Socio-Economic Aspects of Chalcolithic (4500-3500 BC) Societies in the Southern Levant: A lithic perspective 9781407301860, 9781407332314

This work summarizes a techno-typological analysis of Chalcolithic (c. 4500-3500 B.C.) lithic assemblages from Southern

180 79 77MB

English Pages [213] Year 2008

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Front Cover
Title Page
Copyright
Acknowledgements
Dedication
Table of Contents
CHAPTER I: INTRODUCTION
CHAPTER II: AIMS AND METHODOLOGICAL FRAMEWORK
CHAPTER III: NEGEV SITES
CHAPTER IV: CENTRAL ISRAEL SITES
CHAPTER V: NORTHERN ISRAEL SITES
CHAPTER VI: SITES IN JORDAN
CHAPTER VII: DESERT SITES OF ISRAEL ANDSINAI
CHAPTER VIII: DESCRIPTION OF RESULTS
CHAPTER IX: CONCLUSIONS
REFERENCES
LIST OF SITES AND THEIR COORDINATES
PLATES
TABLES
FIGURES
MAPS
Recommend Papers

Socio-Economic Aspects of Chalcolithic (4500-3500 BC) Societies in the Southern Levant: A lithic perspective
 9781407301860, 9781407332314

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

BAR  S1744  2008   HERMON  

Socio-Economic Aspects of Chalcolithic (4500-3500 BC) Societies in the Southern Levant

SOCIO-ECONOMIC ASPECTS OF CHALCOLITHIC (4500-3500 BC) SOCIETIES

A lithic perspective

Sorin Hermon

BAR International Series 1744 B A R

2008

Socio-Economic Aspects of Chalcolithic (4500-3500 BC) Societies in the Southern Levant A lithic perspective

Sorin Hermon

BAR International Series 1744 2008

Published in 2016 by BAR Publishing, Oxford BAR International Series 1744 Socio-Economic Aspects of Chalcolithic (4500-3500 BC) Societies in the Southern Levant © S Hermon and the Publisher 2008 The author's moral rights under the 1988 UK Copyright, Designs and Patents Act are hereby expressly asserted. All rights reserved. No part of this work may be copied, reproduced, stored, sold, distributed, scanned, saved in any form of digital format or transmitted in any form digitally, without the written permission of the Publisher.

ISBN 9781407301860 paperback ISBN 9781407332314 e-format DOI https://doi.org/10.30861/9781407301860 A catalogue record for this book is available from the British Library BAR Publishing is the trading name of British Archaeological Reports (Oxford) Ltd. British Archaeological Reports was first incorporated in 1974 to publish the BAR Series, International and British. In 1992 Hadrian Books Ltd became part of the BAR group. This volume was originally published by Archaeopress in conjunction with British Archaeological Reports (Oxford) Ltd / Hadrian Books Ltd, the Series principal publisher, in 2008. This present volume is published by BAR Publishing, 2016.

BAR PUBLISHING BAR titles are available from:

E MAIL P HONE F AX

BAR Publishing 122 Banbury Rd, Oxford, OX2 7BP, UK [email protected] +44 (0)1865 310431 +44 (0)1865 316916 www.barpublishing.com

Acknowledgements The work could not have reached this final version without the contribution of the following persons, to whom I express my gratitude: My family: Clara Goldstein (grandmother), Lia Hermon (mother), Eva Hermon (sister) and Daria Ruggeri – I love you all. My Professors: Prof. Isaac Gilead - advisor, Prof. Steven A. Rosen, Prof. Eliezer D. Oren –Ben Gurion University, Prof. Ana Belfer-Cohen, Prof. Nigel A. Goring-Morris, Prof. Na’ama Goren-Inbar –the Hebrew University of Jerusalem and Prof. Ofer Bar-Yosef, Harvard University. My colleagues and friends: Ofer Marder and Hamudi Khalayili, Peter Fabian, Nimrod Getzov, Uzi Avner, Rina Baskirer, Flavia Zontag, Pirchia Nachshoni, Ya’akov Baumgarten, Aaron Sugar, Ran Barkai, Lior Grosmann, Yossi Garfinkel, Vladimir Zbenovich, Sorin D. Iordache and wife and Herbert Permuter and wife, and Emil Alag’em. Particular thanks to Franco Niccolucci, Cristina Pugi, Ginevra and Virginia.

Graphics: Patrice Kaminski, Daniel Ladirai, Tali Fabian, Michael Smilansky and Leonid Zeiger.

My particular gratitude to Jean Perrot, who kindly permitted me to study the lithic material of Safadi and Abu Matar and expressed useful comments and advice during this work.

In memory of my grandmother, Clara Goldstein

Contents

CHAPTER I

INTRODUCTION ..........................................................................................................................................3

DESCRIPTION OF ENVIRONMENTS ............................................................................................................................................3 THE NEOLITHIC BACKGROUND ................................................................................................................................................4 EARLY SYNTHESES OF THE CHALCOLITHIC .............................................................................................................................5 CURRENT EXPLANATIONS ........................................................................................................................................................7 THE CHALCOLITHIC CULTURAL MOSAIC ................................................................................................................................9 CHRONOLOGICAL FRAMEWORK AND PERIODIZATION ..........................................................................................................10 HISTORY OF CHALCOLITHIC FLINT INDUSTRIES RESEARCH...................................................................................................11 CHAPTER II

AIMS AND METHODOLOGICAL FRAMEWORK ........................................................................13

AIMS .......................................................................................................................................................................................13 THE CHAÎNE OPÉRATOIRE CONCEPT AND LITHIC INDUSTRIES ANALYSIS ..............................................................................14 DEFINITIONS AND ATTRIBUTE ANALYSIS OF WASTE PRODUCTS AND TOOL-TYPES ...............................................................16 COMPARATIVE FRAMEWORK..................................................................................................................................................20 CHAPTER III

NEGEV SITES ........................................................................................................................................23

ABU MATAR (128.7/71.5) AND BIR ES-SAFADI (129/71.2)....................................................................................................23 NEVE NOY (CA. 129/71) .........................................................................................................................................................31 TEL-SHEVA (135.3/73) ...........................................................................................................................................................32 NEVATIM (139.6/70.25) .........................................................................................................................................................36 MITHAM C (130.9/72.5) .........................................................................................................................................................37 SHIQMIM (115/67) ..................................................................................................................................................................40 RAMOT NOF (CA. 131.7/76.5) ................................................................................................................................................44 RAMOT 3 (131.65/76) ...........................................................................................................................................................45 TEL ARAD (162/75) ................................................................................................................................................................49 MEITAR (143.83/82.2) ............................................................................................................................................................49 GRAR (143.83/82/2) ...............................................................................................................................................................52 WADI GHAZZEH SITES (CA. 101/75) .......................................................................................................................................55 NORTH-EASTERN SINAI SITES (CA. 53.54/73.64) ...................................................................................................................60 CHAPTER IV

CENTRAL ISRAEL SITES ..................................................................................................................63

GAT-GOVRIN – WADI ZEITA (NAHAL KOMEM) (129/116.8) .................................................................................................63 KHIRBETH ALIYA EAST (CA. 146/123)...................................................................................................................................66 CHAPTER V

NORTHERN ISRAEL SITES ...............................................................................................................68

TEL TEO (CA. 210/285)...........................................................................................................................................................68 BEIT NETOFA VALLEY SITES (CA. 170/210) ...........................................................................................................................68 EINOT KAUKAB (174.45/248.45) ...........................................................................................................................................68 TEL TURMUS (210.68/290.89)................................................................................................................................................73 NEVE UR (CA. 202/232)..........................................................................................................................................................79 GOLAN SITES (CA. 221.55/253.35) .........................................................................................................................................80 THE BURIAL CAVE IN PEQI’IN (181.35/264.40) ......................................................................................................................84 CHAPTER VI

SITES IN JORDAN ................................................................................................................................86

TELEILAT GHASSUL (CA. 208/135) ........................................................................................................................................86 EAST JORDAN VALLEY SITES ..................................................................................................................................................87 TELL SHUNA NORTH (CA. 205/225) .......................................................................................................................................87 ABU HAMID (CA. 210/220) .....................................................................................................................................................87 WZ121: TUBNA – WADI ZIQLAB ...........................................................................................................................................89 PELLA .....................................................................................................................................................................................90 KERAK PLATEAU SITES ..........................................................................................................................................................90 TELL EL-HIBR .........................................................................................................................................................................90

Contents

SOUTHERN JORDAN SITES (CA. 210/1) ...................................................................................................................................90 SOUTHEAST ARABAH SITES (CA. 180/1) ................................................................................................................................92 CHAPTER VII

DESERT SITES OF ISRAEL AND SINAI .........................................................................................93

NAHAL NEKAROT (174.5/105) ...............................................................................................................................................93 EASTERN SINAI SITES (CA. 8.39/79.77) ..................................................................................................................................93 TIMNA 39 (CA. 143/92) ...........................................................................................................................................................96 THE CEMETERY AT EILAT (14.32/88.52) ................................................................................................................................97 CHAPTER VIII

DESCRIPTION OF RESULTS .............................................................................................................98

CHAPTER IX

CONCLUSIONS .................................................................................................................................. 114

CHAPTER I. INTRODUCTION

CHAPTER I

Sinai desert. The region’s geographical setting is “marginal”, situated at the south-eastern corner of the Mediterranean and the north-eastern extremity of the Red Sea. The deserts surrounding it from the southwest, south and east pass into the sub-humid and humid Mediterranean climate. The boundary of aridity oscillates with annual rainfall fluctuations (Karmon 1971:6).

INTRODUCTION

The work presented below summarizes a technotypological analysis of Chalcolithic (ca. 4500-3500 B.C.) lithic assemblages from Southern Levant (sites from Israel, the Golan heights, the Jordan valley, Southern and eastern Jordan and eastern and north-eastern Sinai). This period witnessed major changes in the lifestyles of inhabitants in this region, representing the peak of a long development in the rural life, a process that started with first Neolithic villages and ended up in the Early Bronze Age (henceforth EB) period, with the establishment of first towns. All accessible assemblages dated to the above mentioned period have been studied in the laboratory.

Israel is divided into 4 major regions: continental shelf, coastal plain, mountainous backbone and the JordanArava Rift Valley. Its drainage system has three base levels: the Bay of Eilat, the Mediterranean Sea and the Dead Sea. The catchment area leading to Eilat is restricted to the Southern Arava and part of the Eilat block. No perennial streams occur and rare, heavy rains result in short flooding of wadis. Floods from the Eilat block occasionally reach the sea, while those flowing to Southern Arava disappear under surface gravel, enriching the groundwater or forming seasonal playas. The Northern Negev is divided between the Mediterranean and the Dead Sea, via Nahal Zin to the Dead Sea and Nahal Besor and its catchment to the Mediterranean. Negev wadis do not maintain a perennial flow, but are subjected to seasonal torrential floods. The eastern sector is drained to the Jordan Valley by dry wadis to the south and perennial wadis to the north. The Jordan drainage system has two intermediate lakes, Hula and Tiberias (Horowitz 1979:19).

More than two hundred thousand flint artefacts were included in this work, among them ca. twenty thousand tools, the rest being equally divided between debris and débitage. Approximately a third of this database was directly analyzed by me, the rest was studied from publications. A major problem encountered during the synthesis was the variety of excavation and collection methods and the lack of a standardized research methodology of the Chalcolithic lithic material culture. Researchers apply various research standards, which are not always highlighted or explicitly described. Thus, a large-scale synthesis as such, has to take into consideration these factors and adjust the data according to one standard, which is presented below.

Three main topographic divisions are identified: mountains (the oldest), rift and coastal plain. Most erosion channels (wadis) form swamps in their lower forms, while in their upper stretches kick-points and waterfalls occur, more frequent in the eastern drainage, where wadis form deep canyons. The dominant features of the coastal plain are ridges of calcareous, a red sandy soil and sand dunes (kurkar ridges).

The first part of the work presents the Neolithic background, the state-of-art in Chalcolithic research, with an emphasis on lithic aspects and environmental aspects of the study area. The next chapters are devoted to the presentation of the work’s main goals and the methodological background, followed by a presentation of research standards applied in this study.

The climate is transitional between Mediterranean and desert. Two seasons are expected: warm and dry (summer) and cool and wet (winter), with short periods of transition. During the summer the region receives hot air masses from the west (tropical), slightly cooled and moistened by movement over the Mediterranean Sea. During winters, the area is under the influence of two cold air streams: from the Central Asia High and from the Balkan area. Winter conditions are marked by instability and irregularity, seasonal and from year to year.

The main body of the work is devoted to the presentation of lithic assemblages, in a geographic order, with an emphasis on assemblages analyzed by me; the results of this research are compared, according to variables assumed to have socio-economic and cultural implications. Finally, conclusions are presented, according to the proposed goals of this work. Description of environments Modern Environments

There is a gradual rainfall decrease from north to south. Variations also occur within the region, rainfall may vary from place to place, as different air masses may lie simultaneously over different parts of the country. Thus, the average temperature differs locally, the coastal zone being influenced by the sea, the mountainous area being influenced by elevation, the Jordan valley maintaining a higher temperature, and the Negev being influenced by a continental temperature, with its southern areas showing higher temperatures (Karmon 1971:24) (Map 1a).

The Southern Levant is situated between the Mediterranean Sea, the Red Sea and the Persian Gulf. It consists of lowlands and tablelands, bounded in the north and east by high mountain chains (Taurus and Zagros) and to the south by the Arabian Desert. The southwestern edge is separated from the rest by a deep ditch of the Great African Rift (Aqaba Gulf and Jordan-Arava depression), which has created a smaller geographic unit, along the Mediterranean coast. Its Northern boundary is the Lebanese mountains, while the Southern limit is the

3

CHAPTER I. INTRODUCTION

In light of the above presented research stage, climatic reconstruction should be carefully used for understanding the cultural history of the research area. The abundance of Chalcolithic settlements in the Northern Negev may be a response to an amelioration of climatic conditions. However, their contemporaneity and life span can hardly be evaluated. Hence, the suggestion of a Chalcolithic climate similar to the modern one cannot be neglected.

The Southern Levant includes three main environmental zones: Mediterranean, Irano-Turanian and SudanoSindian, while tropical Sudano-Deccanian appears in some desert oases. A Mediterranean environment surrounds the Mediterranean Sea. Rainfall averages from 400 mm in the south to 1200 mm/year in the Northern mountains. The flora’s main characteristics are evergreen forests and maquis, batha and barigue (Zohary 1959), which covers almost half of Israel: the mountains down to the Be’er-Sheva Valley, the Northern Jordan Valley and the coastal plain down to the Western Negev.

The Neolithic background Around 6000 B.C., a community of villages with an economy based on agriculture developed in the Southern Levant (Moore 1985). Stekelis (1972) defined one of the first cultural entities of this period – the Yarmukian, amply studied (Gopher 1993, Gopher and Gophna 1993:306, Garfinkel 1993) and radiocarbon dated (Gopher and Gophna 1993:305). Its time-span was from 7500 to 7000 B.P. (Garfinkel 1999a). Stratigraphically, it is above the latest pre-pottery Neolithic entities and below Wadi Rabah layers; apparently it co-existed with desert and northern entities (Gopher and Gophna 1993).

The Irano-Turanian is characterized by low rainfall (ca. 350 mm/year) and extreme temperatures: cold winters and hot summers. Plant activity is restricted to spring and early summer. Common vegetation is steppe, accompanied by shrubs and trees, grading to deserts (low precipitation) and forests, mainly in mountains, where rainfall is heavier. Characteristic areas are the Central Jordan Valley, the Judean Desert and parts of the Western Negev, including the Be’er-Sheva basin (Horowitz 1979). The Saharo-Sindian domain is characterized by low precipitation, usually less than 200 mm/year. Summer is long and dry, while the winter is mild. This environment can be detected in almost the entire Negev, the Dead Sea, Southern Jordan Valley and the coastal dunes. Vegetation is usually restricted to the large floodplains of the wadis. Microenvironments develop, mainly due to springs and wadi beds, where water is available throughout the year. The vegetation resembles the African savannah; it is referred to sometimes as Sudano-Deccanian (Horowitz 1979:31) (Map 1b).

Flakes dominate the waste products of the flint industries; blades were obtained by a standardized production from bipolar cores. New types of arrowheads appear, smaller than in previous periods, along with new types of sickleblades, coarsely denticulated by bifacial retouch. The tool inventory is completed by bifacial knives, borers, scrapers and notches (Gopher and Gophna 1993:306). Yarmukian dwellings were either rounded, apsidal or rectangular. Houses had stone foundations, floors were occasionally plastered and pits extensively used. The economy was based on the cultivation of cereals and lentils, and on mainly raising sheep/goats and a few pigs. Sites are distributed over ca. 10,000 km2, across the centre of Israel and Jordan, in various topographic units: coast, mountains, valleys, the Jordan valley and the Jordanian Plateau (Gopher and Gophna 1993:308).

Palaeo-climates Several methods were applied to study past climates in the research area. In addition to palynology (Baruch 1986, Horowitz 1974, 1989, Rosen 1989), and geomorphology (Goldberg 1987, Rosen 1989), other studies are available: diatoms (Ehrlich 1973), isotopic composition of rain, groundwater and sea water (Gat 1981, Isaar and Bruins 1983, Luz 1982), biogenic stone weathering (Danin 1985) and land snail studies (Goodfriend 1988, Goodfriend at al. 1986).

Another Late Neolithic entity is Jericho IX, or Lodian (Gopher 1993), found below layers of the Wadi Rabah culture. Its geographic distribution overlaps the Yarmukian in some areas, but extends further south, along the Mediterranean coast and the Dead Sea, but not in Jordan (Gopher and Gophna 1993:324). A single radiocarbon date places it in the first half of the 7th millennium B.P. (Gopher and Gophna 1993:318), being apparently later than the Yarmukian (Aurenche et al. 1981, Kafafi 1987). Others suggest that Yarmukian and Jericho IX entities are contemporaneous (Stekelis 1972, Moore 1982, Stager 1992, Garfinkel 1993).

The results of these researches and relevant synthesis are presented by Goldberg and Rosen (1987) and Gilead (1988). The main conclusion that can be drawn from these studies is that there is a disagreement among the various researches, some claiming for amelioration of the conditions, the Chalcolithic period being more humid than today (Horowitz 1974, Rosen 1989, Goldberg and Rosen 1987, Copeland and Vita-Finzi 1978, Goodfriend 1988), others claiming that conditions were similar to the present time (Baruch 1986, Horowitz 1979, Kislev 1987, Frumkin et al. 1991) or even drier than now (Geyh 1994, Wreschner 1977, Neev and Hall 1977, Henry 1985).

Jericho IX lithic industries were only partially described: bipolar cores went out of use, there are small leaf-shaped arrowheads and transversal arrowheads (Gopher and Gophna 1993:319) and coarse–denticulated sickle-blades (Garfinkel 1993:123). A Jericho IX assemblage was published by Khalaily (1999), who suggested that

4

CHAPTER I. INTRODUCTION

Yet a different picture was proposed by Garfinkel (1999b), who views the Qatifian as a cultural entity post Wadi Rabah and pre Ghassulian Chalcolithic, including sites labelled as Besorian by Alon and Gilead (1988). According to this reconstruction, the Southern Qatifian cultural entity is contemporaneous with a northern one, comprising sites with a pottery assemblage belonging to the Beth Shean ware group (Garfinkel 1999b).

ceramic and flint evidences, as well as the geographic distribution, enable the definition of Jericho IX as a distinct culture of the 6th millennium B.C., postdating the Yarmukian (Khalaily 1999:47). Another opinion states that all the defining elements of Jericho IX appeared already in the Yarmukian, and can be found in later Wadi Rabah assemblages (Gopher and Gophna 1993), the differences between Yarmukian and Jericho IX reflecting regionality rather than chronology (Garfinkel 1999b).

The lithic industry is flake dominated, with tool-types similar to the Wadi Rabah. Continuity is also seen in the Qatifian subsistence economy (Gopher and Gophna 1993:337). A single radiocarbon date places the Qatifian at the end of the 7th millennium B.P (Gilead 1988).

The major culture that predates the ones studied in this work is Wadi Rabah, first identified in the Tel-Aviv area (Kaplan 1958a, 1958b) and later on in Central and Northern Israel. Stratigraphically, it lays below Ghassulian and above Yarmukian/Jericho IX layers. Its earliest appearance is at the beginning of the 7th millennium B.P., overlapping later manifestations of the Yarmukian, while the latest dates are from the end of the same millennium, possibly overlapping with the Ghassulian (Gopher and Gophna 1993:306).

The picture emerging from the reconstruction of the 7th and the 6th millennia B.P. is of small villages, with an economy based on agriculture, herding of sheep/goat, cattle and pigs, and the use of secondary products, such as wool and milk. Hunting gradually decreased until its almost total disappearance towards the end of this period.

The lithic industry is oriented towards the production of flakes, with on-site manufacture of most tool-types. Arrowheads are absent; sickle-blades are rectangular, with a denticulated working edge, backed and double truncated, being different from Yarmukian ones. Other tool-types are borers, bifacials, retouched flakes/blades, notches and burins (Gopher and Gophna 1993).

The Ghassulian, a major Chalcolithic culture, emerged from a mosaic of cultural entities that flourished in the Southern Levant at the end of the 6th millennium B.P (Bourke 1997, Goren 1990). It reflects the peak of a cultural evolution of small rural communities, which started 2000 years earlier and ended at the end of the 4th millennium B.C., with the rise of EB urban complexes.

Dwellings are rectangular, with stone foundations and mud-bricks walls, with an internal subdivision. Floors are of trodden earth. Pits are common, some lined with plaster and paved. Occasionally, shaped paved areas are found outside structures (Gopher and Gophna 1993:332).

Early syntheses of the Chalcolithic Early identifications of Chalcolithic cultures differentiated between unpainted ‘Neolithic’ pottery, and painted ‘Chalcolithic’ pottery (Lloyd and Safar 1945). A different approach was adopted by Mellaart (1975), who refers to the painted pottery as ‘Neolithic’. Albright was among the first to discuss the Chalcolithic period and its cultural entities, dating them back to the 2nd half of the 4th millennium B.C (Albright 1932, Hanbury-Tenison 1986).

The economy was based on sheep/goat herding and occasional cattle and raising pigs. The absence of arrowheads and the scarcity of game bone suggest that hunting was not important in the Wadi Rabah subsistence economy. Artefacts described as spindle whorls and loom weights indicate processing of sheep/goat fur. Agriculture was practised, as cereal pytolith remains, seeds and sickle-blades indicate. Apparently, olives were used for the production of oil (Gopher and Gophna 1993:332).

An early defined cultural entity of the Southern Levant, initially recognized at the type-site of Teleilat Ghassul, was the Ghassulian. Following excavations at this site, the researchers noted differences between houses, some “…était on briques pour les maisons bourgeoises, en pisé pour le commun du people” (Mallon et al. 1934:32). Houses of upper layers were built with stone foundations, while those of lower ones were entirely from mud-bricks. Some houses were larger than others, being interpreted as storehouses, or shops. Open spaces between houses were assumed to be streets. The Ghassulians were viewed as “…agriculteurs et de menuisiers, de potiers aussi…un peuple laborieux et pacifique…sédentaires…” (Mallon et al. 1934:64); their origin was unknown, not local, the concluding remark being that “…il nous reste beaucoup de problèmes a résoudre pour avoir une idée concrète sur la civilisation du Teleilat Ghassul et sur l’origine de cette civilisation si étrange et si inattendu dans ce petit coin de Palestine” (Mallon et al. 1934:140).

Sites were mainly found between Nahal Soreq in the south and Upper Galilee and Huleh Valley in the north, covering the Jezreel Valley, the coast and the Jordan Valley, the southernmost extension being north of the Hebron mountains and the northernmost appearance being in Lebanon. Apparently, the Wadi Rabah pottery was influenced by the Halafian, especially when compared to sites in Syria (Garfinkel 1999b: 151). A major cultural entity that co-existed with the Wadi Rabah is the Qatifian (Gilead 1990), mainly recognized at sites in the Besor area and north-eastern Sinai (Epstein 1984). Goren (1990) suggested a wider distribution, including Southern Jordan and the Dead Sea basin.

5

CHAPTER I. INTRODUCTION

and cultivated. Society became more complex, with villages similar to those of modern Arabs, seasonal settlements of clans and homesteads of extended farming families. Negev settlements are half-sedentary and seasonal, consisting of no more than ten household clusters each, with animal pens. The only site where “evidence for a persistent, fully sedentary life is justified” is Teleilat Ghassul (Anati 1963:306).

The reports of site excavations, located mainly in the Negev area and culturally related to Teleilat Ghassul, (MacDonald 1932; Perrot 1955), served as a basis for several syntheses of the Chalcolithic in the Southern Levant. The following is a summary of some of these works, reflecting the state-of-art in the 70’s research. Kenyon’s description Kenyon (1960) views the Ghassulians as newcomers from the east or northeast, who settled in sedentary farming communities, such as Teleilat Ghassul, or in camping sites, such as Wadi Ghazzeh, being nomadic groups, possibly practising a seasonal agriculture and never settling long enough to build permanent houses. The copper industry of Abu Matar is evidence of trade and specialization, markers of a complex community, while the presence of olives at Teleilat Ghassul reflects a non self-sufficient economy, based on relations with other regions. The cultural mosaic is composed of a number of groups of diverse origins, living side by side during the first half of the 4th millennium, which, toward its end, simply died out (Kenyon 1960:82).

The period is characterized by a cultural diversity, the more varied cultures of the south influencing Northern ones. Four major provinces, all sharing general traits in their ceramic and flint industries, but differing in settlement patterns, art, economy and social organization are defined: Wadi Shallale, the Be’ersheba plain, the Judean Desert and Moab. Southern Chalcolithic cultures were in contact with Proto-Dynastic Egyptian cultures, as manifested by “pressure-flaked” flint tools, flint raw materials from Fayum and a granite violin-shaped figurine, found near Gilat. A network exchange of goods, or trade, was regularly carried on (Anati 1963:296). The period’s end was assigned to an invasion of ProtoUrban populations from the north. Judean desert sites were inhabited in a later Chalcolithic period, following this invasion. Ghassul was destroyed and abandoned before 3300 B.C.; Southern culture settlements (such as Kh. Bitar) persisted a little longer (Anati 1963:314).

Perrot’s description For Perrot, the Chalcolithic culture, dated by him to the 4th millennium B.C. is a result of newcomers from the Trans-Jordanian Plateau. The Ghassulian, as represented at Teleilat Ghassul, was their formative phase, being an attempt by an originally semi nomad population at a more permanent settlement (Perrot 1962a:158). The economy was based on stockbreeding; cultivation of wheat, barley and lentils, served as a complement to permit an “almost sedentary life”. The society was “á tendance égalitaire et peu hiérarchisée”, as reflected in burials (Perrot and Ladiray 1980:131). The population of villages, such as Safadi, did not exceed 200 inhabitants, in 15-20 dwellings. Each agglomeration shows a certain specialization degree; apparently, the Be’ersheba group formed an independent economic and social unit (Perrot 1962a:158). Its lifespan did not exceed 200-300 years.

deVaux’s description deVaux, basing his conclusion on earlier works of Perrot (1968), describes a new culture, homogeneous in time and place, the Ghassulian, found in three areas: the Dead Sea, the Negev and the Mediterranean coast. The typesite, Ghassul, was a village of farmers, who practised hunting; their economy was not self-sufficient, as evidenced by olives, brought from elsewhere (deVaux 1970:523). Underground structures of the Be’ersheba sites were interpreted as dwellings adapted to the surrounding conditions (protection against heat and dust) and cannot be used as indication for the inhabitants’ origins (deVaux 1970:524). The presence of a cemetery along the coast led deVaux to reconstruct a movement pattern of the Be’ersheba inhabitants, from the Negev to the north, in the dry seasons, with their flocks (deVaux 1970:528).

At the end of the 4th millennium B.C. an abrupt collapse occurred: sites were abandoned and the sedentary life disappeared in the semi-arid zone, probably as a result of a decline in the security situation in Syria and Palestine, which led Northern populations to adopt an urban mode of life, while the Southern people, unable to adapt to the new conditions, returned to a fully pastoral semi-nomadic way of life (Perrot 1962a:160, 1984, 1987).

The Ghassulian culture, established in new, previously uninhabited areas, under which new styles and techniques were introduced, was intrusive in Palestine, possibly from Armenia or Anatolia, as suggested by the presence in this area of brachycephalic people. Other evidences were the adoption of metallurgy and the presence of arsenic in copper artefacts from Nahal Mishmar, indicating its Anatolian origin. The end of the Ghassulian is marked by abandonment of sites, without any signs of destruction – “the Ghassul-Be’er-Sheba culture disappears without any sequel” (deVaux 1970:530).

Anati’s description According to Anati, during the Chalcolithic period (dated by him to 3700-3200 B.C.) a cultural process of invasion of peripheral nomads and their settlement in villages occurred (Anati 1963:285). The Neolithic – Chalcolithic transition was a gradual change in material culture, hamlets evolved into villages and new areas were settled

6

CHAPTER I. INTRODUCTION

Ussishkin’s description

Gilead’s approach

Ussishkin adopts previous conclusions regarding the Chalcolithic (see above) and relates its beginning to the first use of copper, which was also the main factor for the development and the economic prosperity of the Ghassulian, which developed after a gap of several hundred years following the end of the Neolithic (Ussishkin 1970:110). The production of basalt vessels was performed in sites located in Northern Israel, for a probable kind of family-cult. The presence of fenestrated bowls made both from basalt and pottery, raised the possibility of different economic statuses among Chalcolithic inhabitants. (Ussishkin 1970:118).

Gilead, following earlier suggestions proposed by Kaplan (1969) and Moore (1973), and basing his analysis on fieldwork at various Chalcolithic sites (Gilead 1990), views the Northern Negev Chalcolithic as a local continuation of Neolithic cultures, such as the Qatifian (Gilead 1995:478). Chalcolithic inhabitants practiced a sedentary mode of life, as suggested by the pig remains (Gilead 1988:421). Milk and its by-products were another element in the Chalcolithic subsistence economy, as indicated by research on remains of sheep/goats from Shiqmim (Grigson 1987). The horse was apparently domesticated in this period (Grigson 1995:416). Indirect evidence for dairy product use is the appearance of churns, a typical Chalcolithic pottery-type (Gilead 1988:420). An additional dietary supplement was the exploitation of fruit trees (Kislev 1987). A modern parallel resembling Chalcolithic economy is that of recent farmers in the Levant, known as “Fellahin”. Seminomadic pastoralism was probably practised in Southern arid regions, such as Southern Jordan, Sinai and the Arava (Gilead 1988:421).

The life span of the Ghassulian was a few hundreds years. Villages were gradually abandoned in an organized way, following a continuing cultural decline, which ended with the replacement of new EB age cultures (Ussishkin 1970:125). Summary The approaches presented above, which reflect the stateof-art of the seventies, agree on an intrusion of newcomers, from the north or northeast (Kenyon), the Transjordan Plateau (Perrot), Anatolia or Armenia (Anati and deVaux), or from unknown regions (Ussishkin), which served as a trigger for the appearance of the Chalcolithic culture in the region. Another source of migration, responsible for the rise of the Chalcolithic culture in the Southern Levant was from the east, which also gave rise to the Mesopotamian Ubaid (Elliot 1978).

Copper production was performed by part-time specialists, with a social status similar to others, without signs of monopoly or centralization of production. A similar picture emerges from the analysis of flint or ceramic industries (Gilead 1988:423). Most products were locally made. Signs of artefact movements were noted, such as tabular scrapers (Rosen 1987), “cream ware” pottery, phosphorite / basalt vessels (Gilead and Goren 1989), molluscs from the Nile or the Red Sea and ivory at Abu Matar (Perrot 1955). The economic importance of this exchange network was low, suggesting it fulfilled a different, unknown social role (Gilead 1988:426).

The economy was based on herding and agriculture (Ussishkin), organized in clusters of independent units (Perrot), non-self sufficient (Kenyon, deVaux), similar to that of local modern Arabs (Anati). Life-style was sedentary or in camping sites (Kenyon), almost sedentary (Perrot) or half sedentary, in seasonal settlements (Anati).

Social stratification was weak: “…an aggregate society…numerous small semiautonomous units…demographically coincident with the multiplicity of villages…”, a “primitive democracy”, or “an assembly of elders, heads of households that could have been a social regulator of such societies” (Gilead 1988:434).

Social organization, with signs of trade and specialization, was complex (Kenyon, Anati) or with an egalitarian tendency (Perrot). Ussishkin interprets the presence/absence of fenestrated bowls as markers of different economic status.

Joffe’s approach

The end of the Chalcolithic period is uncertain (Kenyon), it disappears as a result of political instability in the north (Perrot), an invasion from the north (Anati), “without sequel” (deVaux), or a gradual decline and abandonment (Ussishkin).

Joffe (1993) proposed a new periodization for the Chalcolithic, dividing it into three: Early (including Wadi Rabah and Qatifian), Developed (occasionally termed Late Chalcolithic) and Terminal (transitional to the EB). The Developed Chalcolithic was characterized by a number of regional cultures or complexes. The Be’ersheba – Ghassul cluster consists of large, planned villages along the banks of major watercourses, with “community structures, including the ‘temple’ at Ghassul, highly organized ceramic production and…specialized pastoralism”, with evidence of specialized workshops for ivory and copper processing (Joffe 1993:32). Another site complex, related to Be’ersheba – Ghassul, is in Judah,

Current explanations The following is a survey of most recent explanations concerning the origins, the socio-economical organization and the collapse of Chalcolithic cultures in the Southern Levant, presented in chronological order, from the 80s to the present period.

