.L. 2014. Contextualizing early urbanism: settlement cores, early states and agro-pastoral strategies in the Fertile Crescent during the fourth and third millennia BC. Journal of World Prehistory 27: 43–109.

Woolley, L.C and Lawrence, T.E. 1914. The Wilderness of Zin (Archaeological Report). London.

Xie, X., Yan, M. and Wang, C. 1989. Geochemical standard reference samples GSD 9–12, GSS 1–8 and GSR 1–6. Geostandards Newsletter 13: 83–179.

Xie, X., Yan, N., Li, L., and Shen, H. 1985. Usable values for Chinese standard reference samples of stream sediments, soils and rocks: GSD 9–12, GSS 1–8 and GSR 1–6. Geostandards Newsletter 9: 277–280.

Xu, B., Toffolo, M.B., Regev, L., Boaretto, E. and Poduska, K.M. 2015. Structural differences in archaeologically relevant calcite. Analytical Methods 7: 9304–9309.

Yahalom-Mack, N., Galili, E., Segal, I., Eliyahu-Behar, A., Boaretto, E., Shilstein, S. and Finkelstein, I. 2014. New insights into Levantine copper trade: analysis of ingots from the Bronze and Iron Ages in Israel. Journal of Archaeological Science 45: 159–177.

Yekutieli, Y., Shilstein, S., and Shalev, S. 2005. ‘En Yahav: A copper smelting site in the ‘Arava. Bulletin of the American Schools of Oriental Research 340: 1–21.

221 Zhang, C., Lv, J., Li, F. and Li, X. 2017. Nucleation and growth of Mg-calcite spherulites induced by the bacterium Curvibacter lanceolatus strain HJ-1. Microscopy and Microanalysis 23: 1189–1196.

Zilberman, E. and Avni, Y. 2004. Geological Map of Israel: Mizpe Ramon (Sheet 21-II) (1:50,000). Geological Survey of Israel, Jerusalem.

Ziv, B., Saaroni, H., Pargament, R., Harpaz, T. and Alpert, P. 2013. Trends in rainfall regime over Israel, 1975–2010, and their relationship to large-scale variability. Regional Environmental Change 13: 1–14. Zohary, M. 1962. Plant Life of Palestine: Israel and Jordan. New York

Zurro, D. 2018. One, two, three phytoliths: assessing the minimum phytolith sum for archaeological studies. Archaeological and Anthropological Sciences 10: 1673–1691.

222 11 Appendix A: Mineralogical and microremain data for all Nahal Nizzana 332.1 samples.

Ash Dung Phytolith Pseudomorph Spherulite Locus Square Sample # Context Description Mineralogy Concentrations Concentrations PSR Concentrations (million/g) (million/g) (million/g)