7

CHAPTER I. INTRODUCTION

According to Levy, the settlement pattern was organized in a system of two-tier hierarchy, with large, planned village centres (ca. 10 ha in area), such as Shiqmim or Safadi, and adjacent, smaller dependent satellite sites. The developmental settlement sequence at large villages span over a period of ca. 700 years. The Be’ersheba sites underground structures were defending systems, built on a pioneer occupational stage to protect against aggressors. The need to organize defence of key resources promoted the growth of leadership in the region (Levy 1995:240).

especially the cave sites of the Judean Desert. The coastal plain is another complex, consisting mainly of burial caves. A further one was in the Jezreel Valley, similar in its cultural remains to that of Be’ersheba – Ghassul. The Golan sites are yet another one, with “highly standardized broad-room modules…work and storage areas…and large pithoi for grain storage. The Timnian is a different complex, desert oriented (Joffe 1993:33). The material culture is highly regional, reflecting an expression of cultural or community identity. A high level of material culture consistency is recognized across the Southern Levant. The iconography reflects similarities between complexes. There is evidence for “large villages and specialized sites, agro-technology and specialized pastoralism, craft production, local and longdistance trade…The Chalcolithic was highly developed in terms of material culture and ideology…linked in politico-religious ‘superstructure’ of society…” (Joffe 1993:36), the Chalcolithic being the peak of a long local development from the 6th through to the 4th millennia B.C., during favourable environmental conditions.

Agriculture was improved by check dams and floodwater farming, to support a large, sedentary population. Cattle replaced or accompanied hoe cultivation. There are evidences of “second products revolution”: sheep/goats were exploited for milk and wool; pastoralism evolved, and specialist “herders” took village animals on an annual cycle for seasonally available pasture (Levy 1995:232). Craft specialization is exemplified by the discovery of workshops for processing ivory, copper and stone. The presence of gold artefacts and copper ores is evidence of long-distance trade. Copper was probably brought from the Wadi Feinan area on donkeys, smelted and cast “under very tightly controlled circumstances and not shared with neighbouring Northern Negev societies”. The emergence of craft specialization is a response to the promotion of change and stability assurance, secured by risk management, resource competition and gift-giving, in order to maintain an elite group among the Be’ersheba valley inhabitants (Levy 1995:234).

In Joffe’s opinion, the decline and collapse of the Chalcolithic were caused by three factors: a climatic fluctuation during its Terminal phase, which affected the agriculture, an attenuation of the socio-political organization in the Northern Negev, elites being unable to respond to change and the emergence of commercial influence from Egypt, which caused a social fission, breaking the Chalcolithic patterns of trade and authority (Joffe 1993:37).

A completely different picture of the Chalcolithic emerges from the work of Levy (1986, 1995). According to the author, the end of the 5th millennium B.C. witnesses a number of new and vibrant societies; extended farming communities expanded into semi-arid environmental zones. Their growth and complexity can be best explained by focusing on the interrelationship between environment, population increase and changes in technology and production systems (Levy 1995:226).

The collapse of Chalcolithic societies may be related to various factors that worked together. A climatic deterioration towards the end of the Chalcolithic period destroyed the floodwater farming systems, disrupting the highly specialized farming methods that crystallized during this period. This led to the collapse of the sociopolitical system, tightly stretched over Southern Palestine and unable to evolve alternative avenues of social power. Increase of commercial links with Egypt initiated the erosion of the debt-based Chalcolithic social system. Another factor was warfare with expeditionary Egyptian military organizations (Levy 1995:243).

Marked differences from the Neolithic are: demographic growth, establishment of sanctuaries and cemeteries, emergence of craft specialization and division of sites into spatial hierarchies with settlement centres that coordinated social, economic and religious activities. These changes reflect the emergence of chiefdom organizations (Levy 1995:226), operating on the principle of ranking and lineages graded on a scale of prestige. Levy and Alon, basing their research on the burials discovered at Metzad Aluf cemetery and ethnographic parallels with modern Hawaii societies, described Chalcolithic societies as a low-order rank; membership in some components of the social system was ascribed, rather than achieved (Levy and Alon 1982).

Another research was presented by Golden (1998). In his opinion, there are two industries, one devoted to producing elite goods, while the other produced goods for local consumption: “there is evidence that attached and independent industries of the Northern Negev…operated parallel to each other…the elite body…incapable of extending control to every facet of production …or…there was less interest in controlling the more mundane and less politically fertile utilitarian industry…”. Several workshops, randomly distributed, operated from small villages, while others, small and highly specialized, were controlled by elite clients (Golden 1998:407). Copper industry and complex social formations co-evolved with one another.

Levy’s approach

8

CHAPTER I. INTRODUCTION

A major site concentration was discovered along main Northern Negev wadis. Various settlement types were noted: habitation sites - Safadi (Perrot 1984), Shiqmim (Levy 1987), Gilat (Alon and Levy 1989), Grar (Gilead 1995), ephemeral-herding stations (Nahal Sekher, see Gilead and Goren 1986), workshops, (site A in the Besor area, see Roshwalb 1981) and cemeteries (Metzad Aluf, see Levy and Alon 1982) and Kissufim, see Goren and Fabian 2002). Many sites were noted during surveys (Alon and Levy 1980, Levy and Alon 1982, 1987a, Joffe 1993). The Negev sites were grouped in two principal clusters, with distinct cultural material characteristics and chronological differences: the Besor-Grar and the Be’erSheva ones, the first being earlier than the second (Gilead 1995:475, Gilead 2007). They may represent two human groups, sharing a similar cultural background, but with their own traditions and modes of life.

A doctoral thesis, focusing on metallurgical aspects of the Chalcolithic copper industry, in particular at Abu Matar, concluded that “technological aspects…were far more complex than previously believed…long distance trading for specific copper ores…the ability to smelt sulphiderich coppers ores, the production of an arsenical copper alloy by co-smelting…Abu Matar and Bir Safadi contained a major industrial activity centre…regional craft centre…” (Sugar 2000: 258). Summary Two (different) reconstructions of the socio-economic organization emerge: Gilead’s and Levy’s. For Gilead, Chalcolithic people practiced a sedentary mode of life, being part-time specialists, with a weak social stratification, organized in aggregates of semiautonomous units. The Chalcolithic subsistence economy of the Negev is a continuation of previous Neolithic Qatifian: based on agriculture, herding, milk products and semi-nomadic pastoralism (in arid regions). Social organization was not complex, without evidence for ranking or high-social statuses (Gilead 1988:397). A modern analogy are the Fellahin. There is a possible gap between the Chalcolithic and the next period, the EB-I.

Typical habitation sites consist of dwellings with rectangular houses, used for activities of nuclear families, and open courtyards with pits and installations. Walls were built of mud-bricks, occasionally raised on stone foundations (Gilead 1988:416). A unique feature of some Be’er-Sheva sites is underground structures, interpreted as dwellings of an early stage of occupation (Perrot 1984, Levy 1987). This interpretation was questioned by Gilead (1987), who suggested they were used as storage facilities. Recently, their contemporaneity with the above buildings was suggested, based on radiocarbon dates (Gilead 1994) and stratigraphic considerations (Gilead 1988). No planning can be observed; the duration of occupation and contemporaneity of structures cannot be estimated. Villages, such as Bir es-Safadi, Shiqmim or Grar were composed of a limited number of “household clusters” (Winter 1976), typical of agricultural, pre-urban societies (Gilead 1988:418).

Levy describes sedentary, extended farming Chalcolithic communities, with spatially arranged sites in a system of hierarchy, with central, planned villages and their satellites and a society organized by principles of chiefdom, with ranking based on lineages and prestige. The economy was based on agriculture, improved by irrigation systems, on specialized herders, on craft specialization and on long-distance trade. The collapse of the Chalcolithic culture is due to climatic deterioration, which led to agricultural failure, which caused the breakdown of the elite society. Another factor is the increasing Egyptian economic influence, the trade with this region altering the Chalcolithic social system. A third possible factor was warfare with Egypt.

Despite intensive settlement in the Be'er-Sheva valley and surroundings (Gilead 1988, 1995, Levy and Alon 1987a), only a few Ghassul - Be'er-Sheva sites were discovered in other areas: along the North Sinai coast (Oren and Gilead 1981), in the north-western Negev (Gophna 1979), and the Negev Highlands (Cohen 1988). Few sites were reported from east and south Sinai and the Arava valley (Avner et. al 1994), but none exhibit typical Ghassulian - Be’er-Sheva flint assemblages (Gilead et al. 1995, Gilead and Hermon, in press). They were related to the Chalcolithic because of their pottery assemblages or by radiocarbon dating (Avner et. al 1994).

The Chalcolithic Cultural Mosaic Several cultural clusters have been identified (Kerner 1997b). In the Golan Heights, some 20 sites had been discovered and summarized (Epstein 1978, 1998). They are characterized by stone-built structures, arranged in a long, chainlike pattern, interpreted as habitations for extended families. They represent a distinct cultural entity, given its symbolical representation (Epstein 1998) and cultural material. Another cluster is around Tiberias Lake, where more than 30 sites were recorded (Tsori 1962, Zori 1958) and some excavated (Perrot et al. 1967, Amiran 1977, Epstein 1978). A third cluster, of caves for secondary burial (Perrot and Ladiray 1980), was discovered along the coast plain. Few habitation sites were found in this region (Gilead 1988:412), without a clear link between them and the burial caves.

Two main flint industries had been described from protohistoric Sinai Peninsula: the Eilatian and the Timnian. The Eilatian’s geographic distribution is limited to the coast of the Red Sea, over the west coast of Sinai into the south and Southern Aravah (Ronen 1970, Kozloff 1974, Rothenberg and Glass 1992). The Timnian covers regions in the Timna valley (Rothenberg 1972, 1978, 1988, 1990), Southern Jordan (but see Genz 1997) and Sinai (Kozloff 1974, Henry 1995).

9

CHAPTER I. INTRODUCTION

The presented above dates and their interpretations enable a reconstruction of the occupation history of the Chalcolithic Northern Negev: an early cultural entity, the Besorian of the Besor basin, north-eastern Sinai and Be’er-Sheva valley (Ramot-Nof) existed around 4500 B.C., followed by the Be’er-Sheva - Ghassulian entity, first established at Shiqmim around 4300 B.C. and a couple of a hundred years later at Safadi and its surroundings. Latest Chalcolithic occupation occurred towards the end of the first half of the 4th millennium B.C. at the Cave of Treasures. Thus, the Northern Negev Chalcolithic period did not exceed 500 years of existence.

Kozloff (1974) suggested a contemporaneous date to the Chalcolithic Be’er-Sheva for the Timnian. A rough dating to the 4th millennium BC was proposed for the Eilatian by Ronen (1970), both industries existing in parallel and covering similar areas. An alternative periodization locates the Eilatian “...from the end of the Neolithic to the Late Chalcolithic, whilst the Timnian...extended into the EB...including EB IV...” (Rothenberg and Glass 1992:145, but see Eddy et al. 1999). Some singular sites merit a separate attention. Near the Dead Sea, in the Southern Jordan Valley, a cemetery was found at Adeymeh (Stekelis 1935), possibly linked to Teleilat Ghassul (Mallon et al. 1934). Two sites, near the south-western coast of the Dead Sea, are the Cave of Treasures, where a cache of hundreds of copper items was discovered (Bar-Adon 1980) and En-Gedi, described as a cultic place (Ussishkin 1980). Another site is the burial cave Peqi’in, in Northern Israel (Gal, et al. 1997).

Several C14 dates of the Golan Heights Chalcolithic were published (Epstein 1998). Rasm Harbush was occupied between 4000 and 3600 B.C, representing a later occupation than Northern Negev sites. However, other sites were occupied earlier, such as the “Silo Site”, abandoned sometime between 4496 and 4254 B.C., and site 21, inhabited between 4455 and 4355 B.C. (Carmi et al. 1995:208). Thus, the Golan culture was, at least partially, parallel to the Be’er-Sheva Ghassulian.

Chronological Framework and Periodization The chronological framework was summarized by Levy (1992), Gilead (1994) and Joffe and Dessel (1995). Some dates are available from the Golan sites (Epstein 1998).

Avner (et al. 1994) places the Timnian and the Eilatian of the Southern Negev, the Arava valley and Sinai to the 5th - 4th millennia B.C. Timnian sites of Southern Jordan were C14 being dated to 4000 – 5700 B.P (Henry 1995), and in Eastern Sinai to 5700-3200 B.P., while Eilatian sites were dated to 6575 – 6160 B.P. (Eddy et al. 1999).

Absolute dating A list of C14 dates was discussed by Gilead (1994); three dates are from Horvat Beter (Rosen and Eldar 1993). Their mean is 3977 B.C. (sigma range: 4035-3954 B.C). Seven dates from Bir es-Safadi can be clustered to a mean of 3981 B.C. (sigma range: 4036-3961 B.C.). They support the suggestion that underground structures do not represent a distinct, early occupation stage at Be’er-Sheva sites (Gilead 1987, 1994:4), as proposed by Perrot (1984) and Levy (Levy et al. 1991, Burton and Levy 2001).

A different interpretation of C14 dates was proposed by Joffe and Dessel (1995:511), who suggest a tripartite division of the period: Early (end of the 6th millennium B.C.), Developed (4500-3700 B.C.) and Terminal (37003500 B.C.). The first, transitional from the Neolithic, appears at Tell esh-Shuna North, Tell Abu Habil and Tell es-Sa’idiyeh el-Tahta (none C14 dated); 5 dates from Ghassul span between 5440-4950 B.C., but their context is unclear, “Qatifian”, or Ghassulian (Goren 1990, but see Joffe and Dessel 1995:511). Additional dates (5330-5020 B.C.) are from Tell Wadi Feidan, which has “Qatifian tradition ceramic material”, from the Qatifian site Y-3 and from a probe at Shiqmim.

The largest corpus of C14 dates is of 28 samples from Shiqmim (Levy 1987). Apparently, the site was inhabited from 4357 to 4015 B.C., earlier than Safadi and Horvat Beter (Gilead 1994), for ca. 200 years. This conclusion contradicts an early interpretation of C14 dates by Levy (1992), who suggested a three-phase occupation history, with underground structures used as dwellings at an early stage of occupation. This reconstruction ignores some published C14 dates and their statistical implications.

The Developed Chalcolithic is the longest in the sequence (4500 – 3700 B.C.). Several clusters of C14 dates, corresponding to different cultural development stages (but see Gilead 1988, 1994) were defined (Joffe and Dessel 1995). The first one spans from 4400 to 4100 B.C., consisting of samples collected from underground and above-ground structures at Shiqmim, the Golan Silo site and Nahal Kana cave. The second cluster (4100 to 3900 B.C.) is represented by dates from the same sites. It indicates the final stage of the northward expansion of the Ghassulian tradition and the supplanting of the local Wadi Rabah influenced tradition; habitation continued in the Golan sites and cultic-use of Nahal Kana cave (Joffe and Dessel 1995:511). The third cluster is (3900-3700 B.C.), reflects the final occupation stage at Shiqmim.

A radiocarbon date was obtained from Ramot Nof, located close to the Be’er-Sheva sites. It has a pottery assemblage resembling sites in the Besor area (MacDonald 1932) and was dated to 4681-4464 B.C. (Nahshoni et al. 2002). It indicates an earlier cultural entity in this area, prior to the Be’er-Sheva sites, related to the Besor area area (Gilead 1994). Another cluster of C14 dates comes from the Cave of Treasures (Bar-Adon 1980), which anchors the site around 3700-3650 B.C., later than the Be’er-Sheva sites (Gilead 1994:10).

10

CHAPTER I. INTRODUCTION

The Terminal phase lasts ca. 200 years, is C14 dated by samples from Nahal Mishmar cave and evidenced at sites such as Tell Halif/Lahav terrace, Lachish 1500 and Gat Govrin. Cultures declined and eventually collapsed. The employed term stresses “the fragmented and decayed character of settlement and material culture after the collapse of the main Chalcolithic regional centres…” (Joffe and Dessel 1995:514).

History of Chalcolithic flint industries research Chalcolithic flint assemblages were a minor subject of study during the first half of the XX century. For example, the one of Tuleilat Ghassul was only partially published, and by various authors (Neuville 1934, Mahan 1940, North 1961, Henessy 1969). Other publications, focusing on tool description, were of sites from the Judean Desert (Neuville and Mallon 1931), or surface collections from sites in Northern Israel and Transjordan (Nasralleh 1948, Stekelis 1967, Perrot et al. 1967).

Finkelstein (1996) proposes to term the Chalcolithic as Proto-Urban I (Finkelstein 1996:112), being part of a larger, proto-urban period, which includes the EB.

Flint assemblages from the Besor area (Wadi Ghazzeh) were published a long time ago (MacDonald 1932). Perrot presented a brief synthesis of Chalcolithic flint assemblages, based mainly on collections from the 30’s (Perrot 1952). During the fifties, the flint assemblage of the site Horvat Beter was published (Yeivin 1959).

Garfinkel (1999b) proposed to divide this period into 3: 1. Early, including “Neolithic entities”, such as Jericho VIII and Wadi Rabah, 2. Middle, including sites such as Tell Beith She’an layer XVIII and Qatif, and 3. Late, as Ghassul, Be’er-Sheva sites and Golan sites (Garfinkel 1999b: 104).

In the last 30 years, several publications described Chalcolithic flint assemblages. Lee published a typological study of the Teleilat Ghassul assemblage (Lee 1973). However, important aspects, such as quantitative or contextual, were neglected, thus limiting its use for a synthesis. The flint assemblage of Tel Arad was published by Shick (1978), but its small size reduced its potential for defining characteristic attributes of Chalcolithic flint assemblages. The flint assemblages from Macdonald’s excavations in the Besor area were reanalysed in a doctoral thesis, which emphasized technological aspects, but still concentrated on typological issues (Roshwalb 1981). The analysis of Sinai Peninsula flint assemblages (Oren and Gilead 1981) yielded, among others, the definition of a new tool-type, the micrograttoir (micro end-scraper) (Gilead 1984). The flint assemblages of 'En Esur, a site on the coastal plain of Israel, was recently published (Milevski et al. 2006).

This periodization is mainly based on pottery typology grounds. Terms such as Besorian are not nominated, while others, such as Late Chalcolithic, which refer to late 4th millennium B.C. assemblages (EB) in some publications, are used here for the definition of typically Chalcolithic sites (such as the eponym site Teleilat Ghassul and Be’er-Sheva sites). In order to avoid complications and mis-interpretations, ambiguous terms such as Early Chalcolithic or ProtoUrban were abandoned in this work, the premise being that if the lithic typo-technological research supports chronological differentiations, they will be adopted and labelled consequently. The general term “Chalcolithic” was adopted following the proposal to regard Chalcolithic cultures as of the “fourth millennia B.C.” (Gilead 1988:399), expanding from the mid 5th to the mid 4th millennia B.C. Assemblages belonging to this timespan were analysed according to similar criteria and differences were accordingly interpreted in temporal, socio-economic, cultural or ideological terms.

Some flint assemblages from the Be’er-Sheva area were published, the most recent one is from the site of Gilat (Rowan 2006). The large assemblage of Shiqmim was summarized (Levy and Rosen 1987, Rowan 1990), but a final report is not yet available. Moreover, some technological aspects were neglected, while the application of certain typological criteria limited their integration into a synthesis. The flint assemblage of part of the Safadi site, excavated during the 80’s (Eldar and Baumgarten 1985, Rosen and Eldar 1993), was subject of an inter-site variability research (Hershman 1987). A detailed techno-typological description of the lithic assemblage from Grar, followed by a synthesis of other Chalcolithic assemblages, was published by Gilead et al. (1995).

Relative dating So far, no site has been reported to allow a secure relative dating of the Chalcolithic (Gilead 1995:476). Even though Chalcolithic remains were found on many tells, their context is unclear. Only at a few sites, such as Beth-Shean (Fitzgerald 1934), Megiddo (Shipton 1939), Jericho (Garstang 1935), Tell Farah North (deVaux 1961), Arad (Shick 1978), Shuna north (deContenson 1960), Chalcolithic layers were explored. Apparently, (sporadic) Chalcolithic remains were found below EB or above Late Neolithic layers. However, they cannot clarify the chronological position of the Chalcolithic period, and thus absolute datings are necessary to anchor it in time.

The flint assemblages of some twenty Chalcolithic sites from the Golan Heights were published by Noy (1998), focusing on morphological descriptions of tool-types, but rather neglecting quantitative and technological aspects of the flint industry.

11

CHAPTER I. INTRODUCTION

Two flint industries from sites in the Arava, Southern Jordan and Sinai, Eilatian and Timnian (Ronen 1970, Kozloff 1974, Henry 1995) were dated to the Chalcolithic given characteristics of their pottery assemblages or by C14 dates (Avner et. al 1994).

A doctoral thesis debating cultural aspects of bifacial tools was recently completed, its main hypothesis being that bifacials express existence modes and world perceptions of their producers, as functional tools of everyday life, but also as tools used in the social and ideological struggle (Barkai 2000:315).

From the above presented studies, a general picture can be drawn: several groups of flint industries can be discerned, mainly defined by typological and quantitative characteristics: a Southern group of the Arava, Sinai and Southern Jordan, a Western Negev and possibly Northern Sinai group (Gilead 1995:473), the Be’er-Sheva valley sites, and a Northern group consisting of sites in the Golan Heights and surroundings.

A recent research (Gilead and Hermon in press), focused on detailed techno-typological aspects of the flint assemblages from Bir es-Safadi and Abu Matar. Another reference worth noting is the detailed handbook of protohistoric and historic flint assemblages from Southern Levant, published by Rosen (1997).

12

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

The second goal is to determine the lithic industry level of specialization, as reflected in methods of lithic production and recognition of workshops for the production of specific products. These aspects will be used for discussing social and economical aspects related to the appearance of “specialists”. Respectively, the recognition of “domestic” modes of production of flint tools may reveal aspects of an economy based on the household economy, where each household is an independent economic unit.

CHAPTER II AIMS AND METHODOLOGICAL FRAMEWORK Aims Evaluation of the lithic analysis’ potential for the reconstruction of Chalcolithic societies The tight relations between culture and techniques were emphasized by French structuralists, such as LeviStrausss (1976) or Mauss (1941). Following this theoretical framework, Lemonnier described an “anthropology of technical systems”, arguing that the study of relations between material culture and society becomes the study of coexistence conditions of reciprocal transformation of a technical system and of the society’s socio-economic organization in which it operates (Lemonnier 1986). These relations are emphasized in two main research domains:

Economic organization may be reflected in differences in the tool assemblages from the different sites: differences in the relative quantity of selected tool-types may reflect different activities, or differences in the intensity of activities performed at sites. Thus, aspects of economic organization may be reconstructed. Another possibility is the recognition of “specialized” sites, where a limited number of tool-types predominate.

1. Examining the social control of the technical process or strategic moments is a fertile mean to bridge technical and other social phenomena. 2. The observation of technical variants designates social realities. Explaining them is to try to explore their socio-cultural context, which leads to revealing pertinent links between a technical phenomenon and factors of social order. Irregularity observed in technical behaviour may point toward socio-cultural differences that have hitherto escaped observation.

Another aspect of flint production that may reflect socioeconomic characteristics is its archaeological context – the discovery of caches or the recovery of flint artefacts from non-domestic contexts, such as burials or “spiritual places”. Surplus, represented by possible caches of flint artefacts (end products as well as particular blanks) is characteristic of complex societies (Renfrew and Bahn 1991:156). The recovery of flint artefacts from burial contexts may represent another role, symbolic, which these artefacts played within the society. Moreover, these artefacts, when recognized as grave-goods, may reveal disparities in social status. Flint artefacts, when used as grave-goods, may also reveal gender discrepancies, which can reflect different social status of females/males within the society.

By belonging to a cultural system, technical knowledge is a bridge between techniques and society. If societies exercise “choice” in a universe of possible techniques, this leaves traces in the systems of representation, and the technical solutions must be in harmony with the later.

Reconstruction of the Chalcolithic cultural mosaic in the Southern Levant

The first goal of this work is to explore the potential of techno-typological analyses of flint assemblages for the reconstruction of various aspects of Chalcolithic societies in the Southern Levant, mainly socio-cultural and economic. This capability will be verified from the perspective of logical positivism and analysed within the methodological framework of empiricism, as formulated by Popper (1975) and based on the chaîne opératoire concept (Pelegrin et al. 1988).

The possibility that patterns of flint tool production reflect various cultural entities will be examined. Similarities and differences between patterns will be used for the reconstruction of the cultural milieu in Chalcolithic Southern Levant. Possible inter-connections between the various cultures will be proposed. Successively, each cultural entity will be anchored in time, and its origin from Late Neolithic and its development to EB will be explored. Each entity will be described in its static term and the observed changes will be evaluated whether they represent different entities or variations among a single one.

Description of Chalcolithic socio-economic aspects reflected in lithic assemblages The purpose is to describe aspects of social and economic organization of Chalcolithic societies within the study area, as reflected in the flint assemblages collected from various sites. These aspects may be revealed by various properties of flint assemblages: their process of manufacture, their relative quantity within the different sites and by their archaeological context (Gero 1989).

It should be remembered that gaps exist between anthropological definition of culture (a set of ideas, or symbols acquired by humans as members of a society, Trigger 1968:3) and the definition of prehistorians, who study “material culture” (i.e. lithic artefacts). Thus, from an anthropological point of view, prehistorians do not

13

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

have cultural data to work on. Therefore, any reconstruction of prehistoric cultures will be influenced by these limitations, largely depending on the knowledge about relationships that exist (but which their nature cannot always be correctly evaluated) between material objects and human behaviour in the past.

Chaîne opératoire and lithic industries analysis “…Even the simplest techniques of any…society take the character of a system that can be analysed, in terms of a more general system. The techniques can be seen as a group of significant choices that each society has been forced to make …” (Lévi-Strauss 1976:11).

Keeping this in mind, a suited definition of culture would be: “an archaeological unit possessing traits sufficiently characteristic to distinguish it from all other units…spatially delimited…and chronologically delimited” (Willey and Phillips 1958:22, see also Clarke 1978:490).

Mauss called technique “any effective traditional act” (Mauss 1941). It involves manipulation of materials, sequences of action, tools and particular knowledge (know-how, manual skills, procedures, etc.), forming a system - each technique is a locus of multiple interactions and constant adjustment among its elements. Technique is constantly adapted to material’s transformations, tool characteristics, evolution of know-how; technical knowledge takes into account the tool, the effective action, the material worked, etc (Lemonnier 1986).

The definition and description of prehistoric cultures will be accomplished by applying the chaîne opératoire concept, as a mean for clustering assemblages within a cultural entity. While chaîne opératoire can be considered a common denominator, the rest of the reconstruction is built upon typological elements, including qualitative traits, identified by various attributes (Bar-Yosef 1994:6) and quantitative methods.

Already in the 60’s Leroi-Gourhan highlighted the weakness of the typologically oriented research: “…la typologie ne tient qu’imparfaitement compte du déterminisme imposé par la matière, le geste technique, le degré d’exhaustion des outils, ce qui introduit un élément indépendant du temps et des cultures” (LeroiGourhan et al. 1966). He proposed a research framework, the chaîne opératoire, aimed at producing a “biology of techniques” – a structural and functional awareness of techniques, concluding that “'techniques are both gestures and tools, organized in a true syntax” (Leroi-Gourhan 1964:164). The fabrication act is a dialogue between the knapper and the worked material (Leroi-Gourhan 1965:132) implicating the natural determination of the tendance and the cultural idiosyncrasy of the fait (LeroiGourhan 1943:23-43).

Definition of cultural flint hallmarks (fossiles directeurs) A basic goal of prehistorians is the delineation of cultural configurations, or patterns, from the recovered assemblages (Trigger 1968:15). In order to achieve it, artefacts are analysed by defining a system of ordered units (such as typological groups), used for a comparison between assemblages. The primary comparative unit is the component, defined as a single occupation in the history of a given site (Willey and Phillips 1958:21). For a component to be meaningful, the time segment that is isolated must delimit a static cultural situation, without any significant change. Thus, the cultural material associated with this component should represent the people who inhabited the site, at a given point in their cultural development.

The chaîne opératoire concept assumes that material culture has a relevant history (from natural raw material to cultural matter). Mauss emphasized this point and its implications: “Man creates and at the same time he creates himself” (Mauss 1927: 120). Moreover, Mauss argued (1936) that a technical act is a conscious one, rising from individually and collectively constituted “practical reason”. As a methodological consequence, the need was to pay close attention to all states and events of technical actions and to comprehend their enchaînement organique and their moments essentiels as elements and reflections of this totality (Schlanger 1994:144).

The aim is to define characteristics of lithic production for each cultural entity, in order to set up a database of unique traits to be used as cultural hallmarks, to be described by exploring their unique identity in terms of technology, typology or style. The term “fossile directeur” is thus enlarged, in order to include other dimensions of flint assemblages (see Goring-Morris and Belfer-Cohen 1998:75). Each set of characteristics will be confronted with stratigraphic and chronologic evidences in order to establish its validity. This concept was formulated by Gallus (1977), who described culture as “…an assemblage of objects and observable manifestations or traits…which according to form and expression (observable externalisation) can be regarded as different (disparate) from any other such assemblages of traits…the objects and traits, which document a culture, are the synchronic end results…of a typological series in a particular time period…” (Gallus 1977:143).

The beginning of the 80’s marked a major development in the lithic analyses, highlighting technological aspects, i.e. the preliminary stages before typological studies (Tixier et al. 1980). Research was based on experimentation of flint tool production and analysis of technological attributes. A research scheme was elaborated, its methodological framework being established by the concept of chaîne opératoire (Pelegrin et al. 1988). It included all stages of production, from the acquisition moment of raw materials to the final one of discarding waste products and tools (Garanjer 1992).

14

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

Chaîne opératoire is part of the anthropology of technology; the assumption is that technology embraces all aspects of the process of action upon material, being a "social production" (Lemonnier 1990). Technique is “a socialized action on matter … explained by: suites of gestures and operations - technical process, objects means of action and specific knowledge - connaissance (Schlanger 1994:145). Thus, a research goal would be to describe in which way a given technology is a social production.

Description of research levels of the chaîne opératoire The definition of a chaîne opératoire aiming to describe phases in lithic tool production was recently elaborated (e.g., Boëda et al. 1990, Pigeot 1990, 1991). It includes all stages, from raw material procurement, shaping the core, production of blanks and modification of selected ones, to their use, re-use, and discard. The act of knapping involves intentions, concepts and evaluation of constraints, preferences within a group of equivalent methods as well as technical decisions. This order can be termed a “conceptual operative schema” (Karlin and Julien 1994:154). Thus, a technical scheme can be reconstructed; variations in the technical production depend on differences between levels of technical skill. Its descriptive criteria are:

The reasons for adopting one knapping technique rather than another are often related to constraints imposed by raw material availability (procurement energy expenditures) and characteristics (its mechanical and physical properties). The knapper is also limited by the knowledge (savoir fairs) of knapping methods. The possibility that these constraints emerged not only from the nature of the environment, functional needs, and knowledge, but also from the social system within which chaîne opératoire was practised, cannot be excluded. This approach is rooted in the “conception of techniques as social products: the “objectification” of what are socially elaborate thoughts” (Lemonnier 1990). The reconstruction of chaîne opératoire allows arranging information in a coherent order and to rediscover processes involved in production techniques and conceptual patterns from which they sprang (Pelegrin et al. 1988). It also allows the reconstruction of “production systems” (Karlin and Julien 1994), a knapping process involving intellectual operations: abstraction, anticipation, working out of solutions to problems encountered and the construction of models.

1. Complexity of conceptual scheme –mental construction that guides execution. 2. Degree of chaîne opératoire preconception – precision of forward planning. 3. Discrepancy between a project and its achievement – indicative of ability to take into consideration the data concerning problems to be solved. 4. Core’s final stage – several causes for its shape. 5. Quality of production – the result of the compromise between technical skills and the planned work. 6. Utilitarian productivity – the capability to adjust production according to need Ploux (cited in Karlin and Julien 1994:158) argues that the technical behaviour of an individual is characterized by the phenomena of stability and psycho-motor originality. Each aspect of the technological analysis can be comprehended in terms of psycho-motor operations: reactions, reflections, decisions, execution. A parallel is drawn between three orders of facts and three levels of analysis:

Pelegrin (1985, 1991) distinguishes a scale of “conceptual knowledge” (connaissance) acquired through the memorization of concepts – “the mental representations of forms considered ideal and the materials involved – and memorization of operation modes and procedural knowledge (savoir faire), that can be either ideational, arising from intelligence and memory, or physical movement, presupposing bodily skilled. A distinction can be made between conceptual, abstract knowledge (connaissances), and practical or procedural know-how (savoir-faire). Both are present in an “elaborate flint knapping activity” (Pelegrin 1990).

• Movement and the production of knapping • Succession of movements and chaîne opératoire • Strategy behind their succession and conceptual pattern (Karlin et al. 1992). A second correlation is between these levels and aspects of know-how:

Thus, research is developing along two paths: • Technical production (techno-psychological axis – Boëda et al. 1990) – it implies the knowledge and know-how at the level of concepts, as of methods and techniques (technical behaviour). • Techno-sociological axis - the cultural, spatial and economic implications on which this production depends.

• Elaboration of an “operating conceptual pattern” an intellectual activity, depending on a conceptualised ideational know-how, which consists of making a selection among assimilated knowledge with a view to implementation. • Realization of a chaîne opératoire, or psychomotor activity, depends on an operative ideational know-how that consists of making selected knowledge match a given reality as well as a succession of technical facts. Decision is a

15

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

Stage 1 – the decortication phase can be reconstructed if tool production was performed on-site (their boundaries depending on surveys and on the size of the excavated area). As presented below, in most cases enough information was collected in order to allow a description of this phase.

crucial factor that may show preferences among the choices made. Gesture comes to a motor faculty know-how, including automatic actions that denote habits and motor idiosyncrasy (Karlin and Julien 1994). The principal stages of the chaîne opératoire concept for the analysis of flint tool production are presented below (Mellars 1996:58): Stage 0 Acquisition

Stage 2 – this phase is of major importance for the reconstruction of technological aspects of the lithic industry, which are apparently more sensitive to sociocultural factors on one hand, and to technical skills on the other hand. Since a large amount of débitage was available for a detailed attribute analysis from various sites, aspects of this stage could be reconstructed and their implications discussed.

Extraction and testing of nodule Decortication of nodule

Stage 1 Production

Initial shaping of nodule Preparation of striking platform

Stage 3 – the process of retouching can be reconstructed by observations on tools. Since during early research stages the collecting method focused on assembling an as large as possible amount of tools, and during the last two decades a total collection of artefacts was common, a large amount of tools was available for study.