17/NIZ/01 N16 N16.1 Outside structure Surface sediments Ca Cl (ua) Dol Q 0.48 0.05 0.00 0 17/NIZ/01 N17 N17.1 Outside structure Surface sediments Ca Cl (ua) Dol Q 0.17 0.24 0.00 0 17/NIZ/01 N18 N18.1 Outside structure Surface sediments Ca Cl (ua) Dol Q 0.01 0.26 0.00 0 17/NIZ/01 N19 N19.1 Outside structure Surface sediments Ca Cl (ua) Dol Q 0.14 0.09 0.00 0 17/NIZ/02 N10 N10.1 Courtyard Surface sediments Ca Cl (ua) Dol Q 0.17 5.63 0.00 0 17/NIZ/02 N11 N11.1 Courtyard Surface sediments Ca Cl (ua) Dol Q 0.29 1.08 0.00 0 17/NIZ/02 N12 N12.1 Courtyard Surface sediments Ca Cl (ua) Dol Q 0.09 1.18 0.00 0 17/NIZ/02 N14 N14.1 Courtyard Surface sediments Ca Cl (ua) Dol Q 0.40 0.93 0.00 0 17/NIZ/02 N15 N15.1 Courtyard Surface sediments Ca Cl (ua) Dol Q 0.17 2.72 0.00 0 17/NIZ/02 N16 N16.4 Courtyard Surface sediments Ca Cl (ua) Dol Q 0.20 0.18 0.00 0 17/NIZ/03 G16 G16.1 Inside Structure Surface sediments Ca Cl (ua) Dol Q 0.12 0.17 0.00 0 17/NIZ/03 H15 H15.1 Inside Structure Surface sediments Ca Cl (ua) Dol Q 0.21 0.13 0.00 0 17/NIZ/03 H15 H15.8 Inside Structure Surface sediments Ca Cl (ua) Dol Q G 0.13 0.83 0.00 0 17/NIZ/03 H15 H15.9 Inside Structure Surface sediments Ca Cl (ua) Dol Q G 0.20 0.52 0.00 0 17/NIZ/03 H16 H16.1 Inside Structure Surface sediments Ca Cl (ua) Dol Q G 0.25 0.09 0.00 0 17/NIZ/03 J15 J15.1 Inside Structure Surface sediments Ca Cl (ua) Dol Q 0.24 0.05 0.00 0 17/NIZ/09 K16 K16.2 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.10 0.72 0.00 0 17/NIZ/09 K17 K17.3 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.17 0.47 0.00 0 17/NIZ/09 L16 L16.2 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.14 0.44 0.00 0 17/NIZ/09 L17 L17.2 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.29 0.55 0.00 0 17/NIZ/09 M17 M17.2 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.00 0.34 0.00 0 17/NIZ/09 M18 M18.2 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.03 0.16 0.00 0 17/NIZ/09 M19 M19.2 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.79 0.13 0.00 0 17/NIZ/09 N16 N16.2 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.11 0.16 0.00 0 17/NIZ/09 N16 N16.3 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.13 0.59 0.00 0 17/NIZ/09 N17 N17.2 Outside structure Sediment accumulation Ca Cl (ua) Dol Q 0.22 0.42 0.00 0 17/NIZ/09 N17 N17.3 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.17 0.19 0.00 0 17/NIZ/09 N18 N18.2 Outside structure Sediment accumulation Ca Cl (ua) Dol Q 0.11 0.17 0.00 0 17/NIZ/09 N18 N18.3 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.05 0.16 0.00 0 17/NIZ/09 N19 N19.2 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.13 0.00 0.00 - 17/NIZ/09 N19 N19.3 Outside structure Sediment on bedrock Ca Cl (ua) Dol Q 0.08 0.00 0.00 - 17/NIZ/10 K10 K10.3 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.27 14.49 0.05 0.003533569 17/NIZ/10 K15 K15.2 Courtyard Sediment on bedrock Ca (geo) Cl (ua) Dol Q 0.17 26.71 0.18 0.006896552 17/NIZ/10 L10 L10.2 Courtyard Sediment on bedrock Ca (geo) Cl (ua) Dol Q 0.60 6.10 0.00 0 17/NIZ/10 L11 L11.2 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 2.12 41.67 0.04 0.000990099 17/NIZ/10 L15 L15.2 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.13 14.44 0.00 0

223 17/NIZ/10 M09 M09.1 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.35 3.38 0.00 0 17/NIZ/10 M10 M10.1 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.87 20.60 0.00 0 17/NIZ/10 M11 M11.1 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 1.91 25.21 0.00 0 17/NIZ/10 M12 M12.3 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.35 49.85 0.04 0.000729927 17/NIZ/10 M14 M14.2 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.23 29.36 0.15 0.004942339 17/NIZ/10 M15 M15.1 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.19 5.55 0.10 0.017857143 17/NIZ/10 N10 N10.3 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.60 11.91 0.00 0 17/NIZ/10 N11 N11.2 Courtyard Sediment accumulation Ca Cl (ua) Dol Q 0.49 33.61 0.04 0.001240695 17/NIZ/10 N11 N11.3 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.07 21.73 0.00 0 17/NIZ/10 N12 N12.2 Courtyard Sediment accumulation Ca Cl (ua) Dol Q 0.55 15.40 0.00 0 17/NIZ/10 N12 N12.3 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.60 21.22 0.00 0 17/NIZ/10 N14 N14.2 Courtyard Sediment accumulation Ca Cl (ua) Dol Q 0.19 56.71 0.00 0 17/NIZ/10 N14 N14.3 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.06 8.82 0.04 0.004739336 17/NIZ/10 N15 N15.3 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.20 7.75 0.00 0 17/NIZ/10 N16 N16.5 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.19 1.82 0.00 0 17/NIZ/10 N16 N16.6 Courtyard Sediment on bedrock Ca Cl (ua) Dol Q 0.15 3.25 0.00 0 17/NIZ/11 G16 G16.2 Inside Structure Grey feature Ca Cl (ua) Dol Q 0.08 0.19 0.00 0 17/NIZ/11 G16 G16.3 Inside Structure Grey feature Ca Cl (ua) Dol Q 0.04 0.00 0.00 - 17/NIZ/11 G16 G16.4 Inside Structure Sediment on bedrock Ca Cl (ua) Dol Q 0.04 0.40 0.04 0.1 17/NIZ/11 G16 G16.5 Inside Structure Grey feature Ca Cl (ua) Dol Q G Humic acid 0.14 0.21 0.00 0 17/NIZ/11 G16 G16.6 Inside Structure Grey feature Ca Cl (ua) Dol Q G 0.23 0.25 0.00 0 17/NIZ/11 G16 G16.7 Inside Structure Grey feature Ca Cl (ua) Dol Q G 0.39 0.27 0.05 0.2 17/NIZ/11 H15 H15.2 Inside Structure Sediment on bedrock Ca Cl (ua) Dol Q 0.08 0.31 0.00 0 17/NIZ/11 H15 H15.3 Inside Structure Sediment on bedrock Ca Cl (ua) Dol Q 0.21 0.62 0.00 0 17/NIZ/11 H15 H15.4 Inside Structure OSL-1 (sediment on bedrock) Ca Cl (ua) Dol Q 0.35 0.81 0.00 0 17/NIZ/11 H15 H15.5 Inside Structure OSL-2 (sediment on bedrock) Ca Cl (ua) Dol Q 0.13 0.52 0.04 0.083333333 17/NIZ/11 H15 H15.6 Inside Structure OSL-3 (sediment on bedrock) Ca Cl (ua) Dol Q 0.16 0.46 0.00 0 17/NIZ/11 H15 H15.7 Inside Structure OSL-4 (sediment on bedrock) Ca Cl (ua) Dol Q 0.09 0.96 0.00 0 17/NIZ/11 H16 H16.2 Inside Structure Sediment on bedrock Ca (geo) Cl (ua) Dol Q G 0.12 0.44 0.00 0 17/NIZ/11 H16 H16.3 Inside Structure Sediment on bedrock Ca (geo) Cl (ua) Dol Q G 0.10 0.79 0.00 0 17/NIZ/11 J15 J15.2 Inside Structure Sediment on bedrock Ca Cl (ua) Dol Q 0.16 2.05 0.00 0 17/NIZ/11 J15 J15.3 Inside Structure Sediment on bedrock Ca Cl (ua) Dol Q 0.20 2.01 0.00 0 17/NIZ/11 J15 J15.4 Inside Structure Sediment on bedrock Ca Cl (ua) Dol Q G 0.34 1.65 0.00 0 17/NIZ/11 J15 J15.5 Inside Structure Sediment on bedrock Ca Cl (ua) Dol Q 0.18 8.45 0.00 0 17/NIZ/11 J16 J16.2 Inside Structure Sediment on bedrock Ca Cl (ua) Dol Q 0.12 5.48 0.00 0 17/NIZ/11 J16 J16.3 Inside Structure Sediment on bedrock Ca Cl (ua) Dol Q 0.35 1.41 0.00 0 CONTROL NIZ-C1 CONTROLS alluvial fan Ca Cl (ua) Dol 0.11 0.04 0.00 0 CONTROL NIZ-C2 CONTROLS ridge Ca Cl (ua) Dol Q 0.00 0.04 0.00 0 CONTROL NIZ-C3 CONTROLS alluvial terrace Ca Cl (ua) Dol Q 0.04 0.00 0.00 - CONTROL NIZ-C4 CONTROLS hammada Ca Cl (ua) Dol Q 0.01 0.00 0.00 - CONTROL NIZ-C5 CONTROLS alluvial wadi (N. Zin) Ca Cl (ua) Dol Q 0.00 0.00 0.00 - CONTROL NIZ-C6 CONTROLS ridge Ca Cl (ua) Dol Q 0.13 0.00 0.00 - Note: mineralogical data follows publications presented above. Dolomite was only identified by minor peaks at 728 cm-1. 224 צקת י ר

תדובע טרוטקוד ז ו יצמ הג תא תואצותה לש רקחמ ורקמ - ורקימו - יגולואיכרא לש תלכלכ יקה ,םו ורכה נ היגולו היגולו נ ורכה ,םו יקה תלכלכ לש יגולואיכרא