Stage 2 Production of primary blanks Stage 3 Shaping/retouching of tools Stage 4 Utilization

Use of artefacts Resharpening/reworking of tools

Stage 4 – the use of tools can be inferred by their archaeological context and by use-wear analysis. Since most flint items come from pits, and no detailed use-wear analysis was performed so far, only partial information can be recovered at this stage.

Breakage Stage 5 Discard

Terminal edge-wear / damage

Stage 5 – this analysis is limited to interpretation of flint artefacts’ location within sites. Edge-wear patterns were not studied and they are beyond the scope of this work.

Discard

lithic

Definitions and attribute analysis of waste products and tool-types

Most of the past analyse of Chalcolithic flint industries focused on tools description (e.g. Perrot, 1967, Lee 1973, Roshwalb 1980) or followed a specific oriented research (Hershman 1987, Rowan 1992). Recently, technological descriptions of flint assemblages were published (Gilead 1995, Gilead and Hermon in press, Noy 1998). Research based on refitting (Cziesla 1990, Villa 1991) or experimentation (Inizan et al. 1992) has never been performed on Chalcolithic assemblages, while use-wear analysis was applied on a limited scale (cf. Rosen 1997).

The lithic analysis follows definitions used in the local Chalcolithic research (Roshwalb 1981, Gilead et al. 1995, Rosen 1997) and the Levantine prehistoric research in general (Marks 1976, Bar-Yosef 1970, Goring-Morris 1987). Consequently, assemblages were subdivided into two categories, tools and waste products. Artefacts with intentional retouch (a series of removals) were classified as tools. Exceptions are blanks with gloss along their edges - they were added to the tools. The remaining artefacts were defined as waste products.

Given the limitations presented above, what information can be obtained from Chalcolithic lithic assemblages, following the framework of the chaîne opératoire concept?

The waste products were divided into debris and débitage. All shapeless fragments, for which their method of knapping cannot be identified, nor be assigned to any other category, were classified as debris (Inizan et al. 1992), further subdivided into “chips” (fragments smaller than 20 mm.) and “chunks”, items larger than 20 mm. Flakes, primary elements, blades, bladelets, cores, core trimming elements and burin and bifacial spalls form the second waste class – the débitage. Flakes with more than 50% cortex (natural crust) covering their dorsal face were classified as primary elements. Blades are typologically defined as special flakes, with length at least twice their width (Tixier 1963).

Limitations assemblages

concerning

proto-historical

Stage 0 – in most cases, the raw materials used by Chalcolithic knappers can be identified, therefore their location of acquisition can be determined, although no systematic survey of flint sources used during the 5th – 4th millennium BC is available and no Chalcolithic flint quarry sites have been found so far. Nevertheless, as it will be shown, flint was mostly collected from the immediate vicinity to habitation sites.

16

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

Description of attribute analysis of waste products

Typological considerations

Blank Flakes and Blades

Since the beginning of the century, and under the influence of François Bordes, typological research was regarded as the main tool in establishing chronological frameworks and comparisons between sites: “la typologie est la science qui permet reconnaître, de définir, et de classer les différentes variétés d’outils se rencontrant dans les gisements” (Bordes 1961). Selected flint tool types were regarded as fossile directeurs, useful for dating archaeological horizons, in a similar manner to fossils, used for dating geological layers. Utmost effort was directed to the definition of chronological sequences, based on the “fossile directeur” concept, supposedly reflecting cultural affinities, and for standardization of the terminological vocabulary (Brezillon 1968). Symposiums were held, for the definition of “typological lists”, suited for regions, in a specific time period (Hours 1974).

Samples of flakes and blades were randomly chosen and measured. Length is the maximum extension along the percussion axis, from the bulb of percussion to the distal end. Width is the maximum extension along the edges of a perpendicular line to the length axis. Thickness is the maximum height, perpendicular to the plan of the ventral face of the item. These measurements help to identify knapping techniques, evaluate knapping technique’s “efficiency”, i.e. how many blanks, to be used as tools, were produced from each core and estimate the knapping scheme. Moreover, it is an useful comparison index. Other observations were: 1. raw material type (differentiated by flint texture) allows the identification of raw material procurement strategy, 2. butt type (the part of the core’s striking platform, detached during the removal) –enables the recognition of knapping techniques (possible use of hard/soft hammer) and knapping methods (such as preparation of the striking platform of the core), 3. number of scars on the dorsal face (negatives of previous removals) – determined by the position of the analysed item into the reduction stage and by the number of reduction series, 4. orientation of scars (unidirectional, opposed, radial or multidirectional) – influenced by the reduction scheme applied to the core.

The recognition of repetitive morphological forms enabled researchers to group tools into “types”, defined by characteristic “retouch”. The terms for their definition were taken from ethnographic analogies, preparation mode or supposed function (e.g. burin, scraper, etc.). Types were sub-divided by morphological features (convergent scraper), modes of preparation (dihedral burin) or toponymical relation to a site (Kebara point). Bordes, who concluded that “fossiles directeurs” are meaningless in their singular form, developed a method of comparison between assemblages (Bordes 1950, 1961). The idea was to quantitatively compare between the inventories of tool types from sites. Similarities and differences between the (cumulative) graphs were used for definition of cultures and cultural regions. Following intensification of statistical methods, alternative typological approaches were proposed, such as the analytical typology of Laplace (1966) or the descriptive morphology of Leroi-Gourhan (1966).

Cores Cores of each assemblage were two-dimensionally measured. Length is the perpendicular line from the most used striking platform, to the surface of the core. Width is measured perpendicularly to the length, at the widest plane of the core along this axis. Cores’ size are influenced by several factors: original size and shape of nodule (before modification), exploitation method (influenced by social factors and personal skills of the knapper) and removals’ intensity (dictated by type and number of needed blanks, the characteristics of the raw material and by social factors).

Two properties characterise types: identity, or identifiability, and meaning, relevant to some purpose. A major “weakness” of typological research is the nature itself of types – partly intuitively and partly rationally, partly essential and partly instrumental. Most typological lists are polythetic, with no fixed criteria of “typehood”. Types are distinguished by norms and tendencies, rather than distinct boundaries (Adams and Adams 1991). Every type necessarily has to have a diagnostic attribute, or a cluster of attributes, that sets it apart from all other types.

Additional observations were: 1. Raw material 2. Relative amount of cortex (the intact surface of the core) – influenced by the degree of exploitation of each core and by the nature of the raw material. 3. Type of removals (flakes, blades or bladelets) – hints at the utilization of the core and thus the industry orientation. 4. Number of the striking platforms and their orientation – reflects the degree of exploitation of the core and the knapping method.

The classification approach adopted is the “lumping” one (Dunnell 1971, Everitt 1974), in which the emphasis is on “external isolation” of types (Adams and Adams 1991), rather than searching for minor variations, which can be the result of “internal cohesion” among similar entities. Following Rosen (1997), tools were divided between ad hoc (scrapers, borers, retouched flakes, notches and denticulates, burins and truncations) and all others (tabular scrapers, retouched blades, microliths and celts).

17

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

The following attributes were recorded for all tool-types subjected to analysis:

Tabular scrapers Tabular scrapers, also known as fan-scrapers (racloir en éventail), were defined by Mallon as a large, broad and thin flake struck from a large plaque of flint, with the cortex on the dorsal surface left intact. Usually the edge opposite the plane of percussion is retouched, giving the implement the shape of an open fan (Mallon et al. 1934). A great diversity of shapes and sizes have been observed: oval, round, elongated or irregular, with sizes from 50 to 150 mm. The length/width ratio suggests four types: fan, round, oval and elongated (Rosen 1983). However, their apparent re-use and sharpening and the great number of broken items recovered from excavations makes such a typology limited for descriptive purposes (Rosen 1987).

1. Measurements – length along the striking axis, width at the maximum perpendicular plane to length and maximum thickness 2. Raw material - pebble, nodule, table, etc. for each type of flint 3. Type of blank Morphological description of tool-types Scrapers Rosen (1997) defines various sub-types differentiated by position and morphology of retouch and by the blank’s type and size. This approach was not adopted since, as shown below, most scrapers were obtained applying a characteristic reduction sequence, oriented for the production of a particular type, the “Be’er Sheva scraper”, recently defined by Gilead and Hermon (in press). Its characteristic attributes are:

Tabular scrapers are usually produced from large flint nodules with flat cortical surfaces. Cores may show preparation on both upper and lower surfaces, and faceting on the striking platform; average dimensions are about 20x15x10 cm. Flakes were struck off by the use of block-on block technique, direct percussion or punch (Rosen 1989). The bulb of percussion is sometimes pronounced, possible indicatives of hard hammer technology (McConaughy 1979:301). In some cases, the bulb of percussion was thinned. Tabular scrapers appear either on longitudinal or transversal flakes. Retouch varies between nibbling and semi-abrupt, almost always on the dorsal face.

1. medium sized wadi pebbles were commonly used as the source for blanks. 2. Its blank is knapped off the wadi pebble with a hard hammer, most pieces feature flat or cortical striking platforms and protruding bulbs of percussion. 3. The common blank type is a primary, rounded flake. 4. The typical tool is massive: long, relatively wide (blades are hardly used), and thick - the average width/thickness ratio is 3:1. 5. The most recurrent type of retouch is scalar, followed by steeped. 6. The location of retouch varies considerably but most common is a convex part of an edge, either at the distal end or the sides.

Given the great variety of tabular scrapers’ shapes, the only recorded attributes were technological observations. Burins Burins are not a typical tool-type of Chalcolithic assemblages (Gilead et al. 1995). Their sub-type classification follows Upper Palaeolithic research (senso lato), as exemplified in works of Gilead (1981), BarYosef (1970) or Goring-Morris (1987).

Other scrapers, which do not fall in the sub-type defined above, were classified according to Tixier’s (1963:54) definition and further described by:

Borers All borers were treated as one group, even though subtypes were proposed and several terms were suggested (Tixier 1963, Bar-Yosef 1970, McConaughy 1979, Rosen 1997). Borers include artefacts with one or more retouched points, commonly sub-divided in awls and drills, even though this separation is not always clear (Rosen 1997:68). Since their definition is related to their point, it was decided that items with a point longer than a third of their length would be classified as drills and the rest as awls. In order to validate this, all borers were first classified and then measured. The results showed that for more than 90% of the items, the results coincided with this criterion (Gilead and Hermon, in press). Therefore, it is suggested to use the value 0.3 of the ratio length-ofpoint/length-of-item as a dividing criterion between awls and drills. Other sub-types would appear following the attribute analysis presented below:

1. Amount of cortex covering the dorsal face of the items – reflects the stage of removals from the core 2. Plan view (rounded, ovaloid, rectangular or amorphous) – determined by the type of retouch and its intensity 3. Type of retouch (for each type of retouch see Inizan et al.1992) – possibly influenced by cultural aspects 4. Shape of the retouched (working) edge (straight, concave, convex or a combination of these types) - possibly reflects different functions

18

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

1. Narrow, but relatively thick, elongated blades obtained from prismatic or pyramidal blade cores, mostly of banded flint (see below). 2. One steeply backed side edge. The cross-section is rectangular or triangular, with a perpendicular side formed by backing. 3. Mostly straight, but sometimes finely denticulated working edge. 4. Truncations on proximal, distal or on both ends are common.

1. Plan view – may represent different functions, and/or different blanks 2. Retouch that shaped the point – may embed cultural aspects, or personal imprints of the knapper 3. Position of point, related to the longitudinal axis of the tool – possible various modes of use/hafting 4. Length and diameter of the point (in mm.) – if used for drilling, these measurements reflect the size of the holes

Regarding the morphology of sickle-blades from other regions, the situation is more complicated, since little research has been done on Chalcolithic assemblages north of the Be’er-Sheva basin. The adopted approach was first to classify artefacts as sickle-blades only items with sickle-gloss in order to obtain a set of morphological characteristics that would enable the definition of this type. Only when a standard set of characteristics is defined, can it be applied to the classification of sickleblades, without using gloss as a defining criterion.

Retouched flakes Retouched flakes are defined as flakes with intentional removals along their edges. To separate scrapers from retouched flakes is problematic. Usually, an item will be classified as a "retouched flake" when the retouch along its edges is irregular, forming a low angle with the edge. Occasionally, "retouch" on flakes can be caused by utilization or by natural factors, such as trampling or rolling (Schiffer 1987). In these cases, it is not always possible to differentiate between intentional and unintentional retouch. Since retouched flakes do not serve as cultural marks, and given their problematic nature, only their quantity was recorded.

Sickle-blades are among the tool-types that are sensitive to chronological/cultural variations in Late Neolithic – Chalcolithic assemblages (Gilead 1990). Moreover, some specialization is reflected by discoveries of sickle-blades production workshops (Roshwalb 1981, Hershman 1987, Gilead and Hermon, in press).

Truncations Truncations are defined as "...pièce dont une des deux extrémités présente une troncature (retouche) normale, oblique, concave ou convexe..." (Tixier 1963:127). The blank can be either a flake or a blade. Artefacts classified as truncations were not further analysed, except the type of truncation, following the definition given above.

Besides measurements, other analysed attributes were: 1. Shape of working edge (denticulated or not) 2. Backing 3. Shape of the distal and proximal ends (truncated or natural) 4. Cross-section (triangle, trapeze, straight angle or rectangular) 5. Appearance of gloss, its amount and its orientation

Notches and Denticulates Notches and denticulates are either flakes or blades with retouch forming one or multiple notches on the edge of the item. When two or more retouched notches form a continuous line, the tool is classified as a denticulate. The possibility that some of the notches are results of postdepositional agents cannot be excluded.

Microliths Microliths are bladelets retouched mostly with partial, fine retouch along their edges. A particular type is a bladelet, either retouched or not, with a modified distal end. It was defined by Gilead (1984) and named microendscraper: a bladelet with a retouched end which forms a distal working end mostly rounded or oblique (see Neuville 1930:69, Nasralleh 1936:306, Gilead 1984).

Retouched blades Retouched blades were divided in two: sickle blades and retouched blades. A common definition of sickle blades rests on the presence of gloss along the working edge and an appropriate morphology for reaping (Rosen 1997:55, Levy and Rosen 1987, Roshwalb 1981). However, since the definition of other types is based on morphological and not on functional characteristics (as for sickle-blades, defined by the presence of gloss, a result of harvesting or reaping), there is no need to apply a different approach for their definition.

Bifacials (Celts) Bifacials have a relatively short appearance in the history of stone tool production. First bifacials appeared in Natufian contexts (Belfer-Cohen 1988), and disappeared at the beginning of the EB (Rosen 1997:98). Most bifacials were found in proto-historic assemblages, either Neolithic or Chalcolithic and their typology was discussed in various publications (Levy and Rosen 1987, Neuville 1934b, Roshwalb 1981, Rosen 1997). Several sub-types were classified as follows:

Negev sites’ sickle-blades were defined on morphological considerations (Gilead et al. 1995:255):

19

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

• Technique is explained by means of:

• Axes are traditionally defined as bifacials with a bi-convex cross-section. However, the emphasis is on the position of the working edge, related to the normal axis of the tools. Looking from a frontal view, it is located at the centre of its cross-section. • Adzes have a plano-convex cross-section. The working edge is at the base of its cross-section. • Chisels are narrow bifacials with straight edges, plano-convex cross-section and one or two rounded working edges. • Picks have a pointed working edge and rounded, cortex covered butt. Their morphology resembles some crude Lower Palaeolithic handaxes. • Another sub-type of bifacials, common in flint assemblages from the north of Israel, is the perforated disc (see Perrot, Zori and Reich 1967). These are flat, bi-convex tools, with rounded edges and hollowed in the middle.

o preliminary knowledge of flint tool knappers (the French terms are savoir-fair and connaissance), influenced by : ƒ mental representations (concepts of toolproduction existing in the knappers’ mind) ƒ memorization of operational modes (the capability of reproducing and repeating schemes of tool production) ƒ flexibility of process-adaptation to constraints (adjustment of tool production to limitations imposed by natural factors, as quality of raw materials, or social factors, as presence/absence of specialists) ƒ capability to adjust production to needs (the relation between tool production, in terms of intensity and morphology of products, and the required needs from them) o the technical process itself, which is a suite of gestures and operations that can be traced in: ƒ complexity of the operational scheme ƒ quality of production (observed on flint supports and tools) ƒ knapping methods adopted for the production of tools ƒ variability of reduction sequences applied to production o the objects (flint tools) themselves, their shape being dictated by: ƒ functional needs of their users ƒ aesthetical requirements, reflected in the final shape of tools ƒ traditional marks, recognizable on the morphology of tools

These sub-types may represent different functions: Egyptian wall paintings representing agricultural works show the use of adzes for ploughing (Hatem 1976). It seems that axes were used for a completely different purpose: cutting hard material (be it wood or stone) (see Mackay 1921, Miller 1987, Setton-Karr 1905). Several attributes were observed: 1. Blank type used (nodule or large flake) – reflects the mode of manufacture, which may have cultural connotations 2. Plan view (triangular, trapeze or rectangular) – denotes typological criteria and possibly is influenced by function 3. Cross section (triangle, trapeze, rectangular, elliptic or rhomboidal) – influenced by different schemes of knapping 4. Delineation of working edge (perpendicular to the longitudinal axis of the item - straight, concave, convex or oblique) – reflect possible various modes of production and/or use 5. Condition of working edge (fresh, damaged or resharpened) 6. Shaping method of the working edge (longitudinal or transversal removals)

• Technique is affected by: o the nature of site (type of habitation site, or a specific task-oriented, as a quarry, a tomb, a workshop, etc.), observable in: ƒ expectable content of the tool assemblage ƒ presence/absence of tool production stages ƒ distinct characteristics of tool production (such as special treatment of tools, choice of raw materials, etc.) ƒ spatial location of the site (accessibility to raw material sources and their quality, proximity to contemporaneous sites, etc.), observable on: • exploitation degree of raw materials • quality of tool-production o economic ties between neighbouring sites o social organization of flint-tool producers (social complexity/stratification, size of human group and its components), reflected in: ƒ centralization of specific tools production ƒ standardization of production ƒ diversification of tools reduction schemes

Comparative framework The analysis and comparison of the flint assemblages presented above is based on the theoretical framework presented above. The basic assumption, quoting Lemonier (1986, 1990) and others, is that technique (in this case of flint tool production) is a social action, and therefore, its analysis may divulge socio-economic and cultural aspects of societies involved therein. The graphical representation of the theoretical framework applied during the interpretation of the analysis results of the flint assemblages is presented in Figure 7. Several steps were taken into consideration:

20

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

o economic factors, as existence needs, economic independence, existence of goodsexchange networks, etc. recognized by: ƒ intensification of manufacture of particular products ƒ presence of products obtained by means of goods-exchange ƒ adjustment of production according to economical needs o traditional/cultural factors that influence the decision-makings of flint knappers during their tool-production, traceable in various characteristics of the industry, which cannot be explained in other terms, as functional or economical.

2 Stage 1 – Production (cores analysis and their exploitation mode): 2.1 Exploitation degree: extent of non-used débitage surface, estimated according to the amount of cortex retained thereon 2.2 Aiming: recognition of removal types knapped from cores – blades, bladelets or flakes, observed on the negative scars on their débitage surface 2.3 Initial shaping: observations of the striking platform, whether is faceted or not, and recognition of further investment in corepreparation, as reflected by the presence of core trimming elements 2.4 Knapping scheme: pre-determined or not, observed on removals type, number of platforms and their orientation and raw materials used 2.5 Core exhaustion: the ratio size of core/blank size, hypothetically removed from that core

• The characterization of tool production is based on: o observations on raw materials and their possible sources o attribute analysis of selected tools and waste products o calculation of various ratios between meaningful variables o comparison between assemblages, at different levels of observations o analysis of spatial distribution of selected variables

3 Stage 2 – Production of blanks (observations on blanks’ morphology): 3.1 Products variability: calculation of the variability of blanks size 3.2 Blank production: pre-determined or not, according to selected morphological characteristics, as butt type and scar patterns 3.3 Production efficiency: the ratio between tools and their blanks, which remained unused. The locality of production of blanks must be specified 3.4 Production standardization: adjustment of knapping method, knapping technique to qualities of raw materials, according to their types 3.5 Production variability: number of knapping methods applied for the production of a specific blank

In order to perform a comparison between assemblages, several variables were defined, related to each step of tool production, as defined by the stages of a typical chaîne opératoire for flint tool production (see above). For each set of variables, a set of logic statements was affirmed, which would eventually lead to a characterization of the flint tool production and thus an understanding of socioeconomic and cultural characterization of human groups of the Chalcolithic period: 1 Stage 0 – Raw materials' acquisition (for each type of raw material):

4 Stage 3 – Tools retouching:

1.1 Variability of raw materials: number of raw materials’ types used, recognizable according to texture and cortex 1.2 Abundance of raw materials: estimated distance between the source of raw materials and the location where they were found (the analysed site) 1.3 Locality of modification: presence/absence of waste products on observed raw materials, at each of the analysed sites 1.4 Use diversity: number of tool-types made on the same raw material 1.5 Use intensity: number of tools knapped from similar raw materials 1.6 Reduction scheme: observed on each type of raw material

4.1 Variability of types: the number of tool-types in each assemblage 4.2 Orientation of tool-production: calculation of the ratio ad hoc / non ad hoc tools 4.3 Production diversity: the number of knapping methods applied for the production of selected tool-types 4.4 Production standardization: the morphologic variability of selected tool-types – high variability meaning low standardization 4.5 The number of Technological Operational Sequences (T.O.S) applied for the production of selected tool-types 4.6 Resharpening degree: the ratio re-sharpened tools / all tools.

21

CHAPTER II. AIMS AND METHODOLOGICAL FRAMEWORK

IV The analysis of the morphology and production modes of selected tool-types and a comparison between tool assemblages may reflect:

Following the comparison scheme proposed above, several relations between variables may reveal aspects concerning the socio-economic organization during the Chalcolithic period and characterization of its cultural diversity: I

IV.1 The definition of cultural boundaries, usually confined to a limited geographical zone. IV.2 Specialization degree of the tool assemblage, reflected in a high number of tool-types, each produced by a specific knapping method (assuming a different function to each tooltype, within the analysed assemblage), and in a low morphological variability of tools, which can be linked to the occurrence of workshops.

A high correlation between raw materials and tool-types denote a deliberate choice, given either the suitability of raw materials, because of traditional constraints or the result of concentration of production in a particular place. This assumption would be validated when: I.1 Some tool-types would be produced on particular raw materials. I.2 The locality of modification of particular tooltypes would be off-site. I.3 The availability of raw materials would not always affect their use intensity. I.4 Particular knapping methods would be applied on specific raw materials.

Standardization of tool-production, reflected by low morphological variability and well-defined technological operational sequences for each tool-type may be the result of adequate knapping skills of the artisans, or in case of a generalization, specialized production. In the following chapters the lithic industries of various sites will be presented, which form the database of the study. The sites were grouped into geographical units; a different chapter was devoted to each unit.

II Evidence for exchange networks will appear when local production zones are defined, characterized by a high concentration of blanks and/or particular types of tools. Cores will be highly exploited, despite high availability of raw materials. A decrease in the morphological variability of certain products will occur and several knapping methods will appear, applied for the production of specific blanks. These characteristics require: II.1 A priori knowledge of raw materials’ location and physical characteristics. II.2 High correlation between the aiming of cores and the knapping method applied for their exploitation. II.3 Cores will often be exhausted, despite a high availability of raw materials and original shape of nodules. II.4 Production centres, recognized by concentration of singular products. III An increase in specialization will be reflected in a low variability of blanks, a low amount of nonused blanks at “habitation” sites, while a concentration of particular blanks will be found at “production” sites and specific blanks will be obtained by particular knapping methods. These aspects will be estimated by: III.1 Measurements of selected blanks and calculation of their variability. III.2 Attribute analysis of selected blanks. III.3 Determination of applied knapping methods and techniques. III.4 Calculation of various technological indices.

22

CHAPTER III –NEGEV SITES

CHAPTER III

Period III: At a certain stage the inhabitants blocked the opening of pits and subterranean structures with cobbles and left the site. During this abandonment phase, period III, a layer of aeolian sediments was deposited over the site and trapped in its underground elements. The extent of this period is unknown, but it is worth noting that no change is discernable between the pottery assemblages of periods I, II and IV.

NEGEV SITES

Abu Matar (128.7/71.5) and Bir es-Safadi (129/71.2) The sites of Abu-Matar and Bir es-Safadi are located in the Be’er Sheva valley (Map 2). Abu Matar is situated on the right bank of Nahal Be’er Sheva and Bir es-Safadi (Safadi) is on the opposite bank, ca. 700 m downstream. Abu Matar was excavated by J. Perrot between 1952 and 1955. In 1952, while working at Abu Matar, the nearby site of Safadi was found and later excavated by the same mission, from 1954 to 1960 (Perrot 1955, 1957, 1984, 1987, 1990). Due to intensive development of modern Be’er Sheva, both sites were recently re-excavated: Safadi by Baumgarten and Eldar (1985) in 1982 and Abu Matar by Gilead, Rosen and Fabian (1993) in 1990-91. Given their similarities, they are presented together.

Period IV: The architecture differs markedly from that of periods I and II. Underground structures went out of use, and now settlements consist of above ground structures, made of mud bricks on stone foundations. They are rectangular, with an entrance in the narrow side. The bases of walls, ca. 50-60 cm above floor level, consist of wadi cobbles, their upper parts, hardly preserved, were of terre pisée. Fireplaces and basins are scattered in and around the room. This surface featured also shallow depressions, 2 metres in diameter, filled with ashes and debris, oval fireplaces, fosses, basins and bell-shaped pits.

Architectural phases and stratigraphy The deeper parts of both sites are ca. 3-4 metres below the top of low hillocks adjacent to the Nahal Be’er Sheva channel. In his syntheses, Perrot (1984, 1987, 1990) divides the sequence of the sites into four phases:

Absolute dating Several C14 dates were obtained for Safadi and two for Abu Matar. Since all are statistically similar, it was possible to average them (Gilead 1994, Sugar 2000). The mean calibrated range display a clustering of dates between 3940 and 4370 B.C, most probably a few decades around 4100 B.C. This suggests that the settlement of both sites has a limited duration, more or less continuous and probably contemporaneous. This agrees well with Perrot’s (1987, 1991) suggestion that the sites were settled by small communities for relatively short time spans. Perrot’s assertion that there is no change in the local material culture was recently backed by Commenge Pellerin’s (1987, 1991) analyses of the pottery assemblages from Abu Matar and Safadi.

Period I: This earliest phase consists of open-air sunken courts, about 1m to 2 m below surface with a diameter of 4 – 6 m. From such a court, a narrow gallery leads into a large, subterranean room, with a ceiling ca. 1.5m to 2.0 m high. Rooms have a sub-rectangular shape with an average size of 7 x 3 m. In the courts and rooms, there are ashy, occupational floor(s) with scattered artefacts, fireplaces, basins and storage facilities, mainly bellshaped pits with volumes ranging between 500 to 1500 l. Period II: The inhabitants of period I probably underestimated the durability of the underground rooms. When ceilings collapsed, the nature of underground structures changed and period II began; after an underground room collapsed, the debris were removed and the surface levelled. On it, a base of large wadi cobbles was laid and a wall of stone foundation and mudbricks erected. This is the basis of rounded structures with walls ca. 1.5 m high, floors with small fireplaces and no bell-shaped pits. The roof was composed of horizontal beams covered by reed and soil, probably supported by a central beam that rested on a large stone.

Digging and flint collecting methods The sites were excavated within a grid system of 5 x 5 metre squares. The elementary excavation unit was the "basket": any removed "slice" of sediment during the process of excavation. Any spatially defined humanmade feature is a "locus": an underground structure, a pit, a fireplace or any other installation. During the 1950s excavations, mostly tools and cores were kept, while waste artefacts were collected, counted and most discarded, keeping a large enough sample in order to conduct a detailed attributes analysis.

A second component of period II underground architecture consists of multi-cellular structures – groups of small chambers, sometimes in a sequence, accessed by rounded and small size galleries. Galleries are very short but sometimes they are long and deep with openings into secondary ones. The area of each chamber is about 9 sq. m and its height is 2 m. Numerous artefacts were scattered on the floors. Fireplaces, basins and bell-shaped pits are common at the complex’s entrance. There are hardly any sunken courts, typical of the previous period, associated with multi-cellular chambers and galleries.

Description of the flint assemblages The flint assemblages collected during Perrot’s excavations of the 1950’s are described below. The flint assemblage collected during recent excavations at Safadi were published elsewhere (Hershman 1987), while the flint assemblage of the renewed excavations at Abu Matar (Gilead et al. 1993), are being analysed.

23

CHAPTER III –NEGEV SITES

Rosen (1997) regards this flint as originating from Eocene outcrops in Har Qeren, western Negev, quarried by hinterland Chalcolithic pastorals.

Stage 0: Acquisition – Extraction and testing of nodules Different sorts of raw flint were used, most of them being collected in the immediate vicinity of the sites, most probably from the nearby wadi of Nahal Be’er Sheva. These are pebbles dislocated from a variety of Upper Mesozoic – Early Cenozoic bedrock strata, most of them with inner impurities and thus of a lower quality for knapping. Common are brown to grey, sometimes brecciated and coarse-grained flint cores and their products.

The use of raw materials is summarized in Table 1. Grey and brown pebble flint dominate, followed by banded and in a much lesser percentage, semi-translucent and tabular flint. A comparison between the relative proportion of the different raw materials used at both sites returned a Χ2 value of 1.709 (for 3 df), which indicates that the distribution of raw materials in both sites is very similar, Chalcolithic knappers using the same sources for flint procurement and using them in similar modes.

A distinct, less frequent raw material, of Neogene origins, is known as “banded” or “striped” (Rosen 1997: 33). It is common in north-eastern Negev, can be found in the channel of Nahal Besor, but not in Nahal Grar. Pebbles of this type are larger than others, the flint is fine-medium grained with alternating bands of different varieties of grey and has higher knapping qualities than common pebbles. It was consistently used to produce sickle blades from Upper Palaeolithic-like blade cores, as noted on most sickle – blades and on the remains of an apparent workshop for sickle – blades production at Safadi, where the excavations of a shallow pit yielded 69 blade cores (all blade cores found at the site) and thousands of waste products. A similar phenomenon was noted at site A in the Nahal Besor area (MacDonald 1932: 10).

To conclude, the use of raw materials indicates that low to medium quality flint was very abundant and accessible and that utilizing flint was a domestic practise. The low frequency of less common or exotic raw materials and their restricted and differential distribution suggests a certain degree of craft specialization. The extraction and testing of nodules consists of collecting suitable pebbles from the adjacent wadi of Nahal Be’er-Sheva, located a few tens of metres from the sites and their modification within the area of the habitation site. Stage 1: Production Cores Almost 400 cores were analysed: 168 from Abu Matar and 313 from Safadi; ca. 90% of them are flake cores (Gilead and Hermon, in press). They are medium-sized wadi pebbles and have flat striking platforms. A subtype of flake cores is a core-chopper (Plate 1), made of similar flint as the flake cores but with a shape resembling Lower Palaeolithic choppers, as defined by Leakey (1971: 4).

Another particular type of raw material consists of Senonian small pebbles, brown to pink and semitranslucent chalcedony, almost exclusively used for the production of bladelets and microliths from typical bladelet cores. At Abu Matar, where only 12 microliths (1.32% of all tools) were recovered, 36 bladelet cores were recorded. At Safadi, where the assemblage is much larger, not even one such core was recorded although there are 26 microliths. It seems that in this case too, the differential distribution of the translucent flint reflects a certain organizational level beyond the domestic production that characterizes the use of the standard pebbles. Besor site A (Macdonald 1932: 10) is relevant again since it yielded hundreds of microlithic cores, “…sometimes scuffed into a broken pot...”, and thousands of blanks removed from such cores.

Flake cores These are highly non-standardized and feature a globular, amorphous shape and short working edges; more than half of them have one flat striking platform. Half of the single-platform cores remained unused on 50% of their débitage surface, some 2% have no cortex while 10% have more than 50% cortex on their circumference. A third of flake cores have two striking platforms, either opposite or dihedral. Least frequent are cores with three striking platforms (7%). They are the largest among the group, (and thus the extensive use); all have less than 50% cortex on their surface. The expenditive method of flint processing, probably facilitated by proximity to raw material sources, is reflected in the limited exploitation of cores (most were discarded still covered by cortex on half of their surface) and the presence of pebbles with a single large flake removal (probably used for the manufacture of scrapers) and by core-choppers, usually with only two or three dihedral removals. The number of scars on the last stage of removals is 2 to 13, most common being 4 to 7, related to the number of striking platforms – three platforms cores have an average of 10-12 scars.

A distinct and unique raw material is tabular flint, used solely for the production of tabular scrapers. The nodules are brown, fine to very fine grained and the flint sometimes fractures along flat surfaces to produce large and relatively thin flakes with a flat ventral aspect. Since there are no waste pieces in the assemblages, it is assumed that finished products were imported to sites. The original location of this raw material and its geological source is not clear. Perrot (1955: 179) reports that R. Neuville discovered near Quseime, in eastern Sinai, a workshop in which only tabular scrapers had been produced. Perrot suggests that tabular flint is of Senonian age and was quarried in the Judean desert.

24

CHAPTER III –NEGEV SITES

According to this approach, core-choppers should be viewed as flakes cores.

Blade cores 8 blade cores were found in the Abu Matar assemblage. The 69 Safadi blade cores were found along with hundreds of blades and blade fragments, in a shallow basin, probably an atelier for sickle-blade production.

Most core-choppers are on pebble flint of various sizes, collected from the adjacent wadi of Nahal Be’er-Sheva. This is also the common raw material, used for the production of most tools; a 1/5 are on banded flint. This amount probably reflects their frequency among other types of pebbles, at their source. Brown flint with limestone was rarely used. Pebbles with a globular shape (range of length/width ratio 0.75 to 1.35) and similar sizes were intentionally chosen to be modified into corechoppers. From these, 4 to 6 large flake removals were knapped. The similarity between the size of the biggest scar on the débitage surface of core-choppers, supposedly used for as a blank for flake tools, and in particular scrapers, and the size of scrapers, indicate that the former may have been used as a blank for the later.

All but 3 blade cores (on brown nodule flint) are on banded flint pebble. They are conical or pyramidal, with only one striking platform. Usually, half of the surface was left intact, covered by cortex (Plate 1:2). Since raw material was reachable within a few minutes walking, this apparent wasteful cores exploitation is not surprising. The length blade cores varies between 31 to 93 mm (average 60 ± 12 mm); width values span between 70 and 22 mm (average 43 ± 10 mm); the distribution of both measurements is normal, indicating that these items form an homogenous group.

Several use-signs were noted: along the cutting edge, i.e. the edge of the striking platform (fractures, striation marks or polish) and on the butt (buttering signs). The shape of some core-choppers, resembling axes, may reflect their use as tools. Modification marks on the cutting edge are small fractures or polish, probably the result of the use on a particular material, and not intentional polish prepared by the knappers. Buttering signs are small fractures on the cortex and may reflect their use as hammerstones. It should be noted that more than two thirds of the analyzed items lack any use-signs.

The common number of scars is 6 or 7 (in almost half the sample). Less than 10% have 2 or 3 scars. A quarter of them display a more intensive exploitation: 8 to 10 scars. Bladelet cores Only few bladelet cores were found (N=35), all at Abu Matar, all on semi-translucent flint, with lengths ranging between 20 - 30 mm. The presence of such cores in one site, along with microlithic tools of the same raw material indicates an on-site production of microliths.

The amount of cortex on core-choppers is similar in both assemblages. The possibility that items with more cortex were used as tools was tested: 50% of items with use signs preserved 75 % of their surface covered by cortex, suggesting that the removals were the result of preparing a cutting edge. The possibility that some of them were occasionally used as tools cannot be rejected, but cannot be statistically proven.

Core-choppers Core-choppers (Plate 1:1) were classified according to the type-list used for the analysis of Lower Palaeolithic sites such as Olduvai Gorge (Leakey 1971) or ‘Ubeidiya (Bar-Yosef and Goren-Inbar 1993). They were defined as “... cobblestones with rounded cortex surface forming the butts end…trimming is bifacial, with multidirectional flaking of the working edges…” (Leakey 1971:4). The possibility that choppers were used as cores was raised by Leakey, who concluded “whether certain specimens should be described as cores is largely subjective…it is inapplicable to any of the material from Beds I and II”. This conclusion was adopted in the analysis of ‘Ubeidiya material as well: “…we employ Leakey’s categories of “chopping tools”, without implying that they were in fact “utilized” (Bar-Yosef and Goren-Inbar 1993:125).

Discussion Pebbles were on-site decorticated by a limited number of removals; some of these served as blanks for scrapers. There is no evidence for a particular shaping of pebbles; in most cases, this stage was limited to the preparation of one unfaceted striking platform, obtained by one large removal. The same method was applied for flake, blade and bladelet cores. The blade cores cluster of Safadi, found in a shallow pit on the surface of the site and the bladelet cores found only at Abu Matar indicate a partial concentration of production of sickle blades and microliths in a knapping workshop. Core-choppers could have been exploited for the production of blanks for scrapers, as heavy-duty tools, or both. The performed analysis could not elucidate on whether they were used solely as cores, or the striking platform served also as a working edge. The presence of damaged working edges and battering signs hint that some activities were performed with these artefacts.

Toth and Shick (1986) and followed by Potts (1988), forwarded the possibility that core-choppers were used as cores: “…the various tool types…do not represent predetermined goals of tool manufacture. Rather, they signify a continuous range of forms produced as a result of making flakes. The shape and edge features…depend on the size and shape of the original pieces of raw material and the degree to which these pieces were subjected to flake production” (Potts 1988:235).

25

CHAPTER III –NEGEV SITES

simple blanks, and not because of limited knowledge of flint knapping.

Stage 2: Production of blanks Flakes

The flint industry is oriented towards the production of flakes. Apparently, most of the typologically defined blades are the result of the same reduction sequence as flakes and can thus be referred to as "elongated flakes". Some blades, most void of cortex, with a punctiform butt and narrower than the others, may have been deliberately produced as blades, involving the use of a soft hammer, at later stages of core exploitation and/or a change in knapping techniques, as core exploitation advanced.

Most flakes have a unidirectional scar pattern on their dorsal face and a flat butt. Cortical butt flakes and punctiform butt flakes indicate that the entire process of core exploitation was normally oriented towards the production of flakes, at each stage of their knapping. Items with a multidirectional scar pattern tend to be smaller than the others. The similarity of means and standard deviations, the normal distributions of both samples and their overlapping, indicate that flakes from both sites are virtually the same, obtained from similar raw material sources and represent a single, homogenous knapping tradition, with similar removal schemes and techniques. Most probably flakes were produced by different knappers, applying a minimal preliminary preparation of the core. All these suggest an opportunistic manufacturing method, with the emphasis on expediency, rather than on optimisation of core exploitation.

The frequent use of flint, its distribution and the simple method of its processing indicate that flint was worked in every household and products were used in the routine domestic activities. In a number of locales or occasions, flint was processed in a more professional manner. Stage 3: Shaping/Retouching stage – retouching of tools There are 3609 tools in the assemblages of Abu Matar and Safadi: 901 in the first and 2708 in the second. Since most waste products were discarded in the field, the relative amount of tool types warrants a note. We know that obvious tools such as bifacials, scrapers or sickle blades were not discarded by the excavator of the sites. In retrospect, we suspect that possible candidates for rejection were some ad hoc tools, such as retouched flakes or blades, notches and denticulates, pieces that even now are problematic. Some, for example, define as tools even pieces with signs of utilization or slight trimming (i.e. Levy and Rosen 1987: 285), while we tend to define such pieces as waste in most cases.

Blades Blades were mainly removed from one striking platform cores or from amorphous cores, as evidenced by their scar pattern. Few were knapped from bipolar cores, but none from radial cores. It is clear that cores used for blades production are basically similar to those from which flakes were removed. The reduction sequence aimed at the production of these blades does not markedly differ from that of flakes. Therefore, the possibility that, technologically speaking, at least some blades are in fact elongated flakes cannot be rejected.

Scrapers

When compared with the flakes, while flat butt is the preferred in both groups, punctiform butt is more common in blades and cortical butt is more frequent among flakes. This comparison suggests that: a) blades and flakes were produced by similar methods b) on few occasions the production technique of blades was different, as reflected by the punctiform butt.

The “Be’er Sheva scraper” (N=974; 27%) Scrapers are among the most frequent tools (ca. 1/4 of all tools in each assemblage) (Table 3). A particular subtype, characteristic of the Be’er-Sheva assemblages, was recently defined (Gilead and Hermon in prep.), based on an analysis of more than 900 scrapers. These are on large (75x60x20 mm) cortical flakes, the larger ones among the available blanks; a few items are massive (lengths exceed 12cm). The working edge is shaped by scalar to steeped retouch and covers the distal , the side or most of the circumference (Plate 2).

Summary All attributes presented above indicate that the inhabitants of both sites shared the same tradition of flint processing, producing indistinguishable waste products. Over 80 % of flint products are on pebbles. In many aspects, the industry may be referred to as a Galets aménagés one, with numerous choppers of a Lower Palaeolithic style, some with battering signs and edge damage, scattered all over the sites. A high amount of pebbles was collected from the adjacent wadi, brought to the site and after a minimal decortication and core preparation, a few flakes were removed. Cores were discarded after a minimal exploitation, probably due to the wealth of available raw materials and the needs for

Perrot referred to them stating that: “…the craftsman has usually confined himself to chipping off one end of a pebble and rapidly retouching the thin edge of the flake.” (Perrot 1955: 78). An analysis of their attributes (Table 4) shows the defining attributes merge one into the other or cross cut assumed sub-types in what seems to be a result of random variability, which indicates that the common mode of defining sub-types, with its roots in the research of Palaeolithic assemblages, cannot be applied here.

26

CHAPTER III –NEGEV SITES

Massive drills (N = 46) were classified according to Tixier’s (1966) definitions. Grand perçoirs are on pebbles with a natural, rounded butt. The point was shaped by two longitudinal removals along edges and abrupt sharpening at the end. Some are on natural backed blades (N = 9), on massive flakes (N = 5) or on ridge blades (N = 3). They were used for drilling holes with the largest diameter, of 10 to 15 mm. One item was smeared with ochre along its faces and edges (Plate 4:2).

The number of scrapers that are of a different type is very limited. These are on flat, non-cortical or slightly cortical flakes or blades with parallel or abrupt retouch. There are a few dozens of such pieces that constitute less than 5% of the scrapers, quantitatively insignificant. Tabular scrapers (N=63; 1%) Tabular scrapers are very infrequent: less than 1% in Abu Matar and ca. 2% in Safadi (Table 3). Since the sample is small and many pieces are broken, no detailed analysis was carried out. All items are made on brown, finegrained, very flat flint that is not available in the vicinity of the site. Usually blanks are completely covered by cortex on their dorsal face. The common retouch is parallel, semi-abrupt. In a few cases, mostly on items on transversal flakes, thinning occurs, the bulb of percussion being removed (Plate 3). The cortex on the dorsal face of a few items was smoothed. Tabular scrapers have various size and shapes: transversal, rounded or amygdaloidal.

Regular drills are on flakes or blades (N = 64) and a few on chunks (N = 5). The point was shaped either by abrupt retouch on both edges, or by two retouched notches. The mean extent of the point is almost half the length of the blank, with a mean diameter of 9 mm. The third group of drills consists of narrow items, which include mèche de forets, with or without a tang (N = 20 and 13, respectively), coarse drills (N = 12) and microdrills (N = 6). The first two sub-groups were made on blades, with abrupt retouch along all (N = 13) or almost all (N = 20) of both edges, the appearance of a tang reflecting probably a different hafting method. There is also a minor difference in the mean diameter of the points, items with tang having a bigger diameter than the other group, by one millimetre (Table 5). Coarse drills are similar to massive drills, both in morphological and technological aspects. These are also made on pebbles, and are the biggest among this group. Micro-drills (N=6) are on narrow blades or bladelets, with the point obtained by abrupt retouch along edges (Plate 4:1). The diameter of the point is the smallest among drills (d=8 mm).

Borers (N=478; 14%) The borers are relatively frequent in the assemblages – ca. 14% (Table 3), awls being slightly more frequent. Micro-borers, common in proto-historic assemblages (cf. Gilead et al. 1995, Rosen 1997) are limited to 6 items. Awls The most frequent awl type (70% of all) is simple, on grey pebble flake (Plate 4:3), mean sizes of 62x46x19 mm. (S.D. is 11, 10 and 5 respectively) and a point obtained by two notches at the items' distal end. Fine ventral or abrupt retouch, or slight trimming was occasionally applied to shape the points. Nine awls are on large chunks, their point being modified either by the retouch of one or two notches, or by abrupt retouch. These are among the largest pieces, with the highest ratio length of point/item’s length (Table 5). Less than 20% are on blades; 10% are on primary elements. The variety of blanks used for their production reflects a repertoire of blanks available to the knapper, and not a particular selection of blanks, nor a particular reduction sequence.

Retouched flakes, truncations, denticulates (N=884; 26%)

notches

and

Retouched flakes, truncations, notches and denticulates constitute ca. a quarter of the tool assemblage (Table 3). The preferred blanks used were flakes. They were grouped under the term ad hoc tools (Rosen 1997: 106) in order to emphasize the non-standardized nature of their manufacture. Probably some items were modified by natural factors, such as trampling (Schiffer 1987). Beyond the fact that such tools are known from any Chalcolithic site, no further comments are warranted.

Drills

Retouched blades and bladelets

Drills were usually modified by a parallel abrupt retouch along both their edges that shaped a tool with a very pronounced point (Plate 4:2). Their average size are 69x36x18 mm (STD = 15, 12 and 7 respectively).

Sickle blades (N=561; 16%) In order to establish our typology we conducted a comparative attribute analysis, between sickle blades with and without sheen (Table 5). The 561 sickle blades (Plate 5) of both sites constitute about 16% of tools. While sickle blades from Abu Matar are almost equally divided between items with and without gloss, there are significantly fewer glossy pieces at Safadi (Table 5), a bias caused by recovering remains of a sickle blade atelier at the site, with 69 cores and hundreds of blades discarded during the manufacture of sickle-blades.

Drills were divided in three sub-types: massive, regular and narrow, on the basis of morphological and technological grounds (Table 5). Almost half of the points were shaped by abrupt retouch along both edges or by the fine retouch of two notches. By averaging width and thickness of the points, the diameter of drills was estimated, mostly being used for drilling holes with a mean diameter of 10 mm (Table 5).

27

CHAPTER III –NEGEV SITES

Almost half of the sickle blades are not truncated. The very high frequency of non-truncated items in the atelier from Safadi is a bias and the value from Abu Matar seems to be the more realistic. One third of the sickle blades have one straight truncation. The remainders have two straight truncations, either dorsal or ventral. The truncations were probably used to control the length of the sickle blades and to shape a blunted, straightened end, to better facilitate the process of assembling a sickle.

Remembering the crude nature of the standard pebble flake industry described above, sickle blades originate from a completely different reduction sequence. This is assessed when comparing attributes of sickle blades with those of typical blank blades. A clear difference is that most sickle blades are on banded flint while blank blades, along with retouched blades, were made of common wadi pebble flint (see above). The differences in size are also evident: while the average width and thickness of sickle blades are 11.23 ± 1.6 mm. and 4.13 ± 1.2 mm respectively, the average width of blank blades is 24.9 ± 13.83 mm and the average thickness is 8.4 ± 4.38 mm. The distributions of the width of sickle blades vs. blank blades (Figure 1) further demonstrate the large gap between the two. The scars pattern on the dorsal face of sickle blades is always parallel, indicative of a systematic blade removal scheme. The scar patterns on the dorsal face of blank blades, and the virtual absence of nonbanded flint blade cores, suggest that many of them are results of a flake reduction sequence.

Two-thirds of sickle blades have a finely denticulated working edge. Probably more items had such a working edge, which abraded during use. This suggestion is supported by the fact that more sickle blades without gloss (unused or slightly used pieces) are denticulated. During the excavation at Safadi, a particular locus was uncovered – Loc. 689. It was described as a “bassin” (a shallow pit), found on the excavation surface; it contained remains of pottery, flint products (N=3777), mostly blank blades, retouched and broken blades as well as finished products (sickle blades). The locus served as a refuse pit, containing the waste of an atelier specialized in the production of sickle-blades. Few conjoins and an analysis of the artefacts enabled us to reconstruct the stages of the sickle blade production sequence. The presence of such a locus is important in terms of socioeconomic organization of the Safadi community.

Table 5 and Plate 5 illustrate some typical features of sickle blades, with and without gloss. Both groups have similar frequencies of finely denticulated working edge (90% of items), backing (all items) and truncations (60% of items). Their sizes are also similar. Their resemblance suggests that three groups form the sickle blade family: used and discarded pieces with gloss, unused or slightly used items and items in the process of production. The cross-sections of sickle blades correlate with type of retouch and number of dorsal scars. The common rectangular cross-section results from a low back and two parallel scars on the dorsal face (Plate 5:7). The next frequent type is a straight-angle triangular cross-section, formed by one scar and a complete backing of the edge (Plate 5:3). These types appear on 98% of sickle blades, in similar proportions. Few items, not backed, have an isosceles triangle cross-section (Plate 5:5). Another rare cross-section is trapezoidal, seen in items with three or more parallel scars on their dorsal face (Plate 5:6).

The lithic assemblage of this locus is dominated by fragments of blades, most of them backed. After core decortication and preparation of a suitable débitage surface, a systematic knapping of cortex – free blades occurred. The most important component of size modification is the use of the microburin technique (MBT), applied in order to standardize the width and the length of blanks, probably relatable to a standard size / shape of sickles. The re-invention of this technique by the Chalcolithic knappers, after a virtual absence of more than 5000 years from the local prehistoric record is indicative of their knapping capabilities.

The number of scars on the dorsal face of a sickle blade depends on its place in the reduction sequence and its location on the débitage surface of the core. At an early reduction stage, items with a triangular straight-angle cross-section and one dorsal scar were produced, followed by blanks with an isosceles triangular crosssection and two parallel scars on their dorsal face, or a rectangular cross-section. A subsequent stage produces trapezoidal cross-section blanks with three or more scars on their dorsal face. The low frequencies of such pieces hint that mostly blanks knapped off during first and second stages of cores exploitation were modified into sickle blades, which denotes a controlled and careful selection of blanks, its result being a morphologically uniform assemblage of sickle blades. The standardized nature of this production is thus well reflected.

In order to reduce the width of blank blades, a notch was retouched at a point along the edge (Plate 6:4). Force was applied obliquely to the edges and a piquant-trièdre was obtained. These items were further backed and truncated till obtaining typical sickle blade shapes (Plate 6:1,2,5-7). Too long pieces were also shortened by the use of MBT, obtaining typical microburins (Plate 6:8-12). Four different directions of break were observed on the basis of refitted items (Plate 6:13-16). The possibility that items regarded as typical results of a MBT (such as Krukowski microburin) are in fact accidental breakage of blades is not supported by the lithic evidence, such as described above, its quantity and the repeated intentional and constant application. Moreover, scars of MBT were discernible on sickle-blades as well, when truncations did not remove them (Plate 5:2,4,10,11).

28

CHAPTER III –NEGEV SITES

Three key elements played a major role in the production of sickle blades: the need for a cortex-free blank, a standard width and a limited length, dictated probably by the hafting method and the shape of sickles. The gloss, along one of their edge, is usually parallel to it, covering only a few millimetres of the dorsal face. This suggests that the Chalcolithic sickle was straight or slightly curved, and the sickle blades were deeply hafted. Such sickles are known from Egyptian tombs (Andreu, 1992).

Bifacial tools (N=551; 11%) Bifacial artefacts (celt tools), are more common at Safadi (16%) than at Abu Matar (5%). They were divided in: adzes (Plate 7:1) axes (Plate 7:2), chisels (Plate 8:1) (found only at Safadi), picks (Plate 9) and two perforated discs (Plate 8:2). Tools are on local large pebbles (Tables 8, 11); the bifacial flaking removed all or most of the cortex. A flat face on ca. half of items indicate that they were made of flakes knapped from very large cobbles.

Retouched Blades (N=81; 4%)

The standardization level of bifacials is low and the dividing line between types is not always obvious. The trivial distinction between axes and adzes may illustrate this point. Axes and adzes share many characteristics, in terms of raw material, size, cortex, cross-section, polish etc. The results of the attribute of each type are presented in Tables 8, 9 and 12 ; our description will treat them together, in order to avoid repetitions.

Since the flint industry is flake oriented, a small amount of retouched blades was expected (Table 3). Their insignificant number suggests that the blade core processing described above was exclusively devoted to the production of sickle blades, a statement supported by the fact that the majority of blade cores and sickle blades are made of the same banded flint, while retouched blades are on regular pebble flint and are practically identical in size to blank blades. The morphological and technological characteristics of retouched blades (Table 7) are greatly different from those of the sickle blades.

Axes and adzes (N=169, N=259; 8%) Cross-section attributes: Observations on the shape of axes / adzes cross-sections and the way they relate to the striking edge have been carried out in order to reconstruct the shaping process of bifacials. Common are triangular or trapezoidal cross-sections (Tables 8, 9). Items with two striking edges have a triangular cross-section. The dorsal removals converge to the longitudinal axis of the items, forming the triangle sides. Triangular cross-section axesadzes were sometimes knapped from three striking edges, the third one running along the apex of the distal face. Along it, flakes were removed to one or two directions.

Almost half of the retouched blades retained cortex on their dorsal face (Table 7). Most are only partially modified along one edge, by fine or scalar retouch. Retouched blades were analysed separately, according to the position and the extension of the retouch along their edges. Given the similarities observed, it seems that the position and the extension of retouch does not imply different sub-types, but rather variations within one type. Therefore, these criteria may reflect variations imposed by blank shape and not functional aspects.

Axes / adzes with trapezoid or rectangular cross-sections were shaped from four striking platforms, located along edges and along their dorsal face. Unique axes feature a rhomboidal cross-section, shaped from four striking edges: the standard ones along the sides, one at the apex of the ventral face and one on the apex of the dorsal face.

The retouched and backed blades are made of two mutually exclusive types of tools: sickle blades and retouched blades. The two groups differ in almost every attribute: raw material, size and extent of cortical cover (Table 6 vs. Table 7). Retouched blades are an ad hoc by product of the pebble industry described above and could have been produced anywhere in the sites, by any community member. Sickle blades are products of knapping skills and techniques restricted to an individual, or few, hence the isolated locales of production.

Working edge: about 50% of them are straight; the rest are divided among convex, concave and oblique. These definitions are somewhat arbitrary since tools are not standardized and for one the slightly convex may be defined as such while for another it is still straight. Apparently, flint knappers attempted to achieve a straight working edge for most tools.

Microliths (N=36; 1%) There are only 36 microliths in both assemblages (Table 3); 5 are micro end-scrapers (Plate 4:4-7). Others are bladelets modified by fine retouch, ventral or dorsal, along their edges. It is possible that some of the broken at their distal end pieces were also micro-end scrapers. All microliths are on semi-translucent flint, indicating a deliberate selection of raw material for their production. This choice was dictated by its qualities, more appropriate for the production of bladelets. The concentration of bladelet cores (N=35) uncovered at Abu Matar and their absence at Safadi, suggests that their handling was practised in workshops.

Working edges were shaped by longitudinal or transversal removals, or a combination of both. Common among axes (Table 8) are longitudinal removals on both faces of the edge, or longitudinal on one face and transversal on the other. 2% of axes have a working edge shaped by two transversal removals. Among adzes (Table 9), the patterns are similar, though transversal blows are more common. The removed longitudinal blanks on 7% of adzes are bladelets. These items were more carefully shaped, as seen in the more regular size and shape.

29

CHAPTER III –NEGEV SITES

sections are triangular or trapezoidal-rectangular and the plan view is triangular, rectangular or trapezoidal. The working edge is mostly straight and polish is very rare.

Plan view: these are trapezoidal, rectangular, or triangular. Their frequencies vary between sites (Tables 8, 9 and 12), although all three shapes are represented. Since there is no clear relation between the plan view of axes-adzes and other attributes analysed, it seems that the plan view depended largely on the original shape of the blanks from which tools were manufactured.

An important aspect is the shaping of bifacials, as reflected in their cross-sections. An angular cross-section is formed by knappers who used three and sometimes even four striking edges, some on the apex of a dorsal face or along one of its sides. Since such striking edges had not been flattened in the final stage of modification, the result is a crude looking end product.

Polish: Polish is rare (Tables 8, 9). It is unknown to which degree polish improves the quality of the working edge, in terms of its durability (Boydston 1989, Hayden 1989). Since most bifacials lack polish, it seems that the effort invested in polishing was higher than its expected gain (performance). Since the manufacture/re-sharpening of bifacials was probably a known and easy task, the working edge enforcement by polish was unnecessary, given it a functional and not a symbolic role.

The scars that cover most of the bifacials are large, with pronounced bulb negatives that indicate that the entire process of production had been carried out only with a hard hammer. The use of a hard hammer is also reflected in the sinuous shape of the side edges. These, and the very scarce occurrence of polish, a time consuming process, indicate that the bifacial production had involved neither efforts of skilled artisans, nor investment of time.

Picks (N=98; 2%) Ca. 3/4 of picks are on large pebbles, with a rounded, natural butt, the bifacial removal being used only for sharpening a point. They resemble Lower Palaeolithic pointed handaxes (Plate 9). Other picks are on large flakes. Most retained cortex on their faces. Commonly, the point was shaped by large, bifacial removals. Picks are among the largest tools (Table 12). Their coarse shape and massive appearance implies that they were heavy duty tools, suggesting an activity on hard material (involved in the digging of underground structures, in the metallurgical industry, etc.).

Summary The tool assemblages of Abu Matar and Safadi may be grouped, along with raw materials and blank characteristics, into four different production sequences: • The most common is a sequence aimed towards the production of scrapers, borers and other ad hoc tools. These, and especially scrapers, dominated by the Be’er Sheva scraper, were made on large flakes produced by a sequence of pebble – core, chopper – flake – retouch, mostly scalar – tool – discard. • A sequence aimed towards the production of bifacials, mostly adzes / axes: large pebble, chunks – core – bifacial – rare polishing – re-sharpening – discard. • A totally different, standardized sequence for sickle blades: banded flint pebble – blade – narrowing and shortening (MBT) – backing and denticulating – sickle blade - use, either long, short or unused kept as replacement – discard. • The least known, since the raw material and tool concerned are rare: Semi-translucent flint pebble – bladelet - fine retouch – microlith – discard.

Chisels (N=25; 1%) Chisels, on cores or on flakes, have an elongated elliptic shape (Plate 8:1), with two rounded working edges. Common retouch along edges is scalar; the working edges were prepared in various ways, which indicate nonstandardization. These tools are longer than axes / adzes, with sizes comparable to picks (Table 13). Perforated discs These are two retouched discs, hollowed in the centre (Plate 8:2). Both are completely bifacially retouched; the hole in the middle was obtained by sharp punches with a fine chisel. One item (Plate 8:2), defined by Perrot (et al. 1967: 210) as casse-tête, has ochre traces along its edge. The second item, defined as racloir by Perrot (ibid.), is less symmetrical and the cruder scars indicate that less attention or skill was invested in its production.

In terms of type frequencies, most common are tools on flakes, such as scrapers, borers and a variety of truncated, retouched and notched flakes. Tools on blades are a rare, except for sickle blades, the most common blade tool type. The microlithic component in Matar-Safadi is very marginal. Typical Chalcolithic elements such as fanscrapers and bifacials are present in low quantities.

Discussion Bifacial are dominated by axes / adzes (77% of group). Picks are the largest tools. The shapes of types and their production modes in Safadi are indistinguishable from those of Abu Matar. In both assemblages, the pieces are mostly made on chunks or on very large flakes and the cortex was removed in most of the cases. Most cross-

There was a clear choice for particular raw materials, related to tool types: cortical pebbles and chunks for scrapers, borers and bifacials, banded flint for sickle blades, and semi-translucent flint for microliths. Tabular scrapers, made on an exotic raw material, were most probably brought to the sites as finished tools.

30

CHAPTER III –NEGEV SITES

The standard lithic industry is common in all features. The availability of the raw material, the basic skills needed to produce it and its frequency everywhere suggests that every member of the communities was a potential producer of common flint artefacts for multipurpose use. The procurement of pebbles, their processing and discard were part of domestic activities. From this point of view, practically no inter or intra-site variability could be traced. All periods and parts of each of the two sites feature the same pattern.

The tool assemblage of Abu Matar and Safadi sites are typologically very similar: there are high frequencies of scrapers and other ad hoc tools and moderate to low numbers of other tool types. The common tools were scattered all over the site and their shapes and techniques of production are crude and non-standardized. This suggests that most of them were produced by the site inhabitants within the range of a domestic activity and were involved in daily or seasonal tasks. In addition, sickle blades, standardized in size, shape and reduction sequence, indicate that part of the tool production was carried out by skilled individual(s), supplying their products to a limited circle of consumers, probably the neighbours of the same site or in its close vicinity. The remains of a flint knapper’s atelier in Safadi, with dozens of Upper-Palaeolithic like blade cores, and the concentration of bladelet cores at Abu Matar support the claim that professional production of tools co-existed along with a domestic flint knapping.

Neve Noy (ca. 129/71) The site of Neve Noy is in fact the east and northeast extension of Safadi, dug as a salvage excavation (Eldar and Baumgarten 1985). Its upper layers were destroyed by modern building; thus, mostly underground structures and their adjacent components were preserved. Eight complexes were exposed, with sunken courtyards, passageways connecting underground chambers and tunnels between them. These were filled with alluvial loess mixed with ashes, stones, pottery sherds and flint artefacts; 23 pits were used as dumping zones.

Stage 5: Discard During the excavation of the sites a large amount of flint artefacts, mostly waste products, were discarded prior to any analysis. From ca. 37,000 flint artefacts collected, less than 20% were kept. The curated assemblages consist mostly of tools. According to the field books, most tools were kept; thus, the analysed tool assemblage reflects its real composition at the moment of its discard.

Following is a summary of the main excavated features: Complex 109 – a sunken courtyard (9 x 3 m), with a hearth at its northern part, and pottery fragments around it; 3 pits with few remains of cultural material were found in the vicinity of the courtyard. One empty underground chamber was found opening from the courtyard.

Flint artefacts, especially tools, are not helpful in an intrasite distribution analysis, due to their scarcity on living / activity floors and due to the fact that items were found in dumping zones, such as underground spaces, silos and pit like installations. The majority of pieces produced in the sites were removed from original production or last use loci, either to dumping places or secondary deposition in fills, which could be the outcome of both anthropogenic and natural agencies (Schiffer 1987). Perrot (1984) described the intensity and complexity of settlement and abandonment processes. Subterranean structures collapsed, their content and debris removed, dumped and new structures were constructed where parts of earlier ones stood. Latter, sites were abandoned and erosion filled pits and the still open underground structures, with debris containing surface artefacts.

Complex 110 – empty underground chambers and tunnels connecting them. No courtyard was found, probably due to damage. Few finds were collected from tunnels. Complex 133 – a passageway connecting 4 underground chambers, the smallest containing the dumping from a collapsed pit dug from surface. Other chambers were empty, excluding an elliptic room, with mud-brick niches built along its western wall. Near the entrance of this room the burnt remains of 4 human bodies (adult male, adult female and two children) were found; scattered among the human remains, bones of animals (cattle, goat and deer) were found. Grains were collected from the floor’s room. Four refuse pits were found in close proximity to this complex.

The settlements terminated with a phase of above ground structures that, again, produced debris dumped in pits, silo and other unused features of earlier periods too, to be eroded again and intermixed by erosion and human activities during the millennia to follow. It is evident, in the nature of fills and the way they accumulated, that the original patterns of flint distribution were wiped-out either immediately after or considerably later, and that the spread of flint observable now is almost entirely a product of re-deposition either by humans or by natural agencies.

Complex 102/106 – two main features: an open activity area (courtyard) and underground structures connected with tunnels, opened from a courtyard. The activity area was connected to the underground structures by a corridor. Several small pits were found on the surface of this courtyard, covered with stone slabs that were probably involved in copper activity. A smelting pit was found at the entrance of one of the blocked tunnels. The underground structures (chambers and tunnels) were empty of finds. Four refuse pits were found close to the surface finds of this complex.

31

CHAPTER III –NEGEV SITES

pits, often opened from an ancient surface, emphasize the hazardous aspect of their chances to be recovered during the excavation (often concentrated on areas next to architectonic structures), and thereafter, its impact on socio-economic reconstructions. Thus, for example, when referring to the concentration of sickle-blades found in L208 and not knowing of the existence of the workshop found by Perrot, Hershman relates the finds to sites in the Besor area, especially Ghazzeh A, where a large amount of blade cores was discovered (Hershman 1987:83, 168).

Complex 144 – an open courtyard on the ancient surface, where several industrial activities took place, among them faience beads production and copper metallurgy. A large amount of stone vessels and a ceramic pipe (hidden in a niche) were found on its surface. A concentration of copper artefacts (maceheads, axes, a small pendant and copper slag) was found in a presumably copper workshop. A smelting furnace, made from mud-bricks and stones, was found in a niche west of the entrance. At the lowest level of the floor’s courtyard, several pits were found, some used for storage, one with a concentration of sickle-blades (Hershman 1987: Table 19).

More than half of the débitage and a third of the tools were found in pits of various size and shapes. No correlation was found between the size of pits, their spatial location and content (regarding other kinds of material remains), and the amount and type of flint artefacts. Possible industrial zones (in the restricted meaning of the word) are close to complexes 102/106 and 144. These included several types of activities, such as copper metallurgy, bead production and possibly flint tools knapping. Underground structures were found void of flint artefacts, except a few intrusions from alluvial activity or collapsed pits from the surface.

Complex 104 – a tunnel leading to an underground chamber, both void of artefacts; 3 storage pits were found at a distance of 10 m. from the underground structures. A lack of consistency regarding shape and structure of the underground complexes could be discerned. The general pattern is of a sunken courtyard, from which underground structures open. Modifications in the content of some of them (as blocking of passages and stone-wall building) suggest several stages of their use. The courtyards served for various activities. Most of the material remains were found in pits of various size and shapes. Some of them were initially used as storage pits. Underground structures were usually void of material remains, excluding intrusions from upper (Chalcolithic) layers or loess mixed with material remains accumulations.

Tel-Sheva (135.3/73) The site is located in the modern town of Tel Sheva, on the Northern bank of Nahal Be’er-Sheva, south of Nahal Hebron, on a small chalky hill between these two rivers. It was dug during 1993 and 1995, as part of a salvage excavation project (Baumgarten, pers. comm.). The Chalcolithic site was damaged by later occupations in the area, from Late Roman to Early Islamic periods, and by modern buildings. It consists of open areas with pits of various activities, such as dumping zones or storage pits (some of them bell-shaped pits), rectangular buildings with pebble stone foundations, and subterranean structures, similar to those found at Safadi. In a few areas, where the ancient Chalcolithic surface was uncovered, cup-marks were found.

The lithic assemblage was analysed by Hershman (1987). It is similar to those of Safadi and Abu Matar. Used raw materials are wadi pebbles, banded flint (for the production of sickle-blades), tabular flint (found only on tabular scrapers) and semi-translucent flint, used for the production of microliths. The flint industry is oriented towards the production of flakes; some resemble Levallois flakes, with faceted butts. The presence of primary elements and cores suggests that most of the tools were locally produced.

At least on one occasion, architectural evidence suggests several Chalcolithic occupation stages (one stone structure built on an earlier one). The location of architectural features suggests that there was no planning, the site consisting of clusters of habitation areas, which their contemporaneity cannot be assessed with certainty. In most aspects, the site is similar to others located along Nahal Be’er-Sheva, such as Bir es-Safadi or Abu Matar.