הירוטסיההו יה תיבושי לש רה בגנה ךלהמב תפוקת ורבה נ הז יבה נ י תימי תפוקת( ב"בה , 1950–2500 "הספל נ הפ

)ךרעל . העברא רתא םי רפחנ ו ו נ דמל ו תרגסמב רקחמה : ע י ן ז י ק באשמו י - דש ה , םיינש םירתאהמ םילודגה םייזכרמהו םייזכרמהו םילודגה םירתאהמ םיינש ,

לש רה ,בגנה ומכ םג לחנ רקוב 66 לחנו הנצינ 332 , נש י םירתא םינטק םייסופיטו .הפוקתל והז רקחמה טרופמה טרופמה רקחמה והז .הפוקתל םייסופיטו םינטק םירתא י

ושארה ן ןדה תלכלכב יקה םו ורכהו נ גולו הי לש ת תפוק ב"בה ינשב םירושעה ורחאה נ ,םי ןושארהו וחבל ן וס ג י תו ולאואת ס חל ואה ם רא יוע יש "התו

ךרד יד ג םו רקחמו ורקימ - כרא יגולואי זרב היצולו .ההובג

,תישאר תדובע טרוטקוד ז ו תראתמ הטיש השדח רשא החתופ לע סיסב רקחמ םיטנמידס רתאמ לחנ רקוב 66 הדעונו הדעונו

תוהזל תוריהמב יללג ילעב יח םי םירתאב םיגולואיכרא צמאב תוע Fourier transform infrared spectroscopy

FTIR( .) הטיש השדח וז הבושח רובע תולאש רקחמ תונדה תלכלכב םויק טרפבו יגהונב לוהינ קשמ יח תייערו תייערו יח קשמ לוהינ

רדע ,םי ו נ ןתי שמתשהל הב הדשב תעב פחה י הר .

נש י ת , ה תנ ו נ י ם רקמה ו - ורקימהו - ולואיכרא ג םי םירתאהמ ורפחנש גצומ םי םישמשמו סיסבכ ל יד ו ן רבדב תולאש תולאש רבדב ן ו יד

ושקה ר ו ת טרטסאל ג י ו ת ק י ו ם ו א ז ו ר י עפ תולי .םימודק ודעה י תו ורקימה - כרא י א ו ל ו ג י ו ת ה ן דח - משמ ע י ו ת ו מ צ ב י ע ו ת ע ל ש

ךכ רתאהש םי נטקה םי לש לחנ ב ו רק 66 לחנו הנצינ 332 וקסע רקיעב נב לוהי קשמ יח רשא ללכ יער תי םירדע םירדע תי יער ללכ רשא יח קשמ לוהי נב רקיעב וקסע

בסב י םתב קמה ו מ י ת . אל האצמנ לכ דע ו ת אלקחל ו ת ד ג ן . ילולכמ ילכ סרחה םירתאב וללה םיבכרומ רקיעב יריסמיימריבםברמ להםרא רה ל יוכ

ב לושי ילכו ,השגה ףוסיאו םוגידו לש םיאצממ זרב היצולו ההובג לחנב נ הנצי 332 הארה יכ ימגד תוליעפה וניפו י יויותלעה מד כהר

הפשאה ירתאב םיעור םימודק ולא םימוד ולאל םידעותמה ירתאב םיעור האמהמ .םירשעה

נב י ג דו ,ךכל םירתאה םילודגה לש יע ן קיז יבאשמו הדש אל וריתוה םוש תודע רוצייל ןוזמ — אל לוהינ קשמ יח אלו אלו יח קשמ לוהינ

ילודיג .ןגד םילולכמה םיימרקה םירתאהמ וללה םיבכרומ רקיעב תורוצמ תורוגס לש ילכ ןוסחא אל םיימוקמ םילדגב םילדגב םיימוקמ אל ןוסחא ילכ לש תורוגס תורוצמ רקיעב םיבכרומ וללה םירתאהמ םיימרקה םילולכמה .ןגד ילודיג

וש נ .םי יליטמ ,תשוחנ ילכ תשוחנ תוסיפו תשוחנ םיחיכש נשב י .םירתאה ףא לע יפ ןכ , רקחמ XRF דמ ו קדק אל א ד