The tool assemblage is dominated by tools on flakes. Notches and denticulates, and retouched flakes dominate the assemblage (40% of the tools). Scrapers and awls are the next common group of tools (14% each). The high number of these tools may represent the frequency of their production and discard, and not necessarily the intensity of their use. Other tool-types appear in similar frequencies as at Safadi. The absence of microliths (except one micro-grattoir) is noticeable.

Description of the flint assemblage The flint assemblage (Hermon in prep.) was collected from both surface and all excavated areas of the site. Since most of the sediments were not sieved, only items discernable by the free eye were collected; intrusions from either earlier or later periods are described separately (see below). The high similarity between artefacts originating from different layers allowed treating them as a homogeneous assemblage.

Sickle-blades represent ca. 10% of tools. Their production place is unknown, since very few blade-cores were recovered. A concentration of sickle-blades was collected from a pit opening from the ancient surface of an “industrial area”. A similar pit, found by Perrot in his excavation at the site (locus 689) contained sickle-blades, waste products and blade-cores. The presence of such

32

CHAPTER III –NEGEV SITES

Blade cores

Stage 0: Acquisition – Extraction and testing of nodules

The blade cores, too few for a detailed analysis (N=7), have one striking platform; the débitage surface covers half of their circumference (comp. Plate 1:2). The high amount of blank blades disaccords with the low number of blade cores (ca. 90 blades for each blade core!). The possibility that some blade cores were discarded in a (not discovered) particular place cannot be neglected. Another possibility is that some of the typologically defined blades are the results of a flake removal scheme.

The dominant raw material originates from wadi pebbles, probably brought from the adjacent wadi of Nahal Be`erSheva and/or its tributaries, distanced a few minutes walk from the habitation area. Common are brown to grey flint, sometimes with microfossils. Brown brecciated flint, eroded from the natural bedrock of surrounding hills, was used as well. Banded Neogene flint, found as pebbles in the local wadi beds, is common for most sickle blades. Senonian semi-translucent flint, found as small nodules on the hills close to the site, was mainly used in the production of bladelets. Waste products on these raw material testify for a local production of most tools. Tabular brown flint was noted only on tabular scrapers. The general picture of the raw material procurement strategy, ways of exploitation and types is similar to the one described for the sites of Abu Matar and Safadi.

Bladelet cores Bladelet cores represent 27% of all cores (Table 2). Most are on semi-translucent flint, either grey (65%) or brown (23%). A dominant type is a single platform pyramidal core (Plate 11). Other types are dihedral and bipolar, and a few amorphous. Bladelet cores preserve a few cortex on their surface, suggesting their exploitation.

Stage 1: Production

The high number of bladelet cores on one hand, and the small number of blank bladelets and microliths on the other, suggest that some of the laters were exported. Another explanation is that most bladelets were used and discarded outside the excavation area and were not found. The high amount of bladelet cores and the absence of most of their end-products suggest a specialization at the site, oriented towards the production of bladelets. This possibility is supported by the fact that circa half of the bladelet cores were found in a shallow pit (L3229), probably a disposal pit from a workshop. A similar phenomenon was described above (see Safadi).

Cores were subdivided into 3 groups, according to their aim: flake (Plate 10:3), blade and bladelet (Plate 11). Flakes were obtained by applying four methods, related to core-type: one platform (Plate 10:3), amorphous, discoidal (N=12, Plate 12:1,2) and core-choppers (Plate 10:4). Flake cores represent ca. 70% of all cores, among them amorphous dominate. Two flake cores are on flake. Cores The flake cores1 (mean size 49.8 x 55.59 mm.) (Table 2), are wider than longer, mainly on pebble grey flint (41%), brown (23%) or banded flint (17%). Most cores retained cortex over 50% or more of the surface, being thus only partially exploited. Only 13% of them have removals from their entire débitage surface.

Summary The presence of all components of a flint industry (primary elements, cores at various stages of exploitation, blanks, tools and debris) suggests that tools were on-site produced. The decortication of nodules consisted of a few removals used for the preparation of usually one striking platform, without an initial shaping of pebbles, sufficient for the preparation of a flat striking platform from which blanks were knapped, apparently without a control of size / shape of desired blanks. Faceting occurs on 10% of cores. A more elaborated preparation of pebbles was applied to produce blades and bladelets, as reflected by the uniform shape of blade cores (Upper-Palaeolithic like) and bladelet cores (resembling Epipalaeolithic cores), each made on characteristic raw materials.

A difference between single platform and amorphous cores is discerned in their preferred raw material: brown for single platform cores and grey for amorphous cores. An explanation is that the application of a particular removal scheme was influenced by the original shape, size and quality of raw material. The various types of flake cores may represent different exploitation stages of pebbles: firstly core-choppers (N=13), single-platform cores (N=45) and ending up with amorphous cores (N=71). The presence of all three exploitation stages imply a non-standardized nature of the flint industry, the choice for a particular reduction scheme being dictated by personal decisions of knappers and their immediate needs, and not by a specific tradition.

Stage 2: Production of blanks The inventory of the flint assemblage is shown in Tables 3 and 13. The industry is flake oriented, flakes being ca. eight times more frequent than blades. This picture is also well reflected in the core assemblage, dominated by flake cores. Debris represent half of the flint artefacts; débitage ca. 40% of the assemblage, flakes being the dominant category (Table 13), the remaining being tools.

1

A group of refitted flakes, removed from a unipolar core is shown in Plate 21. The circumference of the conjoined pieces gives a clue for the original size and the number of removal stages from the core.

33

CHAPTER III –NEGEV SITES

The common butt types are flat (52%) and pointed (32%). Two blade production methods were applied: one similar to flakes, its products being in fact elongated blades, and the second, pre-determined, of blades with a pointed butt.

Flakes A sample of 100 flakes analysed. Most are longer than wider; ca. 30% of flakes are side-struck, forming a distinct group. Some had a predetermined size and form, being products of discoidal cores (Plate 12:3,4).

Summary

Most flakes are on pebble (brown of grey) flint. Only 1% of flakes are on semi-translucent flint, indicating that an initial stage of a bladelet core preparation, performed onsite, included the knapping of a few flakes.

The production of blanks was oriented for the manufacture of flakes, as reflected by the result of the ratio flakes/blades (N=9). A maximum number of six primary elements was removed from each core, many of which were used as blanks for tools. From each core, up to 4 tools were produced, and 36 débitage items; among them, each eight was modified, especially into scrapers.

Ca. 1/2 of flakes have a flat butt. Faceted butt is the next common type, (1/3 of flakes). A comparison of the ratio cores with a faceted striking platform and faceted butt flakes, and simple cores and flat butt flakes shows that more flakes were removed from faceted striking platforms, almost doubling the amount of blanks. Thus, faceting a striking platform apparently increases the potential of cores exploitation. Almost a 1/5 of flakes have a cortical butt, removed at a preliminary stage of core exploitation, during the preparation of the débitage surface. The preparation of cores for systematic blank removals required the knapping of ca. 10 removals from each nodule2. Five flakes have apointed butt, possibly knapped at the initialisation of cores for blade removals.

The main method for producing tools is similar to that of Abu Matar and Safadi: simple flakes, without any predetermined shape or size were struck off pebbles collected from the adjacent riverbeds and knapped with a hard hammer from cores with a simple, non-faceted striking platform. Cores were discarded after a few removals, as indicated by their size, the amount of cortex on their débitage surface and by the low ratio tool/cores. More elaborated methods were applied for the production of blades, probably used for the manufacture of sickleblades, and of bladelets. A cluster of bladelet cores in one locus suggests that bladelets were systematically produced by a part time specialist. It is unclear if the presence of Kombewa flakes reflects the intentional application of this method or not.

The scars pattern on the dorsal side of flakes accords with the core types and their amount: 45% of flakes have either a multidirectional scar pattern (removed from amorphous cores) or a unidirectional pattern (products of single platform cores). Only 2% of flakes have a centripetal scar pattern (products of discoidal cores).

Stage 3: Shaping/Retouching stage – retouching of tools

A distinct group is the Kombewa flakes (N=2). Their removal method consists of using a regularly convex surface to produce circular or oval flakes (Plate 10:1,2). Such a surface can be created by means of percussion, intentionally causing a wide, regularly convex bulb. It is by using this convexity of the lower face to predetermine the shape and the thickness of a second flake, removed from this surface (Inizan et al. 1992). The systematic production of Kombewa flakes is common only in the Lower Palaeolithic (Goren and Belfer-Cohen 1994).

The tools assemblage (N = 748), comprises all common Chalcolithic tool-types. Scrapers (N = 272; 32%) Scrapers, on grey or brown pebble flint, are the most common tool-type (Table 3). The term “Be’er-Sheva scraper” is applicable to the scrapers found at Tel-Sheva. More than half of all tabular scrapers (N=22) are broken. Most are bifacially retouched, the bulb of percussion being thinned (Plate 14:1,2). Few items are elongated, with parallel abrupt retouch along edges; two have a faceted butt, one being side-struck, with a typical fanshape (Plate 14:3). One item has an incision of two intersected lines on the cortical surface, which covers its smoothed dorsal side (Plate 14:5). Along its distal end, sheen can be discerned. Tabular scrapers with incisions are typical of the EB period (Shick 1978, Rosen 1997).

Blades One hundred blades underwent an attribute analysis. The preferred raw material was brown flint, though other raw materials appear as well. The presence of blades on banded flint (18%), characteristic of sickle-blades, indicates some local production of sickle-blades. The low amount of cortex on the dorsal side of blades suggests that most were knapped at later removal stages; more than half of the sampled blades are cortex free.

Burins (N=2; 1%) The two burins uncovered, both on brown flint, are on natural break. The possibility that the burin blow is accidental cannot be rejected.

2

This number was estimated by the formula (primary elements+cortical butt flakes)/cores

34

CHAPTER III –NEGEV SITES

section, with one edge backed. The scar pattern on the dorsal side associated with it is two parallel blade scars. A ¼ of sickle blades have a straight-angle trapeze cross section, associated with three or more blade scars. The remaining items have a straight angle triangle cross section, associated with a plane dorsal side.

Borers (N = 103; 13%) Borers were subdivided in two sub-types: awls and drills. Awls are either on flakes or on blades, with a point obtained by two retouched notches. They are highly nonstandardized, as reflected by the high variation in shape of tool, position of point and size (Plate 15:2-6). Apparently, there was no preference for a particular blank for preparation of awls. Knappers took an opportunistic approach, taking advantage of the preliminary shape of a blank (e.g. natural breakage) to prepare an awl. Awls are ca. 5 times more frequent than drills.

The number of scars on the dorsal side of sickle blades are related to the number of previous removals from the core. An early stage of removals is represented by a straight angle triangle cross section, followed by triangular, the last being trapeze cross-section. Most sickle blades have a triangular cross section, which means that the second stage of removals was more efficient for the production of blades suitable for their modification into sickle-blades.

Drills are made on blades or on flakes retouched until they reached the dimensions of a blade (Plate 15:1). In all cases the retouch is parallel abrupt, forming a back along both edges. The drills are of various size, though most of them were found broken, being defined mainly by their characteristic points. Microlithic drills were not found.

Gloss appears on 84% of sickle blades; 66% of them along the entire working edge, covering it to a depth of ca. 2 mm. The remaining have a partial straight (16%) or (3%) a diagonal gloss. The position and orientation of gloss indicate the position of the item in the haft. Apparently, diagonal or partial gloss indicate end pieces in a curved sickle, and straight gloss indicates the location of sickle blades in the straight part of a sickle. If so, the number of the Tel-Sheva sickle blades could compose 4 to 8 sickles (4 to 8 households harvesting). If the sickle was straight, the number of sickles blades could compose 10 to 30 sickles (with a working edge of ca. 20 cm length). Given a life time span of (at least) several dozen years for the site, this low number of sickles may be surprising, given the presumed importance of agriculture (of cereals) in the economy of the Chalcolithic inhabitants. The possibility that harvest was also done by hand cannot be ignored. Another, more plausible possibility is that the inhabitants took with them utilizable sickles when they abandoned the site.

Retouched flakes and notches (N = 142; 18%), (N = 84; 11%) The main characteristic of retouched flakes, notches and denticulates is their high morphological variability. They cannot be clustered by size or retouch and the notch appears in different places and is of different size. In some cases these modifications may have been the result of post-depositional agents (Tringham et al. 1974). Retouched Blades (N = 27; 3%) Retouched blades have a partial fine retouch along one edge, or are backed blades. The scar patterns on their dorsal side and their size suggest that most are in fact elongated flakes, being removed from flake cores. Two fragments are of interest, being “canaanean blades”, characteristic to the EB period (Rosen 1983). One is on typical red-brown flint, the other on white-striped flint. These are the only items found on this type of flint. They are intrusive in the Chalcolithic assemblage, and should not be included therein (but see Yorke and Levy 1994).

Microliths (N = 30; 4%) Microliths are dominated by micro-endscrapers (Plate 16). A few are partially retouched bladelets. All microliths, except 4% made on brown flint, were produced on semi-translucent flint of various shades.

Sickle-Blades (N = 53; 7%) Fifty sickle-blades underwent an attribute analysis; the results showed their high standardization. More than half of them were produced on typical banded flint. Other raw materials used were grey flint (20% of the sample) and translucent brown flint (16% of the sample). Brown flint, commonly used to produce other tools-types, is almost absent in the production of sickle blades. Probably sickle blades on translucent brown and grey flint were locally produced. The possibility that some sickle blades on banded flint were imported to the site cannot be excluded.

The comparison of bladelets, micro-endscrapers and microliths sizes (Figure 2) shows the high standardization of their manufacture, all belonging to a same production sequence. The high amount of bladelet cores (N=56), compared to the low number of microliths (N=52) point to a local production of microliths and for a possible export of bladelets, either as blanks or as end-products. A comparison between the lengths of bladelets, microendscrapers and bladelet cores (Figure 2) shows a preference for bladelets shorter than 45 mm for the production of micro-endscrapers. The length of bladelet

The common shape of sickle blades is a narrow blade, straight truncated on both ends (2/3 of the complete pieces); more than half have a fine denticulated working edge. Ca. half of sickle blades have a triangular, cross

2

cores follows the length of micro-endscrapers (R = 0.68 for bladelet cores and 0.67 for micro-endscrapers).

35

CHAPTER III –NEGEV SITES

exported. The standardization of sickle-blade production and the presence of a bladelet cores cache can be the result of activities of part-time specialists. However, this specialization was limited, most tool-types being produced on each household, being integral parts of daily activities, tools being made whenever it was needed. This supposition is supported by evidence of the on-site production of most tools, their ad hoc nature, and their high rate of production/discard, such as for scrapers. The massive presence of picks may be related to a particular activity performed with this type of tool, possibly related to the digging of underground structures.

Apparently, below a certain length (25 mm), products of these cores were not used to produce micro-endscrapers. Bifacials (N=42; 6%) The common bifacial type is the pick, similar to Safadi and Abu Matar items: a bifacially flaked pebble with a pointed end (Plate 25). Most (Npicks=27; 4% of all tools) resemble crude handaxes, both morphologically and technologically (Bordes 1961). Other bifacials are adzes (N=10) (Plate 18), axes (N=2) (Plate 19) and chisels (N=3), representing ca. 2% of all tools. The working edge of adzes and axes was modified by longitudinal removals and was polished. The rest of the bifacials were discarded in various stages of production, or are heavily abraded. Their presence suggest a local production of bifacials.

The analysis of the flint material revealed a diversity of knapping methods, characteristic of different prehistoric periods. Thus, Kombewa flakes, core-choppers and picks (resembling crude handaxes) are typical to the Lower Palaeolithic, discoidal cores are typical to the Middle Palaeolithic and blade and bladelet cores are typical to the Upper Palaeolithic.

Summary The flint assemblage is similar in all aspects to the ones of Safadi and Abu Matar. The similarity is expressed both in raw material preference, technological and typological aspects. Thus, microliths were produced almost exclusively from translucent flint, while most sickleblades were manufactured from grey banded pebble flint. While in the first case the preference might be due the physical characteristics of the flint, the case with banded flint is less clear. It can be found as pebbles in the local wadi beds, mixed with other flint pebbles (used in the manufacture of most other tools). The clear adoption of banded flint for the production of sickle-blades can be a result of maintaining a tradition of manufacturing these tools. The possibility that sickle-blades were products of a particular “knapper-school” (its evidence being found in a workshop at Safadi), supports this supposition.

Nevatim (139.6/70.25) The site, excavated and published by Gilead and Fabian (2001), is located east of modern Be’er-Sheva (Map 2), on the eastern bank of the Be’er-Sheva wadi, on a loess covered plateau, ca. 100 m from the river channel. Following is a description of its main characteristics. During a one month excavation, two areas were dug, characterized by a concentration of finds and stone-wall remains (area A) and deep occupation sediments, discernible in a trench dug by a mechanical tool (area C). Adjacent to area A, a depression of ca. 1 m. may have been the entrance, or the collapsed remains, of an underground structure. The surface between the areas (ca. 150 m) was scattered with finds, but no concentrations, or architectural remains were noted. The site was inhabited in limited spots, separated by sterile zones, void of human occupation. The extent of finds cannot be related to the size of the Chalcolithic site, but to various postdepositional processes that deflated its upper parts.

A distinct characteristic of the flint industry is its technological diversity. Several knapping methods were discerned; most common is the production of simple, undetermined flakes from single platform or amorphous cores. Another method was the removal of predetermined flakes from discoidal cores (the shape and size of flakes was controlled by preparations of an appropriate débitage surface on the core). The third is the Kombewa method, in which a predetermined flake is obtained by the convexity of the ventral face of another flake, previously knapped to serve as a core.

The main characteristics of area A (ca. 1 ha), are several stone-built rectangular structures, rounded installations (built of stones) and concentrations of material culture, including copper slag. These features were enclosed by a wall (ca. 20 cm high), discernible on the western and eastern parts. The general shape of the area is of an elongated ellipse, with the poles oriented north-south (towards the wadi). The structures and adjacent installations were concentrated in areas corresponding to the poles of the ellipse; the area between them has a lower concentration of remains. At least two occupation stages were discerned, mainly by superposition of structures. Both phases belong to the Chalcolithic period, and probably reflect local corrections of walls and not two major occupation stages.

Blade/bladelet débitage is another pre-planned organized débitage. Blades knapped from a pre-determined blade débitage are characterized by a pointed butt and parallel scars on their dorsal face. Bladelets were produced from pyramidal cores of translucent flint, almost exclusively used for the production of micro-endscrapers. When translating the results of the lithic technotypological research in socio-economical terms, several conclusions can be assessed. Probably some kind of exchange of lithic products was performed: sickle-blades and tabular scrapers were imported, microliths were

36

CHAPTER III –NEGEV SITES

at Safadi and Abu Matar (14% of the tools). Half of them were produced on typical banded flint. The remaining were made on translucent flint, probably non-used blanks prepared for the production of microliths, or on local pebble flint. The main difference between the tool assemblage of Nevatim and those of Safadi and Abu Matar is the almost total absence of bifacials in the tool assemblage of Nevatim. Only two adzes and three picks were found; axes and chisels were absent. This low amount may be explained in terms of special activities performed at Safadi and Abu Matar, absent at Nevatim.

Metallurgical activity was discerned in at least one spot, where a “furnace” was discovered. It appeared as a small pit (90 x 60 x 35 cm), with mud-bricks supporting its walls. The pyrotechnical activity is evidenced by the burnt soil of the pit (including the mud-bricks), by its content, a very dark grey ash and by slag found in it. The archaeological sediments in Area C were buried in the covering loess, ca. 3 m. from the surface. Two Chalcolithic layers were discerned: the first, heavily disturbed by modern ploughing, contained a “living floor” with a few small stone installations, remains of material culture scattered around and an ash pit; from the second layer a massive wall (ca. 0.8 m wide) was exposed. In its immediate vicinity large pits, containing ashy sediment, pottery fragments, flint artefacts, bones, copper slag and animal bones, were excavated. Additional pits contained “masseboth”, large (ca. half a metre height) rectangular blocs of limestone, bi-facially shaped by large removals. An additional pit contained a burial of an infant, its bones being found in a large jar.

Mitham C (130.9/72.5) Introduction The site, located near the modern market of Be’er-Sheva, was dug during one month of 1992; two areas, A and B, were opened (Fabian, in prep.). It was heavily disturbed by modern development of the area, which damaged several layers of the site, from Byzantine, Iron Age and Chalcolithic periods.

The flint assemblage resembles those of Safadi and Abu Matar, in aspects concerning raw materials procurement strategy and exploitation, technology and typology.

In area A, below a level containing historic material a Chalcolithic layer was found, with a living surface and two rounded pits, dug into the natural loess. Three layers were identified in area B, with Byzantine and Iron Age material periods but lacking Chalcolithic pottery. Flint items were found with Iron Age II period sherds in pits (Fabian, pers. comm.), one of which contained a high concentration of flint artefacts; notable among them were bladelet cores on semi-translucent flint.

A total amount of 4530 flint artefacts was collected and analysed. The flint assemblage is dominated by débitage and debris, found in almost equal quantities. Tools represent 2.8% of the flints. Four main types of raw materials were used: pebble flint, cobble brown flint mixed with limestone, translucent flint and tabular flint. Most common is wadi pebble flint, probably collected from the adjacent wadi of Nahal Be’er-Sheva. Among it, banded flint represents a distinct category, used mainly for the production of sickle-blades. Microliths and few sickle-blades and retouched blades were produced on translucent flint, while cobble flint was used mainly for the production of ad hoc tools, as scrapers, retouched flakes and notches.

Description of the flint industry 1962 artefacts were collected from both areas, mostly from area B (85% of artefacts) and analysed separately, by area. The results showed no marked differences; thus, it was assumed that most flint items were Chalcolithic. Stage 0: Acquisition – Extraction and testing of nodules

All these raw materials were probably brought to the site in their natural state and were modified within the site, as evidenced by the waste and debris collected. Tabular flint was found only on tabular scrapers and was not modified within the borders of the site, a common phenomenon in Chalcolithic sites of this area.

Most raw material originates from local wadi pebbles, recognizable by their colours (shades of grey and brown). Among them, a distinct category is banded Neogene flint. Another distinct type is a grey or brown semi-translucent flint, found in the region as small Senonian nodules, used for the production of bladelets. Other raw materials, found in limited frequencies, are breccious brown flint and flint embedded with microfossils.

The débitage group is dominated by flakes, the common blank used for the production of most tools. Cores were partially exploited, usually from one striking platform and only on a few occasions they were renewed. The presence of all products of flint knapping at its various stages hint at the local manufacture of most tools.

Stage 1: Production A common core type has one unfaceted striking platform (58% of all cores, Table 2 and Plate 20:1); 8 cores are faceted. Opposed platform cores (N=5) were mostly used to produce bladelets. Amorphous cores (N=4) were solely used for the manufacture of flakes.

Common among tools are scrapers (Table 3), usually made on primary elements, borers of various size and other ad hoc tools, as retouched flakes and notches. The relative amount of sickle-blades is somehow higher that

37

CHAPTER III –NEGEV SITES

cores for suitable blade removals, or unsuccessful attempts to produce more blades in the final stages of reduction sequences from these cores.

Flake and blade cores Most cores were used to produce flakes (N=25). However, if adding to the number of blade cores (N=11) other types, partially used to produce blades, such as flake/blade and blade/bladelet cores, the number of cores from which blades were knapped grows up to 38 and forms the majority of cores. The high amount of blades and blade-cores hint at a local production of blades, with all the reduction sequences present at the site.

Blades Blades represent ca. a fifth of débitage (Ntot= 112). Almost half (45%) are on banded flint, used for the manufacture of sickle-blades. Other raw materials are grey pebble flint and rarely brown or breccious flint.

All blade cores, except one, have a single unfacetted platform and cortex on ca. half of their circumference (Plate 20:4). All are made on banded flint. At their last stage of removals, 6 to 13 scars are discernible, most common being 6 to 8 removals on each core. Their mean length is 56 ± 7.5 mm, similar to that of blank blades (see below). It seems therefore that blade cores were not used beyond a certain length limit. A particular length was probably required for the production of blades.

Almost half of the blades have no cortex on their dorsal face, most others have less then 25% cortex (N=39), while only nine have a cortex on ca. 50% of their surface. This picture suggests that blades were removed after an initial preparation of cores and cleaning their débitage surface from cortex. Blank blades void of cortex were preferred for tool production (sickle-blades?). Although most blades are broken on their proximal end, a clear preference for using a hard hammer with a direct percussion, resulting in a flat butt, was observed (Table 28). Pointed butt appears on 9 items. The rest of the butt types are cortex covered, winged, dihedral or linear (3 items each type). Apparently, at least two different techniques were used for the manufacture of blades: most frequently used was a direct percussion with a hard hammer (which resulted in a flat, winged or cortex covered butt); the second one, using a soft hammer with a possible indirect percussion resulted in a pointed or a linear butt. In most cases, cores used for the manufacture of blades had no preparation on their striking platform.

Bladelet cores All but one of the bladelet cores are on semi-translucent flint (N=13). Their mean size are 30.3 x 27.3 x 19 mm. Bladelets were usually made from single-platform cores, though opposed platform cores appear as well (Plate 20:2,3). The amount of cortex on cores shows a different treatment to that of blade cores – most have a few cortex on their circumference, the result of their intense exploitation, that is, more bladelets were knapped from each core. This supposition is also reflected by the high number of scars discernible on their débitage surface - an average of 9 scars on each core. All but two bladelet cores were found in two pits in area B. Bladelets were probably produced for an off-site use, since only a few microliths were uncovered at the site. A similar phenomenon was noted at Tel-Sheva and on a larger scale at Site A (Roshwalb 1981).

The number of scars on the dorsal side of the blades varies between 1 to 10, most common being four scars (Table 3). This picture reflects ca. three to five stages of removals from each core. Most of the scars are negatives of blade removals from the same core. It seems therefore that blades can be defined also technologically as predetermined blades, removed from blade cores, most of them not being elongated flakes. The scar patterns (Table 28) show a preference for unipolar (N=65) or bipolar blade cores (N=15). The remaining scar patterns are in insignificant numbers.

Core-choppers Five artefacts are core-choppers (Plate 20:5): three endchoppers (one with battering signs) and 2 side-choppers. Core Trimming Elements (CTE)

Seventeen blades are over-shot, a possible reflection for the use of a direct punch with hard hammer and low control of blow intensity; 12 blades are hinged. Almost a 1/4 of the blades have a natural back, possible evidence that the last stages of cleaning the débitage surfaces of cores (at their edges) were with blade removals.

Observations on CTE showed a clear dichotomy in the proportions of their types, related aim of cores: bladelet cores, on semi-translucent flint were mostly rejuvenated by tablets, while banded flint cores (used for blade production), were mostly rejuvenated by ridge-blades.

The mean size of blades are presented in Table 28. The average length is 60 mm., with values ranging from 91 mm. to 29 mm. The average width of blades is 22.6 mm.

Stage 2: Production of blanks The industry is flake oriented (ratios flakes/blades=3.3 and flake cores/blade cores =3). An additional 17 cores have both blade and flake removal scars. Most probably, these were also used for the production of blade blanks, while flake removals were the result of either shaping the

One item is of particular interest. It is a segmented blade made on banded flint. On its distal right edge, a retouched notch is discernible. The item was (intentionally) broken in the vicinity of the notch, the result being a transversal

38

CHAPTER III –NEGEV SITES

scar on the ventral face of the item. Similar items were reported by Roshwalb (1981) and from Safadi, where an atelier for the production of sickle-blades was discovered. The technique of deliberate breakage was used to control the length of sickle-blades in a similar way to the one described above in Safadi’s assemblage.

Notches and Denticulates (N=46; 28%)

Stage 3: Shaping/Retouching stage – retouching of tools

Sickle-blades are on banded flint, with one or two truncations and a unipolar back (Plate 22:1-6,10). Two pieces have gloss along one edge. Scars of an apparently deliberated method of length trimming were observed on the distal end of two items (see Safadi and Plate 22:1,6). One item differs from the others. It was made on semitranslucent brown flint, it was modified by alternate back and it is finely denticulated along its working edge (Plate 22:11). Similar items were classified as geometric sickles and attributed to historic periods (Rosen 1997:52).

Notches and denticulates dominate the tool assemblage. Most are on cortical flakes (N= 41); 3 are on bladelets. Sickle-blades (N=14; 8%)

The tool assemblage comprises 9% of all flint artefacts (N=180). Among them, dominant types are notches and denticulates (N=48). Second important are retouched flakes and retouched blades/bladelets and end-scrapers (Table 3). Most tools were made on flakes. Blades were used mainly to produce sickle blades, retouched blades and drills. Despite the relative high amount of bladelet cores on semi-translucent flint (see above), very few microliths were found. Moreover, a typical Chalcolithic tool-type, the micro-endscraper (Gilead 1984) is missing.

Retouched blades (N=11; 7%)

Scrapers (N=18; 11%)

Retouched blades are simple blades modified on their dorsal face along one edge by fine, regular or irregular retouch (Plate 22:7-9).

“Be’er-Sheva” scrapers are most common (Plate 21:2,3); one item is steeply retouched, on a thick breccious flake with ventral retouch (Plate 21:4).

Microliths (N=11; 7%) All items are on semi-translucent flint. Retouch varies from fine to semi-abrupt, on one or both edges, being either ventral or dorsal. The fact that there are more bladelet cores than microliths and the fact that bladelet cores were found clustered together in one locus (as at Tel Sheva) hint at a small-scale production of microliths in a workshop, its products being used outside the site as well.

Two tabular scrapers were found; one is on typical tabular flint while the other is on a limy grey flint. The first item is broken on its distal end and along both edges. Partial fine retouch appears on its left edge. The second item has a rectangular form, the retouch, parallel abrupt, appears on its distal end and on the right edge. Burins (N=5; 3%)

Bifacials (N=1; 1%)

Four are on natural break one on concave truncation. One burin is on a semi-translucent retouched blade.

The item, on light grey flint, is bipointed, resembling picks. Its measurements are 93 x 82 x 50 mm. The tool was shaped from two longitudinal striking platforms, with bifacial removals. One point is smoothed and polished, probably the result of a spinning action. The polish covers parts of the smoothed surface; therefore, it is reasonable to assume that the polish is consequence of a prolonged use, which could have been drilling holes in perforated stone tools. So far, this is a unique type of tool, with such a polish (Plate 23).

Borers (N=24; 14%) The borer family consists of awls on flakes and one drill on blade. The common type is an awl with the point obtained by two retouched notches (Plate 21:1). One item is on core-tablet. The raw materials used vary greatly. Among them, awls on semi-translucent, banded and pebble flint can be discerned, reflecting an opportunistic choice of raw materials and blanks.

Varia (N=6; 4%)

Retouched flakes (N=17; 10%)

The varia group includes a heavy duty side-scraper and a heavy duty denticulate, both made on breccious flint. The third is a core-scraper, made on semi-translucent flint. Two items are broken bladelets, with a semi-abrupt retouch on their distal end. One item has two parallel notches on its distal side. The sixth item is a patinated denticulated Levallois blade, broken on its proximal end.

Retouched flakes are partially modified by fine retouch; they were made on various types of flint. The possibility that some retouch is natural cannot be neglected. Truncations (N=12; 7%) The truncated pieces are mostly on flakes; 3 items are on blades and 1 on bladelet; 4 items are straight truncated, 6 obliquely and 2 have concave truncations.

39

CHAPTER III –NEGEV SITES

upper layers of the lower parts of Shiqmim were dug. This “lower village” extends along the wadi’s terrace, ca. 2 m. above the present channel. The upper part of the site, the “upper village”, is to the north of the ”lower village”, along 10 loessial hillocks, but lacking visible architecture remains on the surface, characteristic to the “lower village”. By the end of the 1st excavations phase, 2.2% of the site was dug (2100 sqm). The goals of Phase II excavations were to dig a trench, cutting the site from north to south. Given the rich archaeological layers, only one square was dug down to sterile soil.

Summary The analysis of the flint assemblage revealed no differences between the two areas; therefore all flint artefacts were treated as a homogenous assemblage. Technological and raw materials observations associate this assemblage to the Chalcolithic. Only one item, a sickle blade is attributable to the Iron Age component. Thus, even if some waste products and ad hoc tools can be associated with any industry, the lithic assemblage was attributed to the Chalcolithic. The possibility that some Chalcolithic tools were re-used during later, historic periods was posed. However, there are few tools with a double patina, or multiple tools; therefore, even if this phenomenon occurred, it had an insignificant scale.

Main architectural features Since the site is on slopes with varying inclines, any assessments concerning contemporaneity of structures and relative stratigraphy are difficult. Three main “building phases” were defined (hereafter BP I-III), with two possible earlier ones identified in a deep probe. BPI was sub-divided into BPIa, the latest occupation at the site, while BPIb is earlier than BPIa, but later than BPII, the main occupation stage. It was impossible to point out time differences between the BP’s; architectural features assigned to different BP were often aligned in a same direction, or are additions to earlier construction stages.

A concentration of waste products on semi-translucent flint was found in Locus 506. It included most components of the reduction sequences of bladelet cores and a few ad hoc tools on the same kind of raw material (N=22). This concentration type is comparable with similar ones found in Chalcolithic sites, such as site A (Roshwalb 1981) and Tel-Sheva. It seems that in the case of Mitham C, Locus 506 stored the refuse of a knapping activity, possibly from an atelier for the production of microliths. It is important to note that very few microliths were found within the area of the site. The possibility that microliths were produced for an off-site use cannot be rejected. In contrast, some of the waste products of the bladelet industry were shaped in different ad hoc tools, such as notches, denticulates and truncations.

Phase I excavations: 18 rectangular architectural units were uncovered, most being found ca. 40 cm. below surface. They were divided into two groups: large (5.5x10 m) and small rooms (2.5x5 m). Apparently, small rooms had a domestic use, as suggested by the cultural material collected there: grinders, spindle whorls, flint and pottery debris, associated with hearths. Other features are an alley, 20 m long and 2 m wide, courtyards with small installations, associated with the large rooms, mudbrick and stone benches, passageways and pits.

Another notable phenomenon is a high number of blade cores and their products. All are on banded flint, typical of Chalcolithic sickle-blades. Even though all the reduction sequence of blade production was found within the site, its end-products, sickle-blades, are very few.

Phase II excavations: two house complexes were found (a courtyard, a number of small rectangular rooms, bound by mud-bricks and cobble wall foundations and installations). In one square, two wall foundations, belonging to BPI and BPII respectively, were found superimposed over a large, crater-like feature, built of stone foundations and a mud-brick wall and of an earlier stage (BPIII). In the western profile of the bulk and east of the wall, a cache of 11 vessels was found, probably placed in a pit and sealed by a living surface. Along the wall’s base of this structure, two earlier (BPIV) bellshaped pits, were found containing cultural material.

The flint assemblage can be assigned to the Chalcolithic period, despite the unclear stratigraphic position of its components. The tool assemblage resembles Be’er-Sheva basin flint industries, with a high frequency of scrapers and a low frequency of sickle blades and bifacials. Shiqmim (115/67) The excavation of the site and its results were published by Levy (1987). Following is its summary, based also on Yorke’s Master thesis (Yorke, 1990). Location of site and history of excavation

An apparent ritual structure (“the altar”) was uncovered in area C, below walls of BPI and BPII. It consisted of a thick ash layer bounded by a circle of stones and plaster. Below it, and possibly joined thereto, a rough semicircular cobble construction was found. East of this area, a large crater was dug by the Chalcolithic inhabitants, in which they built the “altar” structure. The northern part of the area was supported by a “retaining wall”. No other architectural remains were noted, though the filling contained a high quantity of cultural material.

The site, located ca. 18 km. west of modern Be’er-Sheva, along the north bank of Nahal Be’er-Sheva (Map 2) was excavated in two phases: in 1982-1984 and in 1987-1989. Its extension was estimated by a preliminary survey to ca. 9.5 ha., according to artefacts’ dispersion on the surface. The site was divided into five areas (A-E) and excavated in a 5 x 5 m grid system. During the 1st excavations stage, a mortuary complex (Levy and Alon 1987b), and the

40

CHAPTER III –NEGEV SITES

Phase II: chips (N=6060) and chunks (N=10268) are omnipresent at the site.

Remains of underground structures were unearthed: a tunnels complex, subterranean linked rooms, bell-shaped pits in some of their floors, buttressing walls, ceilings and walls; 12 subterranean rooms were found, interconnected by 7 tunnels. Some collapsed in antiquity, others were filled by water activity and some were used as dumping places. After a prolonged period these structure went out of their original use, as suggested by the accumulation of stratified silts, separating between floors and the dumped cultural remains. The subterranean complex was subdivided into 4 sub-phases: (a) and (b) - BPII, and (c) and (d) - BPIII.

Débitage Phase I: this group consists of plain flakes, primary flakes (covered by cortex on at least 25% of their dorsal face), blades, bladelets, cores and core trimming elements (C.T.E.). However, in the inventory list, primary flakes and plain flakes are shown together (N= 27712, 75% of all flints). Blades and bladelets, represent less than 1% of the flint assemblage. The small amount of C.T.E. (0.1%) hints at a practically inexistent application of cores rejuvenation.

The orientation of structures, along a NW-SE axis, suggests a village planning (BPII), the structures of BPI becoming additions and subtractions later on. Despite the large amount of underground structures, no clear stratigraphic relation could be defined between them and the rectilinear architecture. Apparently, underground structures consisted of deep bell-shaped pits, later used as refuse receptacles. Other underground structures are “multi-cellular”, using Perrot’s terminology.

Cores were subdivided into four sub-categories: nodular flake (N=1029), mixed flake/blade (N=60), blade (N=2) and bladelet cores (N=2). Nodular flake cores, made usually on wadi cobbles, have 1 to 3 platforms and display a high degree of hinge fractures. Their common size is less than 10 cm. Mixed cores are similar to nodular flake cores, but show some blade removals. Blade cores were also made from wadi cobbles; they were partially exploited from one striking platform. Their general shape is pyramidal, with an average length of 60 cm. Bladelet cores have a similar shape as blade cores, being made on semi-translucent flint, with one smooth striking platform. Their average length is 27 cm.

Description of the flint assemblage The Phase I flint assemblage was published by Rosen and Levy (1987). It consists of 36838 flint artefacts, among them 5939 tools. A guiding assumption that influenced its analysis was: “…small flake tools reflect general domestic functions, while specialized tool-types produced on blades, bladelets or large primary flakes, manufactured according to specific technological criteria, are the products of craft specialization and served very specific functions” (Levy and Rosen 1987:281).

Phase II: this group consists of primary elements (N=3633), flakes (N=11627), cores (N=971) and C.T.E. (N=550). Sub-types of cores are: flake (N=877), blade (N=15), bladelet (N=7) and mixed (N=72). Various reduction stages were observed on flake cores: from a few removed flakes to exhausted cores. The majority have more than one striking platform. C.T.E. associated with flake cores (N=492) are tablets. Despite proximity to raw material sources, apparently almost half the flake cores were rejuvenated (but see Levy and Rosen 1987). Blade cores have a pyramidal shape and a single striking platform. Rejuvenation of their débitage surface was performed by the removal of ridges (N=29), their higher number than blade cores suggesting a high rate of rejuvenation. Bladelet cores have a pyramidal shape and a smooth striking platform. Core trimming elements (N=29) associated with bladelet cores are flakes and platform removals. Mixed cores were used for the production of flakes and blade.

The flint assemblage collected during Phase II excavations was published by Yorke (1990). It consisted of 38146 flint artefacts: tools represent 8.3%, the rest was equally divided between débitage and debris (Table 6). A synthesis of both publications is presented below, followed by an interpretation of the flint assemblage, according to the methodology of this work. Stages 0-2: Acquisition of raw materials and Production of blanks The débitage reflects 5 technologies: large flakes (tabular scrapers), blade and bladelet tools, core tools and ad hoc tools. Each reflects different reduction sequences, distribution and use (Levy and Rosen 1987).

Choppers – Phase I:, Phase II: (N=61; 1.92%) Phase I: choppers (N=12; 0.2%)were defined as pebbles modified by the removal of two or more flakes from an edge. Their presence in houses reflect their domestic use, probably for processing plant materials, as suggested by use-wear analysis (McConaughy 1979).

Debris Phase I: the low amount of chips (N=422, 1%) is the result of the definition applied – artefacts smaller than 5 mm (Levy and Rosen 1987:283). Thus, chips and chunks, represent only 5% of the flint assemblage of the first excavation phase.

Phase II: choppers are split cobbles with a few flakes removed unifacially or bifacially, forming a rough edge, often blunted and battered.

41

CHAPTER III –NEGEV SITES

Stage 3: Shaping/Retouching stage – retouching of tools

Notches and Denticulates - Phase I: (N=718; 12%); (N=97; 2%), Phase II: (N=124; 4%); (N=79; 2%)

Scrapers – Phase I: (N=255; 4%), Phase II: (N=338; 11%)

Phase I: notches are simple or elaborate concavities with internal retouch, probably used for light scraping of hard objects like wood or bone (McConaughy 1979), being all-purpose tools. Retouch ranges from a few blows to invasive pressure retouch. Their length vary from 0.5 to 2 cm. Denticulates have repeated irregular notches, forming rough teeth on the working edge; their number vary between 2 to 7. McConaughy (1979) suggests they were used for cutting.

Phase I: scrapers were sub-divided into three: end, side (most common - N=200) and round scrapers. They were usually made on thick primary wadi cobbles flakes. The retouch is direct and sometimes scalar, at least on one of the lateral edges, and can be unilateral, bilateral or converge to meet distally. The working edge is convex. The bulb of percussion was not thinned; the butt is cortical, plain or linear. End-scrapers (N=43) have a modified distal end; this group includes also transverse scrapers. They were made on wadi pebbles, with average size of 6x5x1.5 cm. Round scrapers are made on primary wadi cobble flakes, with a round profile and outline.

Phase II: notches exhibit a wide variety of size and characteristics. Most are on flakes; 49 on chunks, 9 on primary elements and 2 on split cobbles. Denticulates are usually made on flakes, with four-five notches. Retouched blades Phase I: (N=172; 3%), Phase II: (N=150; 5%)

Phase II. Scrapers are on primary or simple flakes. Usewear analysis suggests multi-functions.

Phase I: they are modified along one or both edges by nibbling to invasive retouch. This group includes truncated blades as well, showing abrupt retouch on their snapped distal or proximal ends. When discontinuous, the retouch is located usually on the mesial part of the items. Cross-sections are either triangular or trapezoidal.

Tabular scrapers – Phase I: (N=8; 1%), Phase II: (N=36; 1%) Phase I: tabular scrapers are made on cortical flakes. Their length and width vary between 5 and 15 cm. some having a thinned bulb of percussion, with faceted butts.

Phase II: this category includes either utilized or retouched blades. Retouch may vary between slight and discontinuous, to even, direct backing or inverse bilateral.

Phase II: all but one were made on fine-grained brown cortical flakes. The absence of similar raw material cores and waste products suggests that they were brought to the site as end-products. Their morphology and shape highly varies; some show resharpening signs. Common outlines are typically fan-shaped, elongated or oval. The majority have either a completely removed or a thinned bulb of percussion. The retouch is regular, limited to the edge and along the circumference. Approximately a third of tabular scrapers were found broken.

Sickle-blades Phase I: (N=127; 2%), Phase II: (N=192; 6%) Phase I: sickle-blades are defined by two characteristics: gloss, on one edge, and shape – rectangular, though sometimes triangular. The segments are made on blades with more or less parallel straight edges and triangular or trapezoidal cross-sections. Their length is controlled by the use of truncations or by deliberate snapping that produces ends perpendicular or oblique to the tools axis. Working edge modification varies between nibbling to serration; 88 items have no gloss; therefore, they were classified by the authors as backed blades.

Borers – Phase I: (N=177; 3%), Phase II: (N=127; 4%) Phase I: borers are made on flakes, with points obtained by bilateral direct backing, or rarely by bilateral inverse retouch, located at the distal end of the blanks.

Phase II: sickle-blades are narrow, one edge is backed; the working edge has sheen along it, usually on both faces. A typical sickle-blade a straight back, a finely denticulated working edge, is straight double truncated and has a trapezoidal cross-section. Among them, 76 items (40%) lack sheen.

Phase II: borers vary in their size: 122 to 34 mm long and 67 to 17 mm width respectively. One item has two points. Retouched flakes - Phase I: (N=4253; 72%), Phase II: (N=1553; 49%) Phase I: these are miscellaneous trimmed pieces, with minor retouch or utilization damage. They are the dominant category among tools.

Retouched bladelets Phase I: (N=24; 1%), Phase II: (N=30; 1%) Phase I: all were made on semi-translucent flint, modified by direct or fine retouch.

Phase II: the largest group among tools, they have a wide variety of characteristics. Retouch is simple edge damage, to intentional retouch or backing along one edge.

42

CHAPTER III –NEGEV SITES

Phase II: they are on fine semi-translucent pink to grey flint. Retouch is fine and discontinuous, either on one or both edges. A characteristic Chalcolithic microlith, the micrograttoir, is missing at Shiqmim.

Miscellaneous trimmed pieces (N=367; 12%) These are retouched/utilized flint artefacts, either cobbles or broken tool fragments. Nine additional backed tools, defined as reaping knives, complete the group.

Prismatic (“Canaanean”) sickle-blades Phase II: (N=6; 1%)

Stage 5: Discard During Phase II excavations, over 1100 cubic metres of archaeological sediments were removed. A noticeable characteristic is the presence of a bladelet industry and its derivates in upper (BPI) context only, with possible chronological implications. Another distinct trait is the different flint artefacts distribution, related to the dug features: topsoils and fills have low densities of wastes and tools; hearths have a low density of tools, but may have a low to high density of wastes, surfaces and pits have a high density of tools and wastes, while on surfaces there are more tools than wastes. Areas of lithic production were cleaned, pits serving as disposal areas.

Phase I: these items are absent from the assemblage of this phase of excavations. Phase II: they are longer and wider than typical sickleblades, with a trapezoidal cross-section and lack backing. Retouch is dorsal and bilateral, creating a dulled and even edge, 4 items have gloss. Their raw material is finegrained brown, distinct of the pebble raw material of typical sickle-blades. All items were collected from BPI features or from topsoil. Their raw material and attributes suggest a different technology and manufacturing tradition than typical sickle-blades.

The analysis of the spatial distribution was unable to locate activity areas, structural associations or other behavioural information. A few locations revealed activities related to flint production – a small scale one in one of the subterranean rooms (the largest). However, its nature was of a much shorter time than in the stone built structures and related courtyards. The comparison of four floors of rectangular buildings revealed a moderate density of tools, except one, with a high density. This “anomaly” may be interpreted as reflecting the waste from a flint workshop, or the place was not cleaned.

Bifacials Phase I: (N=96; 2%), Phase II: (N=88; 3%) Phase I: bifacials were divided between axes (N=54), adzes (N=38), and few chisels (N=4). The working edge was produced by the removal of a transverse flake; the tip is either slightly convex or straight. Lateral edges are convex and diverge toward the working edge, leaving it wider than the tool’s body. Others have convergent sides, leaving a working edge narrower than the body. The cross-section is trapezoidal or triangular. Chisels have sub-parallel lateral edges, converging at the butt of the tool. Occasionally the working edge is polished.

A unique artefact was found in situ on the surface of the area of “the altar”. It is a complete tabular scraper, the single flint item in this area.

Phase II: they were usually made on cores, though items on flakes appeared as well. The majority were axes (N=53) or adzes (N=21); only 5 picks and 2 chisels were found. Polish is common, mainly on axes, either on the working edge, or in the mesial part, as an apparent hafting sign. Axes have a bi-convex cross-section and a high angle working edge (more than 65º), while adzes have a trapezoidal cross-section and a low-angle working edge (less than 50º).

Summary The following summary refers to the assemblage collected during Phase II excavations, as published by Yorke (1990). Since the results of the first phase of excavations are similar to those presented below, a separate summary of the assemblage collected during this stage of excavations was unnecessary.

Chisels have low edge angles. Picks were shaped by crude flaking, and have a trihedral cross-section and bifacial retouch. Two picks exhibit blunting and damage at the distal tip, indicating a heavy use against hard material. Seven items are non-identifiable, being too fragmented, or are roughouts.

The flint industry is flake oriented, as reflected by the predominance of flakes, flake cores and tools on flakes. Five technologies are evident: bladelets, blades, tabular scrapers, core-tools (bifacials) and flake tools. Most tools were locally produced, used and discarded, virtually in all contexts of the site. The presence of a specialized production of sickle-blades, bladelets and bifacials may reflect craft specialization. The absence of tabular flint and bifacial débitages suggest that tabular scrapers and bifacials were quarried and produced off-site and brought in as end-products. It is important to note that both types of tools were resharpened within the site. The absence of micrograttoirs is noticeable. Other tool-types appear in similar relative amounts to Safadi and Abu Matar.

Varia (N=30; 1%) The characteristic of this group is their possible multifunctional purpose. Six are two-pointed borers, two are borers on notched flakes, four are scraper-like, on denticulated flakes, two are scrapers with a point, and one is a scraper on a notched flake. The reminders are retouched flakes, with notch/notches on the other edge.

43

CHAPTER III –NEGEV SITES

The analysis of the spatial distribution of flint artefacts revealed several differences between excavated features. Underground structures are poor in lithic artefacts, both waste and tools. The high amount of pottery fragments suggests a storage function for these structures, and not daily-domestic activities, or flint tool production. Whereas both rooms and courtyards show proportionally high density of tools and waste products, compared to other features, the lower débitage density in rooms suggest that they were periodically cleaned, or tools were produced and used elsewhere, or both. Courtyards may have served as occasional places for flint tool production.

technological grounds, these items can be securely placed within the retouched blade group of the Chalcolithic assemblage.

Re-classification of the flint tool assemblage

Finally, the presence of a tabular scraper in an apparent ritual-symbolic context is noticeable, highlighting the other possible nature of this tool-type.

The presence of a concentration of flint tools and waste on the floor of one of the structures may represent the remains of a small-scale workshop, phenomenon common at other sites, as Safadi or Tel Sheva. The absence of micrograttoirs may not be related to digging methods and collection, but reflect a real case, comparable to that of Safadi.

During the analysis of the flint tool assemblage from both excavation phases, Levy and Rosen (1987), and Yorke (1990), adopted a different typological approach than the one followed in this work. Therefore, several factors were considered for a re-evaluation of the typological variability of the tool assemblage. Thus, miscellaneous trimmed pieces and varia were re-classified according to their supposedly last aim. For example, artefacts described as borers on notched flakes were classified as borers, scrapers with a retouched point were classified as borers, and retouched flakes with notches, were grouped together with the rest of the retouched flakes. Backed blades were grouped together with sickle-blades, given their similar morphological characteristics. All scrapers were grouped together, excluding tabular scrapers. The amount of retouched flakes were estimated according to the average of the relative amounts of retouched flakes in various Chalcolithic sites of the same geographical area and cultural affinity, such as Tel-Sheva, Safadi and Abu Matar. Finally, the average of the relative amounts of tools from the two phases of excavation were calculated, for each tool-type. Thus, more than two-thirds of the original amount of artefacts classified as tools was rejected as such.

The lithic assemblage of Shiqmim is identical in typotechnological terms to those of other sites in the area, such as Safadi, Abu Matar and Tel Sheva, reflecting a similar socio-economic organisation and cultural affinity. Ramot Nof (ca. 131.7/76.5) The site was excavated by Nachshoni as a salvage excavation (Nachshoni et al. 2002). Chalcolithic remains were found below a Byzantine building, being severely disturbed by this historic settlement. Description of the site The Chalcolithic features exposed are 12 pits, with similar shapes of 1-1.8 m in diameter and 0.5-1.8 m. depth. Noticeable among them is L511, which contained a high concentration of flint artefacts (most of the bladelet cores) and a limestone violin - shaped figurine. Fragmented remains of a living surface were exposed, on which 2 restored vessels were found. The presence of mud-bricks in the filling of pits suggest that built structures existed, but were removed by historic human activity in the area. A calibrated date of 4540 B.C. was obtained from two charcoal samples.

Several points should be highlighted: The presence of so-called proto-canaanean blades (see above) is problematic, since it is a single manifestation in the lithic repertoire of the region. A possible explanation for their appearance is their provenance: since most were found close to the modern surface, they may be intrusions from an EB occupation. Another possibility is that these items are integral of the Chalcolithic assemblage; in this case, they may be classified as retouched blades/sickleblades, their technological attributes being similar to other retouched blades, without any of the characteristics of typical canannean blades, as defined by Rosen (1997).

Description of the flint industry The lithic assemblage consists of 622 artefacts; among them, débitage form 2/3 of the assemblage, and tools 6%, the remaining being chips and chunks (Table 3). Stage 0: Acquisition – Extraction and testing of nodules The common raw material is pebble brown flint, of Eocene origin, found in the vicinity of the site. Additional raw material is semi-translucent flint, used exclusively for a local production of bladelets. Banded pebble flint and tabular flint were found exclusively on tools (sickleblades and tabular scrapers, respectively), suggesting a probable off-site production of these tools.

The C14 dates of the site may also contribute to this issues – since most of them place the site at an early stage of the Chalcolithic period, there is a long gap to the EB, and thus a discontinuity which is hard to explain in terms of cultural material continuity. Therefore, on typo-

44

CHAPTER III –NEGEV SITES

Stage 1: Production

RAMOT 3 (131.65/76)

Among the 37 cores collected, seven were used for the production of bladelets, while all others were used to produce flakes. Apparently, the blades collected (10% of débitage) are in fact elongated flakes, on technological grounds. Flake cores were only partially exploited, their potential débitage surface remained covered by cortex in most cases. Their high diversity of shapes and exploitation modes (orientation and intensity of blow) suggest a local production, within each household, of most of the flint tools. Bladelet cores are small, their size never exceeding 30 mm, with no or little remains of cortex on their circumference. The general impression is that they were exhausted. No blade cores were found. Therefore, the relatively high amount of sickle-blades collected, most on banded flint, were probably manufactured outside the excavation area. The possibility that they were off-site produced cannot be rejected.

Introduction The site (Fabian et al. 2004) is in the Northern Negev, at the foothills of Lahav Mountains, a transitional zone between the hilly area in the north and the Be’er-Sheva basin in the south (Map 2), on a gentle slope of a wadi terrace, facing the north-south axis. Its location was probably influenced by proximity to water sources or large agricultural fields, as observed in other Chalcolithic sites of the Negev (Gilead 1988). The closest Chalcolithic site, Ramot Nof, was excavated a few hundred metres to the south (Nahshoni et al. 2002). Major sites are further south, along Nahal Be’er-Sheva. Horvat Hor (Govrin 1987) and Meitar (Garfinkel and Hermon, in prep.) are other small sites north of Ramot 3. The site was discovered by a survey, during which a concentration of flint artefacts and pottery fragments was noted. Since the scattered artefacts covered a large area, it was decided to delimit the major concentration of artefacts, which roughly corresponds its half (henceforth the “site”) and to perform a systematic collection of all cultural material remains. An additional collection was performed from the vicinity of the site, in order to evaluate its nature and to collect a sufficient assemblage for analysis.

Stage 3: Shaping/Retouching stage – retouching of tools Sickle-blades (28%) dominate the tool assemblage (N=39), followed by notches and denticulated pieces (26%) and scrapers (21%). The sickle-blades are made on banded flint; all but two are broken. They are backed and occasionally the working edge is denticulated (on two items, which display also gloss along the working edge). All other tools, except microliths (a micro-grattoir and a retouched bladelet), are made on crude flakes, usually preserving cortex on their dorsal face. A polished axe and a broken tabular scraper complete the tool assemblage.

Excavation method and collection of artefacts A grid of 5x5m, north oriented, was laid over a surface of 50x40 m., from which artefacts were systematically collected. Excavations in several spots showed that the sediments are void of cultural remains, all items being on the surface, remains of a relatively ephemeral occupation of the area.

Summary The importance of the site described above is highlighted by several points:

Description of the flint assemblage

• It is apparently securely radiocarbon dated, from an otherwise poorly dated cultural entity. • The ceramic assemblage is typical to the Besorian phase, as defined by Gilead and Alon (1988) (cf. Nachshoni et al. in press). • The geographic position of the site extends the location of Besorian sites to other regions in Northern Negev and thus turning it into an apparently well defined cultural entity in time and space, and not a regional manifestation, limited to the western Negev.

Even though the flint assemblage is a surface collection, most artefacts are in mint condition, without signs of alteration of post-depositional processes. Moreover, the assemblage is homogenous, reflecting probably a single, short habitation phase. A number of 5913 flint items were analysed, including 83 uncoordinated pieces. more than half are debris (57% of the assemblage) and débitage (40 %), the remaining being tools. Apparently, most tools were produced and used on-site. The presence of 11 hammerstones supports this supposition.

Several aspects distinct the lithic assemblage from other Chalcolithic ones of the area, among them the relative amounts of different tool-types (especially sickle-blades) and the relatively high amount of bladelet cores.

The flint industry is flake oriented (ratio flakes/blades = 28), flakes being the most commonly knapped item from cores and the most common blank for tool production. The ratio primary elements/cores is 3:1; the small amount of primary elements suggests that cores were cleaned from cortex outside the borders of the site.

45

CHAPTER III –NEGEV SITES

The core measurements showed no marked differences between them, related to their number of platforms. The index values span from 30-70 for both single and two platform cores. Three single platform cores are larger than the mean, with values higher than 70. This large size probably reflects the size of nodules from which they were shaped.

Apparently, a few tools were produced from each core (ratio tools/cores = 3:1), while a large amount of waste was knapped in order to obtain suitable blanks (the ratio (primary elements + flakes)/cores is 23). The implication of these results is that the flint industry was not oriented toward optimal exploitation of cores. This is supported by observations on cores, (not exhausted) and by the few core trimming elements. This strategy was probably influenced by the low quality of the raw material.

There is a tendency for a more elongated shape as the core has more platforms. More than half of singleplatform cores are wider than longer. This picture changes on two-platform cores, where only a third are wider than longer and still more on three platform-cores, all being longer than wider. No relation was observed between shape of cores and their aim. Apparently, a similar reduction scheme was applied for the production of all blanks, from chunks wider than higher (related to position of striking platform). Thus, a more extensive exploitation, from more than one striking platform, resulted in a more elongated core, last reduction sequences narrowing the cores from additional striking platforms, located along the circumference of the chunk.

The availability of raw materials and the low investment in the manufacture of tools dictated the nature of the flint industry, characterized as generalized and nonstandardized, oriented towards the production of ad hoc tools, by several production schemes. Stage 0: Acquisition – Extraction and testing of nodules The dominant raw material type, for most tools, is of Eocene strata. Its characteristics are a light brown colour, mixed with chalk, and thus of bad quality for knapping. It can be found in the surroundings of the site as nodules of various sizes, remains of the bedrock’s disintegration. Other sources, used in a limited scale, are banded wadi pebbles and semi-translucent flints (three and one core, respectively). The exploitation of raw materials shows that most important for Chalcolithic knappers was vicinity of raw materials, and not their quality. They preferred to adapt their techniques to this restraint, even though pebble flint was available in the adjacent wadis, only a few minutes' walking distance.

Cores were mainly used to knapp flakes; only few blades and bladelets were found. Blanks were knapped from cores with single or multiple platforms, discoidal, cores on flakes or core-choppers, with a similar method, using a hard hammer on a non-prepared striking platform. This diversity was dictated by a low flint quality, which required a flexible reduction scheme, changeable according to chunks and their quality. Even though a higher quality flint can be found a few km. farther, Ramot 3 knappers preferred to use a local flint, even if this caused them to continuously change reduction methods and to remove a high amount of waste before producing suitable blanks.

Stage 1: Production 74 cores were found. Single platform ones dominate, (1/2 of all); five are on flake. Two platform cores and discoidal cores appear in similar amounts (15 % each). Remaining types are three platform cores (7%) and corechoppers (1%). The abundance of raw materials and its poor quality forced the Chalcolithic knappers to prepare a large number of cores and at the same time to limit their exploitation, usually from one striking platform.

Stage 2: Production of blanks Flakes Flakes (1/4 of all flints, Table 22) are mostly made on local raw material. A few flakes on banded flint hint for a local exploitation of this type of flint, probably for a very small scale production of sickle-blades.

The dominant removal is the flake – almost 3/4 of cores have only flake scars on their débitage surface. This is also well reflected in the large amount of blank flakes collected (Table 13) and in the tool assemblage (see below). 20% of cores have blades and flakes or bladelets and flakes removals. Only 6% of cores were exclusively used for the production of blades, and 1% of bladelets.

Flakes have a rectangular shape (57 ± 13; 44 ± 9; 15 ± 4 mm.), usually with 4 scars on the dorsal face and a unipolar pattern. Most have no, or a few, cortex on their dorsal face. Only a few items are hinged (N=8). The common butt type is flat (N=88). A typical flake can be described as having a square shape, with a flat butt, no cortex and 4 scars arranged in a unipolar pattern.

Almost half of cores have 25% of their surface covered by cortex, 1/3 have no cortex and 20% of cores are half covered by cortex. Only a few cores have a débitage surface almost completely covered by cortex (cores on flint mixed with chalk). This low amount of cortex is related to the low quality of the flint, which required a prolonged preparation of chunks, until desired blanks were knapped.

Blades 55 blades were collected and analysed (Table 28). Most were produced on the local Eocene brown flint. Occasionally, banded pebble flint and semi-translucent flint were used as well. Since cores on these types of raw

46

CHAPTER III –NEGEV SITES

materials were also found within the site, it is assumed that most blades were produced on-site. More than half have no cortex on their dorsal face; 1/3 has some cortex and 8% of the group have half of their dorsal face cortex covered. According to this picture, blades were usually knapped after the cleaning from cortex of the cores' débitage surface.

Stage 3: Shaping/Retouching stage – retouching of tools Most tools are on local raw material, except tabular scrapers and at least some of the sickle blades. The high amount of cores suggests that most tools were produced on-site. Common among tools are retouched flakes (Table 3). Sickle blades are indicative of small scale agricultural activities.

The ratio length/width shows that most blades are at the limit of their definition as such. Only three items are more than three times longer than wider.

Scrapers (N=35; 24%) Scrapers (¼ of all tools, Table 3) are of the “Be’erSheva” type. Their high amount does not necessarily reflect an intensive activity performed by these tools, but rather a higher than other tools rate of production, use and discard. They do not require preliminary preparation of cores or blanks, and from personal experiments, the time needed for the preparation of a scraper takes no more than a few minutes.

The common butt types are punctiform (19 items) or flat (15 items); an intentional scheme of blade production, resulting in punctiform butt items, cannot be neglected. Most blades were knapped from unipolar cores; centripetal and bipolar cores were seldom used for blade production. This picture corresponds to the noted core types. Half of the blades have 1 to 3 scars, 2/3 have 4 to 6 scars and only 13% of the blades have more than seven scars on their dorsal face. Apparently, blades were mainly produced at first stages of removals, after the cleaning of cortex from the débitage surface of cores.

Tabular Scrapers (N=8; 6%) A relatively high amount of tabular scrapers was collected, but only a few items are complete. One, made on blade, has a shape of an elongated rectangular. Its size are 152 x 41 x 12.5 mm. The item was modified by scaled retouch, covering most of its dorsal face, extending along its circumference. It was removed from a core with a faceted striking platform. Its unique shape emphasises the high variability of tabular scrapers morphological variability.

In order to evaluate whether the differences between blades with a flat butt and blades with punctiform butt reflect technological characteristics, they were analysed separately, according to their butt types. Punctiform butt blades are shorter, thinner and with a higher ratio of length/width than flat butt blades (Table 18). According to these measurements, punctiform butt blades are apparently more “blady” than flat butt blades, being more elongated and with a more “delicate” shape.

The second item is an elliptic tabular scraper. Its size are 103 x 77 x 16 mm. The item was modified by scalar retouch along all its edges, on its ventral face. The dorsal face was left unretouched. A third item is of particular interest, being modified by invasive retouch in the shape of bifacial knives (cf. Rosen 1997:82). Its general shape is of a large lunate. The remaining items are too fragmentary to estimate their shape. Most resemble typical tabular scrapers, thin, with a parallel retouch along their edges.

The dominant scar pattern on flat butt blades is unipolar, appearing on 3/4 of the items. Centripetal scar pattern was observed on ca. 20% of flat butt blades. Only a small amount of flat butt blades was produced from bipolar cores. Punctiform butt blades display a slightly different picture. More blades of this group were produced from bipolar cores (20%) and centripetal cores (27%). The remaining blades (ca. half of them) have a unipolar scar pattern on their dorsal face.

The variability of raw materials used for the production of tabular scrapers suggests different sources of raw materials, and thus, different origins (workshops for the production of tabular scrapers). Since no evidence for their production was found on-site, a different economical mechanism was probably responsible for their import. It is important to note that this mechanism, noted at larger sites, such as Safadi or Shiqmim, is valid also for the smaller sites, such as Ramot 3 or Meitar (see below).

Generally, flat butt blades have more cortex on their dorsal face than punctiform butt blades. The implication of this observation is that punctiform butt blades were produced in later stages of removals than flat butt blades. These observations suggest a different production scheme of punctiform butt blades: they are thinner and longer than flat butt blades and were produced in a higher frequency from bipolar cores and at later stages of removals. Some flat butt blades are technologically similar to flakes, probably being elongated flakes.

Burins (N=1; 1%) Only one burin on flake, made on natural brake, was found. Burins are rare in Chalcolithic assemblages; hence, the presence of only one item is not surprising.

47

CHAPTER III –NEGEV SITES

Borers (N=19; 14%)

Bifacials (N=3; 2%)

The borer family comprises only awls, all on flakes (Table 3). The absence of drills and micro-drills may be related to their specific uses, (such as beads preparations or wood working), activities not testified at Ramot 3. Awls are on the local raw material. Their point was shaped by the abrupt retouch of two notches, usually at the distal end of items, usually extending over a 1/3 of the overall length of items. The production of awls does not follow a rigid production scheme, since the size of blanks, the orientation of the points related to the striking axis of items and their position appear to be random.

All three bifacials are 3 adzes. One, made on a whitish flint, uncommon in the area, is extremely large: 150 x 72 x 45 mm. It was shaped from two striking platforms located along its edges by bifacial large flake removals. A third striking platform is along the working edge, shaped by one longitudinal ventral flake removal and a few longitudinal blade removals on its dorsal face. The second adze resembles Chalcolithic ones, regarding aspects such as raw material morphology and technology. It was partially bifacially polished near the working edge and shaped from a pebble flint, by large bifacial flake removals from two striking platforms along the edges and two additional ones along both ends. The ends were shaped by bifacial longitudinal removals. The third item is similar to the previous one, but lacks polish.

Notches (N=18; 13%) All notches are made on flakes. Their mode of manufacture and general morphology suggest a minimum investment in their manufacture.

Varia (N=4; 1%)

Retouched Flakes (N=46; 32%) Retouched flakes represent 1/3 of the tools, being the most common tool-type. Generally, the retouch is fine, dorsal and partial, along one edge of the items.

The group consists of one tang of a Neolithic arrowhead, one small blade with partial fine retouch at the mesial part of both edges (transversal arrowhead?) and two flakes, retouched along edges and ends.

Retouched Blades (N=5; 4%)

Stage 5: Discard

All have a partial fine retouch along one or two edges.

In order to evaluate whether post-deposition processes affected the distribution of artefacts, the relative height of each square was measured. The topography of the area displays a difference of 6 metres (at max. height) in elevation, between the northwest and the southeast corners. Generally, the east part is higher than others; its inclination is ca. 6o. The highest spot is located more or less in the middle of the site; most flint artefacts were collected from the eastern parts of the site.

Sickle-Blades (N=5; 4%) Five items bearing sickle gloss were found. However, their relative frequency is similar to other sites of the Be’er-Sheva valley (cf. Table 3). All are on pebble banded flint, typical for the production of Chalcolithic sickle-blades in this region (see above). Probably most items were produced off-site, in various places, since each one differs from the other regarding morphological and technological aspects. Only one sickle-blade resembles Be’er-Sheva basin sickle-blades, being backed and truncated. Two others are backed, but they are wider and their back is curved. The remaining sickle-blades are on fine retouched blades, with partial backing.