ףשח ףא תודע וליעפל י תו תורושקה הקפהל וא ביע דו ,תשוחנ ו תאז דוגינב תועצהל תומדוק .רקחמב לע סיסב םינותנה םינותנה סיסב לע .רקחמב תומדוק תועצהל דוגינב תאז ו ,תשוחנ דו

רקמה ו - ורקימהו - ולואיכרא ג ,םי שי רזחשל תא םירתאה םילודגה זכרמהו י םי תודוקנכ רחס רשא ומקומ זאב םירו םירו זאב ומקומ רשא רחס תודוקנכ םי י זכרמהו םילודגה םירתאה תא רזחשל שי ,םי ג ולואיכרא

ומ נג םי יבאשמ( )הדש וא תוכימסב רוקמל םימ עובק יע( ן ז קי .)

225 סבל ו ף , ומ אב םי דל י ו ן הה י טב םי רכה ו נ ו ל ו ג י םי בקתהש ול ירקחממ ןמחפה 14 הו - optically stimulated

OSL) luminescence ) תוריפחהמ ועצובש תרגסמב תדובע טרוטקודה .וזה תואצותה תועיבצמ לע ךכ םירתאהש ךכ לע תועיבצמ תואצותה .וזה טרוטקודה תדובע תרגסמב ועצובש תוריפחהמ )

נטקה םי םה העפות תכורא - וט חו — לחנ רקוב 66 ללוכ תואצות ןמחפ 14 ופמה תורז ךרואל קרפ ןמזה יבש ן –3300

2300 "הספל נ ךרעל — עב ו ד רתאה י ם םילודגה גכ ו ן יע ן ז קי יקתה ומי קרפב ןמז ומ ,לבג תיצחמב הנושארה לש ,ב"בה בב שהואהתצמ ,ב מ מ קפ ויית יז ןי ןוג יוג

2200–2500 נ"הספל .ךרעל םג םירתאה םילודגה זכרמהו י םי ו םג םירתאה נטקה םי נ ושטנ תוביבסב 2200 "הספל נ . הפ

דע תו ז ו תמאות תא דעה ו י תו תאמ יר יפ ןאנ , ימבו דחו 'חמ תברי הרמח ןדפיא . הארנ תוטטומתהש ךרעמ חנה תשו ב נ בג גנב ש

יפבו ןאנ הרושק רקיעב התליפנל לש הכלממה המודקה .םירצמב נב ,ףסו םתודרשיה לש םירתאה נטקה םי תפוקתמ תפוקתמ םי נטקה םירתאה לש םתודרשיה ,ףסו נב .םירצמב המודקה הכלממה לש התליפנל רקיעב הרושק ןאנ יפבו

ורבה נ הז המודקה )ק"ב( רבעמבו ךותל ב"בה תררועמ שדחמ תא וה ו חוכי יבגל הקולחה לש תופוקת ק"בה III II-ו II

בגנב זאבו םירו ,םיכומס תזמרמו ךכל ךלהמבש ףלאה ישילשה "הספל נ ךרעמ תשוחנה םורדב םייקתה דרפנב דרפנב םייקתה םורדב תשוחנה ךרעמ נ

הליפנהמ תיללכה לש זאה םירו נוריעה י םי ופצב ן , וא היה יסח ן נפב .הי

226

וא נ תטיסרבי לת ביבא וקפה הטל עדמל י - רה ו ח ע ש" רטסל ו לאס י א נ ט י ן תיב רפסה יעדמל תודהיה ולואכראו ג הי ש"ע יח םי גרבנזור וחה ג ולואיכראל ג הי וברתו י תו חרזמה םודקה ש"ע בוקעי .מ קלא ו ב

רה ה נ ג ב ב תפוקת ורבה הזנ תימייניבה (1950-2500 נ"הספל ) : הה י טב גה ואי - כרא י ולוא ג י

רוביח םשל תלבק ראות רוטקוד היפוסוליפל תאמ יראכאז קראלק ד א 'תסנ

הב נ ח תי רפ ו פ ' לארשי פ י נ יטשלק י ן ( א ו נ י סרב י ט ת לת בא י ב ) רפ ו פ ' ר תו קחש - סורג וא( נ טיסרבי ת )הפיח

וה שג טאנסל לש וא נ תטיסרבי לת ביבא 2019