The distribution of artefacts was not influenced by postdepositional factors, since various concentrations, each of other flint types, were located, and items were not concentrated on the lowest parts of the site, or gradually arranged according to shape and size, along the slope. Flint artefacts were not discarded into dumping zones. Several episodes of knapping activities were identified: cores were cleaned from cortex in one area, then processed at several metres distance, while tools were discarded in different locations. Since no architectural remains were found, it is unknown whether these concentrations were connected to habitations areas or they reflect knapping activity zones.

The low number of sickle-blades can be interpreted in different modes. Agriculture did not play an important role in the economy of Ramot’s inhabitants. The number of inhabitants was small, comprising only a few families, occupying this area for a short period of time. Sickleblades that went out of use were discarded in the field. The life span of sickle-blades was longer than that of other tools, therefore their production was less intensive. Usable sickle-blades were taken out from the site, when its inhabitants left it. Probably most of these potential factors influenced to a certain degree the presence / absence of sickle-blades. It is important to note that the relative amount of sickle-blades found at Ramot is similar to that found at larger sites in the Be’er-Sheva basin area, such as Safadi, Abu Matar, Shiqmim or Tel-Sheva.

Summary Most flint artefacts were on-site produced from the locally available raw material, a brown and breccious Eocene flint, mixed with chalk and therefore of low quality. The flint industry is flake oriented, with a few deliberately produced blades. Several removal schemes were applied, dictated by the size and shape of nodules.

48

CHAPTER III –NEGEV SITES

Schick (1978). It consists of 47 waste products, mostly flakes, and 14 tools on various raw materials and techniques, resembling Be’er-Sheva basin assemblages. Among tools, worth mentioning are one thin tabular scraper, modified by circumferential fine retouch, two typical scrapers on cortical pebble flakes, one adze and one sickle-blade, bi-truncated and backed. Even though it is not mentioned in the report, the very fragmentary nature of the assemblage may have been biased by the excavation and collection method, not focusing on the systematic exposure of the Chalcolithic strata in general and the sporadic collection of flint artefacts in particular.

The large number of cores was probably the result of low quality of raw material, which required a long knapping sequence, from more than one striking platform, until suitable blanks were obtained. The diversity of modes of production was probably influenced by this factor as well. Only a few tools were produced from each core. The spatial distribution of flint artefacts shows several concentrations, according to different stages of production. Flint tools were concentrated in two other areas, which probably reflect the place where they were used and later discarded. The mint condition of artefacts and the presence of various concentrations, each containing different flint types, reflects the low influence of post depositional processes.

Meitar (143.83/82.2) Introduction

Within the site and its nearby vicinity, no traces of any kind of features, installations or architecture could be discerned. Therefore, it might be concluded that the occupation of the site was short-term, possibly seasonal, in ephemeral structures, such as tents or wind shelters.

The site, located north of modern Be’er-Sheva (Map 2) was excavated during one season of three weeks, in April 1997 (Garfinkel and Hermon, in prep.). The main goal of the excavation, extending on an area of ca. 2 dunams, was to determine the boundary and the intensity of the archaeological occupation in an area planned for urban development. The excavation unearthed a Chalcolithic occupation, consisting mostly of pits of various size and shapes, and fragments of stone built walls and installations on living surfaces.

The flint industry is comparable, in typo-technological terms, to other industries of the area, found at larger sites. Moreover, the relative amount of tool-types is similar to tool assemblages found in these Chalcolithic sites of the Be’er-Sheva basin. Accordingly, the tasks performed within the site were apparently similar to those of the larger sites. These tools may represent the range of types representative of a limited number of households.

Description of main features The excavation area was delimited by trenches dug by a mechanical tool, which restricted an area of ca. 75 x 65 m, where remains of human occupation were discerned and subsequently defined as “the site”.

The presence of tabular scrapers at the site may reflect the low economic value of these items, even though they were apparently achieved by some kind of exchange network. A further conclusion might be that economical differences between households were minimal, each product being accessible to all, and thus social status based on economic inequalities were minimal. Moreover, the economic structure was independent to the nature of the site, large ones, as Safadi, Abu Matar or Tel Sheva (described above), having a basically similar economy with smaller sites, as Meitar (see below) or Ramot 3.

The main feature of the site is the pit; 3 types were recognized: bell-shaped (N=4), pits with a row of stones or mud-bricks at their opening (N=3) and rounded, of 50 cm. in diameter and depth (16 pits). At the entrance of one of the bell-shaped pits, two large grinding slabs were uncovered, apparently blocking its entrance. The depth of the pit is unknown. Due to safety considerations, the excavation stopped at a depth of one metre below surface. In one spot, the NE corner of the site, remains of a “living surface” were identified at 10 cm below surface. On this layer, pottery sherds and stone vessels were found “in situ”. Probably the pits found in the vicinity were dug from this layer. Adjacent to it, several stone built installations were uncovered, as well as the blocked entrance to a natural cave, which was not explored.

The ceramic assemblage resembles “Besorian” ones (Fabian et al. 2004). Thus, together with the nearby site of Ramot Nof, Ramot 3 may represent occupation at the early stages of the Chalcolithic period. It should be noted that at both sites, the small lithic assemblages displayed no marked differences with Be’er-Sheva Chalcolithic sites, such as Safadi or Shiqmim.

Apparently, the upper part of the site was eroded. However, since some pits were certainly excavated from their top, as suggested by the mud-bricks or stones found at their opening, the possibility that this picture represented the ancient occupation at the site cannot be rejected. The density of pits should be noted –sometimes within a few metres from each other. However, large areas between the concentrations of pits was void of cultural remains.

Tel Arad (162/75) Chalcolithic remains were found in different isolated locations at this large site, in layer V (the earliest occupation). Pits of various shapes and sizes, as well as concentrations of finds on the bedrock, are manifestations of the Chalcolithic occupation at the site, of a nature which still remains obscure, given the limited excavated area. The flint assemblage was briefly described by

49

CHAPTER III –NEGEV SITES

2. most of the bladelet products were used and discarded elsewhere, outside the borders of the excavated areas of the site, or 3. bladelet products were taken off site.

Description of the flint assemblage 1727 flint artefacts, among them 73 tools were found. Since during a previous reconnaissance survey on the area only Chalcolithic flint implements were observed, artefacts collected from both surface and the excavated features were treated as a homogenous assemblage.

The first possibility can be rejected because a large many chips were collected and the sieved sediments did not yield different results than the unsieved ones. Since a large area of the site remained unexcavated, options two and three remain available. A similar phenomenon (large amount of bladelet cores and few end-products) was noted at Abu Matar and Tel Sheva as well (see above).

Stage 0: Acquisition – Extraction and testing of nodules The preferred raw material was a breccious flint, found in the area as large boulders, probably the results of the bedrock disintegration, of Meishash formation. The Chalcolithic knappers exploited these boulders exposed in the vicinity of their habitation area, as exemplified by the many broken fragments still discernible on the surface of the site. Large boulders were probably broken, in order to obtain chunks to be later modified into cores. A brief survey of the boulders on the surface of the site, showing scars of removals, relieved no planned strategy for the extraction of large flakes, the “block on block” technique probably being applied (Inizan et al. 1991). Another type is pebble flint. Its closest source is a tributary wadi of Nahal Hebron, circa 50 m. west of the site. Pebbles were brought to the site, where they were modified into cores. The same strategy was employed for semi-translucent flint, found on the surrounding hills as small nodules.

The initial shaping of nodules involved the “block on block” technique; large boulders were broken in order to obtain chunks or flakes, further modified into cores, with flat striking platforms. Different knapping sequences occurred, reflected by the number and sizes of striking platforms. Probably the low quality of the flint imposed the knapping of a large amount of waste, from which only a small amount was suitable for tool production. Stage 2: Production of blanks The flint assemblage is dominated by debris (ca. half of all waste, Table 13). The possibility that some debris (especially chunks) is natural, or part of preliminary exploitation of the large boulders, should not be ignored. Among débitage, flakes dominate (Table 13), as showed by the ratio [flakes/(blades + bladelets) = 15], the low number of blade cores and the low number of blade tools. Primary elements are 5% of all flints; the ratio primary elements / cores (value = 1.5) suggests that few removals were sufficient to prepare a débitage surface. Since some cores were made on chunks obtained from large boulders, and thus lacking cortex on their circumference, a low number of primary elements was expected.

Stage 1: Production Cores All but one of the cores are on breccious flint or on wadi pebble flint, of different shades of grey (N=58). The exception is a blade core on banded flint, probably used for the production of sickle-blades. Most cores were used for flake production (Table 2). Flakes were knapped from flake removals (N= 32), from cores on which blade and flake scars are discernible (N=10) and from cores with bladelets and flakes scars (N=5, Plate 26:1). Probably only the first group of cores was intentionally used for the production of flake blanks. The flake scars discernible on the other two groups of cores may result from reshaping or renewing their débitage surface. Among the first group of cores (used for flake removals only), two striking platform cores dominate (N=21); 13 are bipolar and 8 have opposed platforms. The mean size of flake cores are 46 ± 12 mm length and 45 ± 11 mm. width. Most cores have a few cortex on their circumference.

Flakes 100 flakes were analysed. The preferred raw material for their production is breccious flint; others are on wadi pebble flint or semi-translucent flint. Ca. half the flakes are cortex covered; 40% are void of cortex. Apparently, their production process was performed within the site. The predominant butt is flat, showing that flakes were obtained by direct blows, probably with a hard hammer, from cores without preparation of a striking platform. The number of scars varies between 1 to 11 (Table 27), most common being 3 scars on each flake. This number coincides with the supposition that from each core at least three stages of removals were performed.

Bladelet cores (ca. ¼ of all cores, N=13), are exclusively on semi-translucent flint. Although the number of bladelet blanks and microliths is very small (N=8), bladelet cores were exhausted (Plate 18:2). There are several explanations for the low number of bladelets and microliths:

The scar patterns on the dorsal face of flakes show that in ca. half of the cases flakes were knapped off from singleplatform cores; a 1/3 have a multi-directional pattern, being knapped from amorphous cores, while the rest have a bipolar scar pattern, being removed from bipolar cores.

1. Since not all sediments were sieved, there is a possibility of missing some of the smaller pieces,

50

CHAPTER III –NEGEV SITES

Blades

Notches and Denticulates (N=13; 18%)

Blades (N=40, 2% of all flints) are commonly on pebble flint. Five items are on banded flint. Blades were more often produced on pebble flint, rather than on breccious flint, reflecting a deliberate choice for this type of raw material, due to its quality, suitable for their manufacture.

Although notches and denticulates form the next group in importance, their non-standardized nature renders any further description useless.

Half the blades have a unipolar scar pattern, a third bipolar; multidirectional pattern is rare (Table 28). Three butt types were noted, all in similar amounts: flat, winged and pointed. Flat butt blades were knapped from unipolar cores (2/3 of them have a unipolar pattern). Winged or pointed butt blades were produced from unipolar or bipolar cores. Apparently, blades were mainly knapped using a soft hammer and applying a predetermined production method, from unipolar cores.

The microlith is a micro-endscraper on semi-translucent flint (Plate 26:4).

Stage 3: Shaping/Retouching stage – retouching of tools

Varia (N=3; 4%)

Microliths (N=1; 2%)

Bifacials (N=4; 5%) Two adzes (Plate 26:6), one chisel and one roughout were found. The first three were discarded after they went out of use, as indicated by their damaged working edges. The roughout testifies for a local production of bifacials.

The varia group consists of: one item with 3 longitudinal ventral removals, resembling a pièce esquilles (Bar-Yosef 1970), a core-scraper and a broken blade, which resembles an arrowhead with a tang.

The flint tool inventory (Table 3) is one of the few described from this area. Scrapers (N=25; 34%)

Stage 5: Discard

Scrapers, of the “Be’er-Sheva” type (Plate 26:5) dominate; among them rounded, with parallel semiabrupt retouch covering most of their circumference are common. Their blank is a cortical wadi pebble flake. Others are steep retouched; a few are on blades.

Flint artefacts were collected from three types of features: ca. half are from the surface or subsurface (first 30 cm.), a third was collected from pits of different size, and the rest from various architectonic features. The surface and sub-surface items were concentrated near the entrance to the cave, which was blocked with large boulders and pebbles; the flint artefacts were probably part of this filling material. The remaining flints were equally dispersed in each square, several tens in each one.

Only one tabular scraper was found, made on transversal flake; its bulb of percussion was thinned and the edges were modified by a semi-abrupt parallel retouch. Burins (N=2; 3%)

550 flint items were collected from 26 pits, ca. 20 in each. No correlation between pit size and amount of flints was noted; 4 pits were richer (ca. 50 artefacts per pit). One (L635) contained 4 cores (average is 1core per pit). It is possible that it served as a refuse pit of a flint knapping activity. Pit no. 634 contained the highest number of tools (N=4). The small number of artefacts limit the intra-site distribution analysis. Moreover, since only one living floor was found, and given the clear disposal of flint artefacts into dumping zones, the distribution of flints cannot point at activity areas.

One burin is on natural break and other on truncation. Retouched Flakes (N=9; 12%) Retouched flakes were modified by fine, partial retouch along one edge. Borers (N=9; 12%) Awls (N=7) are on flakes, with the point obtained by the retouch of two notches. The two drills found are made on blades; one is on an abraded sickle-blade, modified by ventral retouch in order to obtain the point.

Summary

Retouched Blades (N=2; 3%) and Sickle Blades (N=3; 4%)

The flint assemblage of bears all the characteristics of Chalcolithic assemblages of the Be’er-Sheva basin. This supposition is reflected in all aspects of the lithic industry, both in technological and typological terms. The flint industry is flake oriented, blades being partially used for the manufacture of sickle blades. Other tools on blades, such as drills, may be by-products of the same sickle-blade production, made on unsuccessful or resharpened pieces. Bladelet production is generally

Retouched blades have partial fine retouch along one edge. Sickle blades (N=3) are on banded flint, with straight truncations, denticulated working edge and unipolar backing (Plate 26:3). The presence of a blade core on banded flint testifies for a local production of sickle-blades.

51

CHAPTER III –NEGEV SITES

Description of main features

restricted to micro-endscrapers or partially retouched bladelets. Although a relative high number of bladelet cores were uncovered, very few bladelets and only one microlith were collected. Since the entire site was not excavated, it is possible that further excavations will uncover a higher number of bladelets. The probability that some of the bladelet products were taken off the site cannot be rejected (such as at Tel Sheva).

Apparently, the site is composed of a cluster of sub-sites, or habitation areas, easily recognizable by patches of ashy archaeological sediments, separated by sterile, yellowish sediments. Therefore, even though the extension of the archaeological remains cover a relatively wide area, any reference to Grar as a large and extensively occupied site is not in concordance with the reality, as seen in the section of the wadi, on the surface of the agricultural field and during the excavation. Observations on the natural section of the wadi and the excavation of the site revealed that the depth of the archaeological sediments are usually one metre, occasionally two to three metres, but all belonging to one period settlement. The various occupation cluster zones cannot be chronologically anchored neither can their contemporaneity be proven. However, the homogeneity of the cultural material shared between the different clusters suggests they all belong to a single period/culture settlement. Therefore, the material culture was analysed as belonging to a single cultural entity, the recording of artefacts from the various clusters being used for a comparison between the sub-sites.

The flint tool assemblage is dominated by scrapers and other flake tools, such as retouched flakes and notches. The scrapers are similar to others found in the Be’erSheva basin. The sickle blades maintain all technological and typological characteristics of Chalcolithic sickleblades of the Be’er-Sheva basin. The number of sickleblades found at Meitar (N=3) could barely fit into one sickle. Given their different provenance within the site, it is probable that each piece belonged to a different sickle. Despite the proximity of the site to open fields, it seems that agriculture did not play an important role in the economy of the inhabitants of Meitar. However, the possibility that the inhabitants took with them the sickles when they left the site, or discarded the unused ones within their agricultural fields, should be kept in mind.

The common excavated feature is the pit, being the richest in finds. An ordinary pit is a depression dug into the soil, containing accumulations of pottery sherds, bones, flint artefacts and occasionally stones, all in ashy layers. Pits have different volumes, shapes and diameters; most of them being eventually used as dumping zones, even if they might have had a different primary function, as reflected by the presence of bell-shaped pits (storage pits?) or burial pits. Remains of poorly preserved stone or mud-brick walls were excavated in various places. Few features, mainly installations made of mud-bricks or stones, complete the architectonic features of the site.

The flint assemblage described herein is one of the few Chalcolithic assemblages reported from this part of the Negev. Its close similarity with Be’er-Sheva basin flint industries reflects both cultural and economic analogies and therefore expands the area of typical Be’er-Sheva flint industries further north. The flint assemblage may represent the tool-kit and the waste of its preparation of a typical Chalcolithic household (or a minimal number of households), of a Be’er-Sheva basin origin. Grar (143.83/82/2) The site (Gilead 1995), was excavated between 1981 and 1987, during which time 5925 flint artefacts were collected from various archaeological features. The site is located on the top of a thick, silty sediment, that forms the right bank of Nahal Grar, in the northern-most section of the loess-covered area of the Negev (Map 2). The site is located in an agricultural field, and since it was found immediately below surface, probably its upper layers were disturbed by modern ploughing activities.

Description of the flint assemblage The analysis of the flint assemblage included all pieces, from all sub-sites. Several tool-types and waste products underwent an attribute analysis. An intra-site analysis of the distribution of flint artefacts was performed, related to features and areas. Stage 0: Acquisition of raw materials

Excavation and recording methods

The large majority of flint artefacts were made on wadi pebbles, mostly of Cenomen-Turon origin, collected from the wadi of Nahal Grar, located few metres from the site and originating from the dislocated bedrock of the surrounding hills and transported by water. The common colour is brown to grey, occasionally brecciated flint, found on both waste products and tools. A distinct type is a banded grey pebble, used for the production of sickleblades. It can be found in the wadis of Nahal Be’er-Sheva and Nahal Besor, in the north-eastern Negev, but not in the wadi of Nahal Grar. Two additional types of flint were recognized: a pink to grey, semi-translucent flint,

The site was divided in areas, or “sub-sites”, which grossly correspond to patches of apparently habitation areas, separated by sterile sediments, discernible on the surface. “Baskets” were defined as the smallest excavation units, while “locus” defines any human-made feature. The excavation was performed within a grid system of 5 x 5 m, covering the entire area of the site.

52

CHAPTER III –NEGEV SITES

originating in small Senonian nodules, and tabular flint, its source being probably in the Negev Highlands. Semitranslucent flint was mostly used for the production of microliths; tabular flint was exclusively used for the production of tabular scrapers.

Stage 2: Production of blanks The flint industry is oriented towards the production of flakes, as reflected by cores, débitage (Table 13) and the high number of tools on flakes. The ratio (primary elements + flakes + flake tools)/cores suggests that 15 blanks were removed from each core. Among them, 10% were modified into tools.

All raw materials, except tabular flint, were collected from the vicinity of the site and brought to the site in their natural form, being later modified within the area of the site. Tabular flint was found only on tabular (fan) scrapers, which were manufactured in an unknown location, probably close to the source of this flint.

Other production schemes, pre-determined, are of sickleblades and microliths, from deliberately chosen raw materials (banded flint and semi-translucent flint, respectively) and from blade and bladelet cores. They are similar to the ones described above, for the sites of Safadi and Abu Matar.

Stage 1: Production Most cores (90%) were used to produce flakes, others being blade and bladelet cores. Flake cores are nonstandardized, with a globular shape (similar size of length and width) and short striking platforms. Cores were exploited from multiple striking platforms, until they went exhausted, most of their débitage surface being used. The prolonged use of some cores is also reflected in the high number of core trimming elements (Table 13). No preparation of the striking platforms was noted. The decortication of pebbles used for the production of flakes took place within the site, as documented by the appearance in high numbers of primary elements.

Flakes Flakes are the most common waste type (Table 13). Almost all (90%) are made on pebble grey or brown flint; flakes on banded or semi-translucent flint probably represent removals used for shaping cores for a predetermined blade and bladelet removals, and not the deliberate preparation of blanks. Almost 3/4 of flakes were knapped with a hard hammer, as suggested by their plain, flat striking platform. Soft hammer was probably involved in the production of 12% of flakes, baring a pointed butt. The possibility that these were knapped during the preparation of blade cores cannot be neglected. The remaining flakes have a facetted striking platform. The sizes of flakes are presented in Table 27. The similar length and width hints at their square shape. This shape is best depicted by the values of the ratio length/width (mean value = 1.13). The mean length of blank flakes corresponds to the mean of the flake cores’ length.

The initial shaping of pebbles included the knapping of a few number of bifacial removals, in order to obtain a straight striking platform/cutting edge. These items resemble choppers, a phenomenon discussed above. Blade cores are too few for a detailed analysis; their importance lies in the fact that they were used probably for the production of sickle-blades, being on the same raw material, banded flint, as the sickle-blades. Their small amount, compared to the number of sickle-blades collected, suggest that either more cores of this type remained buried within the site, or some sickle-blades were brought on-site as end-products. Blade cores have one striking platform; ca. half of their circumference remained covered by cortex. They resemble Upper Palaeolithic pyramidal cores (cf. Gilead and Hermon, in press), being similar to ones found in other Chalcolithic sites of the area, such as at Besor A and the Be’er-Sheva sites (Safadi or Abu Matar).

Blades Most blades are made on pebble grey or brown flint, but in a lesser percentage than flakes (78%, opposed to 90% of flakes). The remaining blades are made on semitranslucent flint. They may represent early removal stages for the production of bladelets. The relative amount of blades with a punctiform butt is higher than on flakes (18%, against to 12% on flakes), suggesting a soft hammer, with a different technique, being involved in their production. This deliberate scheme of production is also reflected in the parallel scar pattern on the dorsal face of most blades.

Bladelet cores were made on semi-translucent flint, their products being found within the excavated area. Their small amount does not allow a detailed description supported by statistical analysis. Apparently, the high number of microliths suggests that at least some of them were brought on-site as end-products. However, the possibility that more bladelet cores are still buried within the site as a single concentration (a phenomenon observed elsewhere, see above) cannot be neglected.

Stage 3: Shaping/Retouching stage – retouching of tools Even though the flint industry is clearly oriented towards the production of flakes (see high amount of flake cores and the dominance of flakes among waste products), almost half the tool assemblage is comprised of tools made on blades (Table 3). The possibility that a certain amount among these tools was brought to the site as endproducts cannot be rejected. However, if assuming that

53

CHAPTER III –NEGEV SITES

among available blanks for a desired width. Ca. ¼ of sickle-blades are on banded flint, while only 2% of the blank blades are on this raw material type. This amount can be related to the small number of banded flint blade cores found at the site, and it supports the suggestion that most of the sickle-blades were not produced at the site.

most tools were made on-site, the investment in the blade production is evident: each third blade was modified in a tool, while only each 7th flake was retouched. Moreover, this observation indicates a “pre-planning” of blades, each blank being suitable to be modified into a tool. Thus, the production scheme of blades was efficient in terms of pre-determination. When compared to that of blades, the non-determined production scheme flakes is apparent.

Sickle-blades were usually backed (Table 15); commonly, the working edge was finely denticulated. The ends were truncated: 73% on one end, 23% on both. Most sickle blades were either broken or intentionally shortened by truncations; only 3% preserved their original size. Almost all (98%) exhibit gloss along their working edge.

Sickle-blades represent almost a ¼ of the tool assemblage, being the most common tool-type. Other types, in similar percentages of 10%, are microliths, notches and retouched flakes. All other tool-types appear in small amounts (Table 3).

Microliths (N = 70; 10%)

Scrapers (N=62; 9%)

The main characteristic trait of Chalcolithic microliths is the raw material – semi-translucent flint. Microliths were divided into two groups: retouched bladelets (N=40) and micro-endscrapers (N=30). Retouched bladelets were finely retouched along one or both edges. A possible subtype is a bladelet with fine retouch on one edge, the opposite edge being obliquely truncated, turning the end into a point.

Scrapers are simple end-scrapers with a distally retouched rounded end. Some resemble Be’er-Sheva scrapers. Side-scrapers (N=12) were modified by evasive or scalar retouch on an edge, either ventral or dorsal. The difference between the sub-types is probably not functional, but rather attributable to the original shape of the blank and a particular choice of the knapper. Tabular (fan) scrapers (N=7) are made on a flat, cortical dorsal face, and a scraper retouch around its circumference.

Bifacials (N = 44; 6%) The majority of bifacials were divided into two groups: axes (N=14) and adzes (N=27). Chisels (N=3), bifacials with pointed ends, appear in insignificant numbers. Polish occurs rarely, on the working edge. Bifacials were usually shaped from two striking platforms, located long their edges, though some adzes have a trihedral crosssection, being shaped from an additional striking platform. No picks were reported.

Borers (N = 37; 5%) Borers were subdivided into three main types: drills – a flake or a blade completely retouched along its edges, commonly with an abrupt retouch, awls – made on flakes, with one or more points, retouched on two sides to form a sharp, piercing tip and micro-borers –bladelet fragments or flakes, with a distal tip modified on both edges.

Stage 5: Discard

Retouched flakes (N = 90; 13%), Notches and Denticulates (N = 181; 26%)

The site was divided into six areas (A – E and G), according to clusters of finds discernible on the surface, separated by sterile sediments. Most flint items were collected from area E, even though area B is larger. Their mean density (calculated as the ratio number of artefacts/number of baskets) was similar in most areas, excluding E, with a much higher density, where apparently the flint production was more intensive than in other areas. However, the difference is only quantitative, since when techno-typological aspects are considered, all sub-sites share similar characteristics, with small, but worth mentioning typological differences.

Their shape is non-standardized and retouch is minimal, showing a low investment in their production. Most notches and denticulates were on flakes. The possibility that the retouch on some pieces was caused by natural agents cannot be excluded. Retouched blades (N=42; 6%) Retouched blades are simple blades with a fine, regular or irregular retouch along one of their edges. Sickle-blades (N = 166; 24%)

Grar C and E are similar in terms of flint typology, excluding denticulates (8.26% in C and 1.52% in E), even though their architectural features differ. The tool-types frequencies in other sub-sites differ from those of C and E, and from each other. These differences are apparent when sub-sites with similar architectural features are compared, such as Grar B, E and G. The variability of D and G may be influenced by their small flint sample.

Sickle blades constitute the most important tool-type, both in quantitative terms and their implication on reconstructing socio-economical aspects of the Chalcolithic inhabitants of Grar. Most have mean size of 25-35 mm. long, 8-12 mm. wide and 2.5-4.5 mm. thickness (Table 15). The homogeneity of the width distribution suggests that there was a careful choice

54

CHAPTER III –NEGEV SITES

Of particular importance is the presumed presence of a workshop for the production of tabular scrapers (at site A), a phenomenon restricted to this site. Apparently, the knappers brought to the site large chunks of tabular flint, which were further modified into tools. Another specialty of the Chalcolithic knappers of the Besor area was the production of bifacials, as testified by the presence of a high quantity of roughouts and discarded bifacials, at various production stages. So far, this is the only place where a bifacials workshop was reported.

Sickle-blades are a frequent tool-type, and in all sub-sites but Grar B (13%), they dominate the tool assemblage (25%-31%). This dichotomy cannot be easily explained, since architectural features did not influence the nature of the tool assemblage. Another difference is the high frequency of scrapers in C – 17%, ca. 3 times more than in any other sub-site, probably due to a low number of retouched flakes and the absence of retouched blades. Summary Pebbles, modified into flake core, are the most commonly used raw material . There is a preference for particular raw materials for the production of distinct tool-types: cortical pebble chunks for scrapers, tabular flint for tabular scrapers, banded flint for sickle-blades and semitranslucent flint for microliths. Important to note is the extensive use of cores, even though the main source of raw materials, the nearby wadi of Nahal Grar is a few metres from the site. Blades, insignificant in the waste products, were extensively used as blanks for most tools. Microliths are also prominent, scrapers and bifacials are present in low quantities. The few standardized blade cores, used for the production of sickle blades, suggests that most sickle-blades were made elsewhere. Worth mentioning is the absence of picks.

The picture described above should be balanced by the presence of a high amount of material culture that can be described as domestic. Thus, the main activities of the Chalcolithic inhabitants of the Besor sites concerned general aspects of a domestic, self-sufficient economy. Support for this suggestion is the architecture and various features found at these sites (similar to common Chalcolithic sites of this area) and the remains of material culture, such as flint and pottery. The products of the various workshops may have had served regional purposes, and were not subject to long-distance, ramified or high-scaled trading, but rather parts of a network of good-exchanges of a low economic importance level.

The entire flint assemblage and assemblages from subsites are heterogeneous; there are moderate to low numbers of most tool-types and similar relative amounts of waste products. This suggests that no specialized activity involving flint artefacts was performed in any area. Apparently, most tools were made by the site inhabitants, each household producing their own tools. Differences in architectural features did not correspond to differences in the relative frequency of tool-types, and vice versa – similar tool-types frequencies were found in different architectural features.

The site is located on the west bank of Nahal Besor, south of Tell Fara (Map 2). It was excavated in three different events, by Macdonald in 1932, Perrot in 1962 and Alon in 1977. The earliest excavation reports floor levels with pits, hearths and stone installations. At least seven flint activity areas were recorded, all at the same depth, among them one with tabular scrapers and long flakes (blades?), one with long flakes (blades?), one of tabular flint and a dump of hoes (related to as core-tools, celts, or bifacials). No architecture was uncovered in any of the excavations.

Site A

Description of the flint assemblage

Wadi Ghazzeh sites (ca. 101/75)

The flint industry can be described as highly specialized, having five distinct fields of specialization, different production methods and end-products. The first is a blade industry, probably for the production of sickle-blades. The second one is a bladelet industry, oriented towards the production of microliths. The third produced various flake tools, as scrapers and retouched flakes. The fourth one produced tabular flint scrapers, while the last one was directed towards the production of bifacial tools.

The following paragraphs summarize the results of flint assemblages analysis from sites in the Nahal Besor area, discussed in a doctorate thesis (Roshwalb 1981). The data analysed in this dissertation originates from early excavations (Macdonald 1932); several sites were reexcavated during the sixties (Perrot 1962b, Alon 1977). These assemblages were selectively collected: probably only a sample of tools, and not all types, and only a few débitage items; hence, quantitative comparisons between sites are meaningless. Moreover, only general characteristics of the flint industry can be assessed, given the fact that most débitage was not collected.

The comparison of the amounts of the final products of each industry reveals that mostly bifacials (36%) and blade tools (23%) were discarded, the rest being less than 10% of the amount of tools (N=1144). However, it should be remembered that these items represent discarded, abandoned artefacts, while most of the finished products were probably exported. Therefore, the relative amount of tools does not necessarily reflect an intensity of the industry, but rather the frequency of discarding, possibly as a result of manufacturing errors.

An important aspect is the presence of workshops for various tool types (concentrations of blade and bladelet cores at site A and the workshop of sickle-blades at site B). The inhabitants were partial specialists in various tasks, such as production of flint tools and beads (site M).

55

CHAPTER III –NEGEV SITES

Stage 2: Production of blanks

Stage 3: Shaping/Retouching stage – retouching of tools4 (Table 6)

More than 9000 pieces were collected by the three excavations. During the earliest one, most débitage was collected from: factory sites 1 and 2, and middle 29, although material was found virtually in all excavated units. Some bladelets cores were found stuffed in broken pots. Among débitage, 80% is on wadi pebble flint, the rest on semi-translucent flint, even though pebble flint flake-cores are in minority compared to other cores.

Retouched blades (N=11; 1%) Retouched blades have unilateral, bilateral discontinuous, continuous or bilateral retouch. No backing occurred. Cross sections are trapezoidal or triangular. Two blades are on tabular flint, the rest are on pebble flint. Tabular flint blades are longer than the others (120 mm average, opposed to 60 mm average, respectively).

There is a clear distinction between knapping wadi gravel and semi-translucent flint. The dominant butt type on semi-translucent flint waste products is punctiform, with little or no bulb of percussion, whereas among wadi gravel débitage, plain butts are most common. Large flakes and blades have mostly plain butts, except blades longer than 80 mm, with punctiform butts. Primary elements (5% of débitage) have large, cortical or plain butts and prominent bulbs of percussion. Prominent bulbs of percussion on flakes as opposed to diffuse or nonexistent ones on blades/bladelets also suggest different techniques for the production of blanks. The scars pattern on the dorsal face of blades is parallel, while flakes have amorphous or unipolar patterns. Most core trimming elements are on semi-translucent flint. They consist of core tablets, ridges, flanc de nucleus and bases of cores.

Truncated blades (N=45; 4%) These items have a straight truncated distal end; occasionally the proximal end was truncated as well (13 items). All but two items (on tabular flint) are on wadi pebble flint. Bi-truncated blades were commonly backed as well. They are larger than sickle-blades, probably being discarded during the process of sickle-blade production. Notched and denticulated blades (N=11; 1%) All items are on wadi pebble flint; notches are on the mesial or proximal parts; 8 items have only one notch. Utilized blades (N=69; 6%)

Over 1400 cores were collected; ca. 2/3 of them are bladelet cores, the remaining ones being blade cores. Only an insignificant number of flake cores, made on pebbles, were collected3. They are by-pyramidal, with an average of nine flakes removed.

Two-thirds of them have wear marks on both edges. Raw material is usually pebble flint, some are on semitranslucent or tabular flint. Butts are usually punctiform.

Bladelet cores are almost exclusively made on semitranslucent flint, used for the production of bladelets. Most have one striking platform and are pyramidal, conical or prismatic in shape. The remaining cores have two or more platforms. All cores were exhausted. Apparently, they were prepared by splitting the original nodule, the part of the resulting section being used as a striking platform. The edges of the striking platform were prepared and spurs removed before knapping. Most cores have no cortex on their débitage surface. The possibility that pressure flaking was used for the production of bladelets cannot be excluded.

Microliths, on semi-translucent flint, were sub-divided into retouched and truncated. They were locally produced, as suggested by a high amount of waste products on semi-translucent flint, from all production stages. Retouch is fine and parallel, on the distal end and occasionally on the mesial part. Truncations are straight, concave or convex. (these are in fact micrograttoirs, see Gilead 1984). Their provenance and their waste were found throughout the site, and in two concentrations.

Retouched bladelets (N=125; 11%)

Sickle-blades (N=109; 10%) These items, made on blades with sub-parallel straight edges and triangular or trapezoidal cross-sections, have gloss along one of their edges. Their length was controlled by either truncation or deliberate breakage. Half of them are backed, either unipolar or bipolar; 1/3 are backed and denticulated; 9% are not backed – the opposite edge to the working edge is retouched by direct or bipolar retouch.

Blade cores, made on wadi pebble flint, were rarely fully utilized, as evidenced by the retaining of cortex on almost half of their débitage surface. Most have one striking platform, a prismatic or pyramidal shape and an average length similar to that of blades. Distal and lateral modification of the core to guide the blank removal occurs. The striking platform is smooth. They were used to produce sickle-blades.

4

The amounts of tools reflect the quantity of items available for analysis by Roshwalb (1981). It should be remembered that probably not all the lithic material was collected during these excavations, therefore their integration into a synthesis is problematic.

3

The possibility that flake cores were not collected cannot be ignored.

56

CHAPTER III –NEGEV SITES

The tip is formed by lamellar convergent or semi-parallel retouch on the upper surface. Secondary short retouch is placed parallel to and over the first flaking on one or both faces of the tip. Over 200 bifacial spalls were collected. Apparently, the knappers of bifacials changed hammerstones during their production: first stages were performed with hard hammer and final shaping with soft hammer.

Borers (N=82; 12%) Borers, on wadi pebble flint, were divided into awls (N=25), made on flakes (except 3 items on blades) and drills (N=53), made on blades with backed edges. A distinct type is the micro-borer (N=4), made on semitranslucent flint. A unique artefact is a “Ghassulian star”, commonly found in Chalcolithic Golan sites5.

Retouched flakes (N=43; 4%)

Burins (N=5; 1%)

This group consists of retouched items (N=19), utilized pieces (N=17) and notches (N=9). The common raw material is wadi pebble flint, or, in lesser occasions, semitranslucent flint. Retouch vary in its location and morphology. Butts are plain, linear or dihedral. Bulbs of percussion are prominent.

All burins are made on breaks. Two burins are made on sickle-blade fragments. Scrapers (N=151; 13%) Several sub-types were defined: end-scrapers on (occasionally retouched) blade (N=9), sidescrapers (N=24) on primary flakes, with a prominent bulb of percussion and a cortical, plain or linear butt, evidence of hard hammer use and scrapers on flake (N=9), with retouch along at least one end and edge. Endscrapers on flakes (N=32) are differentiated from previous ones by location of retouch – only on their distal end.

Multiple tools (N=4; 0.3%) These are two scraper/points, one scraper/truncation and one truncation/point. Perforated discs (N=1; 0.1%) This is a bifacially retouched tool, with a central perforation.

Tabular scrapers (N=77) represent a distinct category. They were produced at the site, as evidenced by the high amount of tabular flint blanks, and of tools discarded at various stages of modification. It should be noted that no cores, primary elements or debris of tabular flint was collected at the site, on neither one of the excavations. Apparently, the site was used for a secondary-stage of tabular scrapers production.

Summary Site A was occupied during several episodes, as testified by its stratigraphy. The lack of architectural structures, the preponderance of lithic waste and utilitarianism of pottery suggest it is a site of industrial importance, best exemplified by the presence of the ca. 1400 cores collected at the site. The presence of polished basalt objects, however, seems to contradict its purely functional nature (Roshwalb 1981:88). The ceramic repertoire conforms to the description of the lower levels of Ghassul6.

Tabular scrapers share some common characteristics: all have faceted butts, similar thickness (between 7-8 cm) and a prominent bulb of percussion; the retouch is parallel and abrupt. Occasionally, the bulb of percussion was thinned. Blanks were usually transversal flakes. Four items have an invasive, bifacial retouch. Two items resemble Egyptian knives. Tabular flint scrapers were found in several concentrations at the site, as evidenced by a large group of unfinished items found in one locus.

Site B The site, excavated by Macdonald, is located northwest of site A, on the west bank of Nahal Besor. It is typologically indistinguishable from site A, though the absolute numbers and relative frequencies of tool-types differ (Table 6). The site consists of pits of various size and shapes; the presence of a small “floor” for the production of sickle-blades was noted by Macdonald.

Bifacials (N=300; 34%) Site A has the biggest amount of bifacials among the sites of this area; adzes dominate (N=106), followed by axes (N=62), ogivals (N=48) and chisels (N=16), choppers (N=5) and picks (N=7). Broken (N=102) and unfinished items (N=61) complete the group.

The flint assemblage consists of 149 tools and 48 waste products. Among tools, several types, noted by Macdonald, were not found during the study of the collection (such as retouched bladelets or microborers).

Bifacials were usually shaped from two striking platforms located along edges, on pebble flint chunks.

5

6

These items are commonly classified as a distinct tool-type; the description here follows the original publication

The site was attributed to the Besorian phase, on ceramic typology grounds by Gilead (1990).

57

CHAPTER III –NEGEV SITES

Sickle-blades (N=28) are usually backed, truncated and the working edge denticulated. Another variant of sickleblades has inverse retouch, though still backed and denticulated. Items of this type are usually longer and wider than the former one.

Sickle-blades (N=13) are similar to those of site A, sharing characteristics such as back, denticulated working edge and truncations. Two items are truncated blades; other two are truncated bladelets, on white patinated flint. 12 items are borers: one is a drill, the rest awls on flakes.

Borers (N=19) are divided between awls (N=15) and drills (N=4). They are on flakes or blades, on various types of raw materials. One item is a microborer.

Scrapers (N=6) are of various size and shapes, retouch being either on the distal end, or along the edges. Some are transversal scrapers, others are made on blades.

Only one burin, on retouched break, was collected from the surface.

Tabular scrapers (N=11) were divided into convex side scrapers (N=9), transversal scraper and oval scraper. Retouch is parallel semi-abrupt, along a cortex covered blank. The bulbs of percussion were thinned.

Scrapers (N=12) have a retouch along the edges, on the distal end, or they are rounded scrapers. All scrapers are made on wadi pebbles.

Bifacials (N=87) are similar to those of site A; 14% are unfinished; others are adzes (N=21), axes (N=15), ogival (N=17), chisels (N=4) and non-identifiable items. The reconstruction of the main stages of bifacials’ preparation is: first the sides were shaped by lateral retouch, from striking platforms located along the edges, then the working edge was shaped by longitudinal retouch; a final stage included a second-stage retouch on edges and extremities. Other bifacials are 3 choppers and 4 picks.

Five tabular scrapers were collected, from the upper layers of the site. They were shaped by parallel semiabrupt retouch; their butt was usually removed. Bifacials (N=60) are the largest group of tools collected. Almost half of them are adzes with a triangular shape; others are ogivals (18%), with a rounded narrow tip, or indeterminate (13%). Remaining bifacials are axes (5%), a chisel and 12 roughouts. Half of the bifacials were surface collection, the rest from intermediate layers and 16 from the lower layers. Two varieties of workmanship are discernible: a gross working of large deep flakes, often scalar and little regularity, and a finer retouch at the tip, used for shaping solely adzes. Longitudinal retouch, used for shaping the working edge of bifacials at site A and B, is absent at site D.

The tool assemblage is completed by one notched flake, one burin on scraper and 6 retouched/broken fragments. Among débitage, cores dominate (30 out of 48 items collected): 17 bladelet cores, 11 blade cores, the rest being flake cores. Cores are similar to those of site A (see above). The remaining débitage consists of blades and bladelets.

Other tools are one chopper, one pick, six retouched flakes, one notch, one bifacial modified into a borer and one flake with bifacial retouch.

Site D The site, on the west bank of Nahal Besor, was dug by Macdonald in the 30’s, by Perrot in the 60’s and by Alon in the 70’s. Macdonald defined two levels; the analysis of finds corroborates the existence of 2 occupation phases. Perrot’s excavation revealed 3 levels: a lower one of pits, covered by later semi-subterranean structures, and a mixed layer between them. Alon’s excavation uncovered a paved surface and a mud-brick superstructure along with semi-subterranean structures. The ceramic material displays a mixture of Neolithic and Chalcolithic types, differentiated by their raw material as well. No clear stratigraphic context or feature could be correlated with Neolithic sherds. The tool assemblage (Table 6) is similar to site A (see above).

Only 8 blades and 14 cores were collected. Cores are blade (N=5), bladelet (N=5) or flake cores. Blade and flake cores are on wadi pebbles, bladelet cores are on semi-translucent flint. Being similar to those of site A. Site M The site is located on the east bank of Nahal Besor, on the southernmost point excavated by Macdonald, who unearthed 5 floors. The lower ones have a mix Neolithic and Chalcolithic pottery, while the uppermost level is similar to site H (EB-I). The main floor is the second, similar to sites A and B. It was excavated in 9 units; one contained a workshop for beads production.

Retouched blades (N=3) are made on wadi pebbles; their retouch is different on each item: scalar retouch, semiparallel along both edges and backed. Butts are linear. Retouched bladelets are on semi-translucent flint, modified by fine retouch. Two additional bladelets have a modified distal end (micrograttoirs). Butts are punctiform and bulbs of percussion small. There are three truncated blades, all backed.

The flint industry is dominated by tools on bladelets, produced in a comparable mode to site A: on semitranslucent flint, with punctiform butts, no bulbs of percussion and probably by pressure knapping. Among them, 19 are retouched bladelets, mostly broken, and 20 truncated bladelets.

58

CHAPTER III –NEGEV SITES

Site E

Sickle-blades (N=9), on pebble flint, are rectangular, long and wide, bi-truncated and backed, with fine denticulated working edge and triangular or trapezoidal cross-sections.

The site is located on the east bank of Nahal Besor, near the spring of Ain Farah. It was excavated by Macdonald, who reported that it was badly eroded. A rectangular structure, pits and hearths were uncovered, all by crosstrenching excavation. The flint assemblage is small (40 tools and 7 débitage), though Macdonald reports a large amount of sickle-blades on grey flint. The analysed tool assemblage, similar to the one described above, is dominated by scrapers, bifacials and sickle-blades.

The borers group consists of 3 awls (on flake) and 8 drills (on blade). All but two items, on semi-translucent flint, are on wadi pebble flint. A huge amount of microborers was collected from the second floor: Macdonald reports of over one thousand pieces, found together with ostrich shells, different stones (carnelian, quartz crystal and feldspar) and beads on various stages of production. Seven types of microdrills were defined (Roshwalb 1981:167), probably reflecting different phases in the manufacture of beads, or different stages of abrasion. All are on blades or bladelets, shaped by bilateral backing.

Summary (of Roshwalb’s conclusions) The Chalcolithic flint industry of Wadi Ghazzeh sites is composed of five major production categories: a blade industry - sickle-blades, a bladelet industry - microliths, a flake industry - scrapers, borers and other tools on flakes, a bifacial industry and a tabular scraper industry. Cores are mainly associated with the first three industries. Blade cores have usually one platform and a pyramidal shape. Bladelet cores are exclusively made on semitranslucent flint, being exploited from various striking platforms. Occasionally, faceting signs on striking platforms are discernible on both types of cores. Flake cores are rare, while tabular-flint cores are absent. Probably tabular flint blanks were brought to the site and locally modified. Bifacial tools were made on-site.

Scrapers (N=13) are on wadi pebble flakes, either cortical or not, with the retouch located on the distal end, along the edges, or both. Some are carinated scrapers; others were shaped by either denticulation or steep retouch. Bifacials (N=28) are coarsely worked large pieces or finely shaped triangular items, with polished tips. Most of the bifacials are roughouts; others are adzes, with a triangular shape and the tip modified by lamellar flaking or by coarse retouch. The rest are axes, ogivals and broken / unfinished items. The tool assemblage is completed by a chopper on pebble, a bifacially retouched blade, two notches on flake and a retouched flake.

Apparently, blades were produced with a soft hammer, bladelets with pressure retouch and/or soft hammer, while flakes with hard hammer. During the production of bifacials, hammers were changed, depending on the production stage. Flakes show no special preparation prior to their removal.

The débitage consists of bladelets and cores, mostly found on the second floor. More than half of the cores are blade cores, even though more than half of the tools are made on bladelets. Blade cores were made on either semi-translucent flint, or wadi pebble flint. Their striking platforms are smooth, though faceting also occurs.

The tool assemblage is composed of various types. Tools on blades are rare, excluding sickle-blades. Retouched blades have not a standard morphology, retouch appearing in various shapes and locations. Retouched bladelets are common in all sites. They are mostly made on semi-translucent flint, with retouch varying from nibbling to abrupt retouch. A distinct microlith has the distal end truncated (micro-endscrapers).

A more recent excavation (Gophna and Gazit, 1985), revealed three pits containing flint artefacts, grinding stones, pottery sherds and animal bones. Above the fill of one pit, below the top-soil loess, a unique flint artefact was found. Its sizes are 76 cm long, 6 cm in diameter and weighs 3 kg. It has an equilateral triangular cross-section. There are no visible signs of use. Its blunt edge, which may be considered its base, preserved the brown cortex of the Eocene flint block from which the tool was shaped. The other end is pointed. This type of raw material is prevalent east and southeast of the site, ca. half-a-day's walk therefrom. The tool was shaped in a similar mode as most bifacials, namely by repeated shallow removals. There are over a thousand scars of removals on its surfaces. Its unique size makes it the largest flint tool ever discovered in the Near East. An apparently similar tool, though only 20 cm long, was discovered in the adjacent site B. The remaining flint implements are similar to the ones collected during the Macdonald’s excavation.

Sickle-blades are backed, bi-truncated and have a finely denticulated working edge, with triangular or trapezoidal cross-sections. Borers are made on non-standardized flakes; among them, microdrills and drills are apparently more specialized, made on double backed blades. Scrapers are made usually on cortical flakes; retouch is semi-abrupt, located at various points and with various extensions, along the edges. Bifacials, rarely polished, are axes, adzes, ogivals, and picks. Their production is not standardized, as exemplified by the various modes of preparation of working edges. They were probably manufactured locally, as testified by roughouts collected from almost each site.

59

CHAPTER III –NEGEV SITES

The second site (R 48 – 180 m2 excavated) is located further north along the coast, some 700 m. from the shore, in a similar environment as A301. It extends below an active sand dune, the remains covering some 3 dunams. Three occupational levels were distinguished, each separated by a thin layer of sand. The uppermost level phase of occupation included remains of mud-brick structures and installations (ovens, pits and hearths). The second phase is characterized by installations, mostly built of broken grinding slabs, hearths and pits. Some refuse and ash pits, dug into the sterile sand, were attributed to the earliest phase of occupation. Small cuplike depressions found in both uppermost levels may indicate the presence of wooden beams for the support of tents or huts (fond du cabane).

Workshops for bifacials production were found at sites A and B. Two different stages of bifacials production were noted: the first consisted of knapping large flakes, from two striking platforms along edges, while the second, involving probably a soft hammer, was used for further shaping. Choppers appear constantly, in limited amounts; they are pebbles modified by a few lateral removals, in order to obtain a working edge. Roshwalb (1981) recognized two groups of sites: sites O and E, and sites A, B, D, and M. The first group displays a crude flaked flint industry and a fine ceramic industry. Tool assemblages are dominated by scrapers, bifacials and sickle-blades. Tabular scrapers and microliths are absent. Semi-subterranean structures were found in both sites. The second group of sites displays a skilled flint industry with flint tools dominated by bifacials; tabular scrapers and microliths appear as well.

Two additional large sites were discovered during North Sinai expedition, both in the same area as R48, but not yet published. The remaining Chalcolithic occurrences were not extensively excavated, or just recorded and a few indicative finds were collected. The general impression is that most were ephemerally occupied, during which restricted economic activities took place.

North-eastern Sinai Sites (ca. 53.54/73.64) Introduction From 1972 to 1980, a survey and excavations along the Mediterranean coast of the Sinai peninsula took place (Oren and Gilead 1981). Two clusters of Chalcolithic sites were investigated (Map 2), south of modern Rafiah: near Haruvit (5 sites) and west of Yamit (5 sites); partial results were published by Oren and Gilead (1981). The rest of the collected data was analysed, but not yet published; the descriptions presented here are with the kind permission of the head of North Sinai expedition research (E. Oren). The lithic material was investigated under the supervision of I. Gilead. Additional to the published material, Chalcolithic remains were collected from two large sites and from 59 find-spots. Among the later, 24 revealed material from other periods as well; therefore, their flint assemblage was separately analysed. Among the remaining 35, half yielded only waste products and isolated tools. Given the similarity of the collections, they were described together.

Description of the tool assemblage Most tools were locally prepared from wadi pebbles, as testified by the presence of hammerstones, cores and other waste products. Fan scrapers were made on tabular flint. Sickle-blades resemble Be’er-Sheva ones: narrow, backed and truncated, with a fine denticulated working edge, their common raw material being banded flint. Outstanding among tools are microliths and microendscrapers, made on semi-translucent flint. Important to note is the presence of two transversal arrowheads, rare tool-types among Chalcolithic assemblages. Scrapers Scrapers represent ca. 8% of tools in the examined sites (Table 3). They were locally made from a variety of wadi pebbles, without a particular shape or retouch: rounded, transversal or side scrapers occur in similar percentages. Steep retouched and scrapers with a denticulated working edge appear as well. Rounded scrapers on primary elements and scrapers on blades are rare.

General description of sites Among the published material, two sites had been systematically excavated, within a grid of 1x1 m. The first (A301) is ca. 500 m. south of the seashore, on a sparsely vegetated surface scattered with relic sandstone concretions. Its west part was covered by an active sand dune, while on the east and southeast stabilized dunes formed its border. Even though only 700 m2 were excavated, it was clear that the site extends further west and was covered by the sand dunes. Surface remains included clay ovens, hearths, ash pits, installations, stone implements, pottery fragments and lithic material. Several probes showed that the sediment is not deeper than 20-25 cm. Conjoinable pottery fragments from surface and probes suggest that there was only one occupational episode, probably of a short-term.

Tabular scrapers Tabular scrapers are rare (ca. 2%), but their constant appearance indicate a probable domestic use, a simple mode of procurement at a relatively low economic price, despite the fact they were not local products. Most items were found broken. They were made on typical tabular flint, cortex covered and modified by fine retouch. The bulb of percussion was thinned in most cases and the striking platform faceted.

60

CHAPTER III –NEGEV SITES

The presence of a high amount of sickle-blades (ca. 15% of tools, Table 3) and their presence in the surveyed sites denote the importance of agriculture in the activities of the Chalcolithic inhabitants of North Sinai. The unusual high amount of sickle-blades at one of the excavated but still not published site, as well as the presence of blade cores made on banded flint, may represent a local centre for the production of sickle-blades, perhaps for the neighbouring sites as well.

Burins Several burins were collected, usually made on natural break; their relative amount never exceeded 4% of all tools. The variety of raw materials and blanks used for their production suggest that their manufacture was not pre-planned. Borers The borers group comprise awls, drills and micro-drills, collected in almost equal amounts (ca. 6.5% of tools). Awls were manufactured from the available blank flakes without a prior preparation, as suggested by their different size and shape. A common awl is a flake with the point obtained by the retouch of two notches at one of the extremities of the blank. Drills were usually modified by abrupt retouch along the edges of blades, while microdrills were commonly made on semi-translucent flint bladelets. All sub-types were probably locally produced, on waste material left over from more specific-oriented productions, as sickle-blades (drills) and microendscrapers(micro-drills). The variety of sub-types and the different diameters of points suggest a wide range of drilling activities performed with these tools. Retouched flakes, notches/denticulates

truncations

Retouched blades Retouched blades represent ca. 10% of tools, apart of site R45 (ca. 1/4 of tools). Retouched blades were modified by partial dorsal fine retouch along one or both edges. They are distinct from sickle-blades, and it is clear they are the results of a different chaîne opératoire. Microliths Microliths dominate the tools of all (Chalcolithic) survey sites (ca. 20% of tools). Thus, a characteristic of these sites can be identified. Most microliths (micrograttoirs) are on semi-translucent flint, modified by fine retouch along one edge and the distal end, which in most cases is straight, and rarely rounded or oblique. The various shapes of the working edges may denote a deliberate retouch, and not differences in the way they were manipulated. The presence of bladelet cores, of the same raw material as microliths, testifies their local production.

and

Even though this group comprises approximately a quarter of the tool assemblage from each of the excavated sites, their non-standardized nature and high morphological variability hinders any attempt for a more focused description. The possibility that some of these items were “artificially” produced by various postdepositional processes, as trampling and rolling, cannot be excluded.

Bifacials Only a few bifacials were found in all excavated and surveyed sites, the group being also typologically restricted to axes and adzes (Table 3). These were made on chunks, modified by bifacial retouch from two striking platforms located along their edges. Polish appears on most items. The working edge was obtained either by one lateral removal on each face, by longitudinal bladelet removals or by a combination of both. The presence of a few items which were apparently discarded prior their use suggest a local production of bifacials. However, their small amount indicates that they were not an important item in the domestic tool-kit of the Chalcolithic inhabitants of north-eastern Sinai.

Sickle-blades Sickle-blades can be divided by their morphology and raw material characteristics in 3 sub-types (all in similar percentages). Among them, backed, double truncated and with a fine denticulated working edge items are distinct, being similar to typical Ghassulian sickle-blades. Diagonal scars of deliberate breakage can be discerned at their ends. A few items were not backed or truncated, their working edge being only slightly modified. A third group consists of items with an arched back, probably hafted at the point of a curved sickle.

Summary Apparently, the sites were seasonal settlements or encampments occupied by small human groups, which abandoned them peacefully after short periods of time. Little evidence, such as at site R48, indicate that the Chalcolithic inhabitants returned to the same spots on several occasions. The presence of pits, grinding stones and other installations suggests that the occupation was not ephemeral, and probably reflects planning of the movement scheme. Smaller sites may reflect short-term occupation, task-specific oriented, part of the same transhumance scheme.

Most sickle-blades were produced on banded pebble flint. Occasionally other types of pebble flints were used as well. Given the high standardization of sickle-blades on banded flint and the low amount of cores on this type of flint, the possibility that these items were brought on-site as end products cannot be excluded. Items on other types of raw materials may denote local production, probably to fulfil a specific need.

61

CHAPTER III –NEGEV SITES

The wide range of the flint tool-types collected at the excavated, larger sites, suggests a variety of activities probably typical of domestic self-sufficient economy, similar to that of larger sites, as those in the Be’er-Sheva area or Grar (cf. above). Most tools were locally produced; other tools were imported, such as tabular scrapers and probably some of the sickle-blades. The presence of a high amount of sickle-blades and grinding slabs suggests that harvesting played an important role in the subsistence economy of the Chalcolithic inhabitants of the above described sites. Some relations with predynastic Egypt are testified by the presence of a Gerzean (Neqada) palette and Egyptian pottery fragments at site R48. However, these relations are not reflected in the lithic assemblages, excluding the high amount of micro-grattoirs, a common tool-type in contemporary Egyptian and Western Negev assemblages. Since most of the cultural material is well connected to the Be’er-Sheva Ghassulian in general and to Besorian sites in particular, the Egyptian relations may have had a goods-exchange nature (Oren and Gilead 1981). The occupation of the north-eastern coast of the Sinai peninsula may be an extension of the settlement in western Negev, as reflected at sites such as Grar and those of the Besor basin. The high amount of sites discovered, and the probable existence of further more sites still covered by dunes reflect the intensity of occupation of this area in the Chalcolithic period. Even though the contemporaneity of sites cannot be evaluated, the homogeneity of cultural material suggests that they all belong to a single cultural occupational phase, of a limited number of hundreds of years.

62

CHAPTER IV – CENTRAL ISRAEL SITES

During works on behalf of Israel Railway Authority, severe damage was caused to the western margins of the site. In the section of a ditch, it was possible to discern several ash pits. The present, salvage excavation, revealed pits and installations, all dated to the Chalcolithic Period (Khalaily and Hermon in press). Some are shallow and wide, while others have a deep bell shape (1.10m), which contained sherds, stone vessels and flint artefacts. The excavation extended over 6 squares, located close to the damaged points, confined to the railway outline in the North - South orientation. A total area of 150 sqm was excavated. The excavation revealed one archaeological layer, subdivided into two levels:

CHAPTER IV CENTRAL ISRAEL SITES Gat-Govrin – Wadi Zeita (Nahal Komem) (129/116.8) Introduction The transition from Chalcolithic to EB periods in the Northern Negev is still in debate, partially due to a lack of periodization consistency (cf. Joffe and Dessel 1995, Levy 1992, as opposed to Gilead 1994, 1995) and interpretation of material culture (Braun 1996, Gilead 1988). At some sites, such as Arad (Amiran 1978) and Tell 'Erani (Yeivin 1961), archaeological data indicate a clear cut between Chalcolithic and EB age strata. There is no clear evidence for either continuity or transition in occupational phases or cultural material. In other sites, such as those near modern Ashkelon, recently investigated by Golani (pers. com.) and Khalaily (pers. com.), there is apparent clear evidence of cultural continuity between the Chalcolithic and EB Age, in some aspects of the material culture.

1. The upper most level is topsoil, with an average thickness of 0.4 m, consisting of a clayey soil (loess), with a brown to grey colour. This level was disturbed by agricultural deep ploughing. 2. The middle level is horizontal; it appeared in the entire excavation area; its thickness is 0.2 m. This level is a crispy, light grey sediment and contains a high concentration of potsherds and flint artefacts. At its base, it was possible to identify outlines of pits with a circular or elliptical shape.

One of the sites that may contribute to clarifying this issue is Gat-Govrin – Wadi Zeita, excavated by Jean Perrot in the early sixties. The collected cultural material was found in pits from two strata. Perrot dated the lower group of pits to the Chalcolithic period and the upper ones to EB I (Perrot 1961, 1962). Based on these finds, he suggested that there is no occupational gap in the sequence of this site between Chalcolithic and EB I periods (Perrot 1968).

Ten pits were uncovered and excavated; six were ash pits, with shallow rounded outlines. They contained a few finds. Two additional pits have a V-shaped outline and contained a grey colour crispy sediment. Both contained pottery fragments, stone vessels and flint artefacts. The pits were sealed with pebbles and fragments of stone vessels. The remaining two are bell-shaped pits, being broad at the base and narrow in the opening. The diameter of the opening ranges between 1 m to 1.2 m, while the base has an incremental diameter of 3 metres. These pits contained dark grey sediment with a few levels of light ash that could be discerned. A number of stream pebbles of various size was collected from the bottom.

Description of the site The site is located in the vicinity of modern Kyriat Gat, circa 2 km west of Kibbutz Gat and 2 km east from the Be’er-Sheva-Tel-Aviv road, in the Southern coastal plain (Map 2). This section is a wide laying coastal plain, extending between Nahal Shiqma in the south and Nahal Lachish to the north. This area is a transitional zone between the Mediterranean climate of the north and the semi- arid of the Northern Negev. The topography is dominated by low loessial hills with an average elevation of 80 m ASL. Ephemeral streams, such as Nahal Komem that runs adjacent to the site, have incised the topography in a deep stream course. Intensive modern farming contributed to the changing of the natural landscape and obscured the natural vegetation.

The bottom of the dig consists of a sterile loess sediment, with no archaeological material, and is most probably the virgin layer (Khalaily and Hermon, in press). Pits were divided into 3 groups: the 1st one is a set of five shallow pits, with a circular outline and a mean diameter of 80 cm, dug into the virgin soil to a depth of ca. 40 cm. These, poor in findings, intersect each other, which indicates they were not opened in the same time. The 2nd group includes 2 pits, ca. 1 m deep. Both have a Ushape, a circular opening and straight walls, while the bottoms are wide. The pit openings were found sealed with large wadi pebbles and fragments of stone vessels. The fill is rich in archaeological material such as pottery fragments, flint artefacts and grinding stones. The third group consists of two bell-shaped pits. They are large, narrow at the opening and wide at the bottom. Both were dug to a depth of ca. 1.5m. The fill consists of loosely dark-brown sediment alternated with fine ashy layers, deposited horizontally.

Members of Kibbutz Gat discovered the site during the late fifties, where many Chalcolithic and EB vessels were scattered over an area of 35 dunams, after deep ploughing. Since then, the site was surveyed each year by Yehuda Dagan (Dagan pers.com). After digging at Abu Matar and Safadi (Perrot 1984), Perrot investigated other Chalcolithic sites in Northern Negev, such as Gat-Govrin (Perrot 1961, 1962). He excavated ca. 50 sq. and uncovered pits of various size, but no architecture remains.

63

CHAPTER IV – CENTRAL ISRAEL SITES

The distribution of raw materials is similar among the various features: some 1/2 of artefacts are surface collection, ca. a 1/3 in the living layer and the shallow pits, and 4-to 8% were collected from the bell-shaped pits. The only discrepancy is the appearance of grey flint in higher quantities in bell-shaped pits than in shallow pits. This picture may be the result of a knapping activity, its waste being disposed in those pits.

Description of the flint assemblage The flint assemblage consists of 241 items (Table 3). Most waste products and tools types appear. Débitage represents ca. half of the flint assemblage, followed by debris (26%) and tools. Among waste products, flakes are dominant (64% of the débitage). Blades represent 17% of the débitage group, reflecting the importance of this kind of blank for the production of tools (see below).

Stage 1: Production

Almost half of tools were made on blades, and only a third on flakes. An apparent discrepancy was noted between the ratio flakes/blades among waste products (3.8) and among tools (0.8). A possible explanation is the import of tools to the site. However, since waste products of all production stages were found, a local manufacture of most tools is expected. Therefore, a high number of flake blanks may reflect the exploitation mode of the cores, and not the orientation of the flint industry.

Only 12 cores were found: 4 are on canaanean flint, 4 on brown flint, two on semi-translucent flint and 2 on pebble grey flint. The presence of a relatively high amount of cores on canaanean flint, a type of raw material that is not common in the area of the site, reflects a deliberate preference for this type of flint, possibly due to its quality. Other factors may be cultural (a changing fashion for the selection of raw materials) or economic (an established network of procurement of this kind of flint). The importance of its presence is reflected also in the high exploitation degree of these cores, being only ones with three striking platforms, used for the production of both flakes and blades (Plate 25:1). Semi-translucent and brown cores were mainly used to produce bladelets (Plate 25:2). Two radial cores were used to produce flakes.

A main characteristic of the assemblage is the presence of Chalcolithic Be’er-Sheva sickle-blades (on banded grey pebble-flint, double truncated, backed and denticulated) alongside EB canaanean blades (on brown flint, with a trapezoidal cross-section), some retouched and with gloss. Their stratigraphic context excludes the possibility of intrusion. This presence will be discussed from two points of view: the appearance of two tool-types, used as cultural hallmarks of two distinct periods, and the presence of a tool-type, with end-products, results of two different chaînes opératoires.

3 core trimming elements were collected: one tablet on canaanean flint, and one tablet and one ridge on translucent flint. They support the suggestion made regarding the deliberate choice for canaanean flint and of semi-translucent flint for the production of blades and bladelets respectively.

Stage 0: Acquisition – Extraction and testing of nodules

A few chips and chunks were collected, even though all archaeological layers were sieved.

Several raw material types were used; most common are Eocene brown (“canaanean”) and wadi pebble flints (grey and brown). A distinct category is banded Neogene flint, found in the local wadi bed, commonly used in the Northern Negev for the production of Chalcolithic sickleblades. Another distinct flint type is grey or brown semitranslucent flint, found in the region as small pebbles, probably of Senonian nodules origin. It was extensively used for the production of microliths. Other raw materials, found in limited frequencies, are breccious brown flint and flint embedded with microfossils, found mainly on waste products.

Stage 2: Production of blanks Although small, the flint assemblage contains all elements of a lithic industry that most of its reduction stages were performed on-site. Therefore, it may be assumed that most tools were locally produced. Some of them, such as tabular scrapers and sickle-blades made on banded grey pebble flint may have been manufactured elsewhere and brought to the site as end-products. Even though no blade cores on canaanean flint were found, several cores on this type of raw material were collected. The possibility that some of them were used for the production of blades at an earlier stage of their exploitation cannot be ruled out.

The distribution of raw material’s types, as observed on all analysed material is as follows: Eocene brown 44%, Senonian brown 22%, semi-translucent 15%, pebble grey 11% and banded pebble 15%. Several collection locations of raw materials may be indicated: Canaanean flint may may originate from the Negev highlands, brought to the site after a modification of nodules at the quarrying place, pebble flint (grey and banded grey), found in the main (nearby) streams of the Negev wadis and brown and semi-translucent flint, originating probably from the hills surrounding the site.

Flakes More than a 1/3 of flakes are on canaanean flint, which hint at the intensity of exploitation of canaanean cores and also the preference of this kind of raw material for the manufacture of tools. The comparison between the amounts of raw materials among blank flakes and tools’

64

CHAPTER IV – CENTRAL ISRAEL SITES

raw materials shows a similarity concerning brown flint, canaanean flint and semi-translucent flint. The high amount of tools on banded flint may reflect an import of these items from an unknown source, with possible relations with major southern Chalcolithic sites, such as the Be’er-Sheva sites. The small amount of blanks on banded flint reflect a low-scale exploitation of this type of raw material. Tools on tabular flint (tabular scrapers) were all probably brought as end-products.

Borers (N=3; 8%) The borers are: two awls on simple flakes, with the point shaped by two notches, and one drill on blade, backed along its edges. Retouched Flakes and Notches (N=8; 20%) Retouched flakes and notches are common in Chalcolithic assemblages (Rosen and Levy 1987). Their low amount indicates a notable difference of the discussed tool assemblage, compared to Chalcolithic flint assemblages. However, given the small size of the tool assemblage, no firm conclusions can be drawn from this kind of comparison.

Flakes on breccious flint, not used to produce tools, may be natural products, or unsuccessful attempts to use this kind of raw material. A few flakes on patinated white flint, collected from the surface, may represent earlier, Palaeolithic intrusion.

Retouched Blades (N=8; 20%)

Several types of striking platforms were noted, the common one being flat (59%), followed by natural (14%) and faceted (12%). Punctiform and crushed butt rare. A higher amount of translucent and canaanean flakes have a punctiform butt related to flakes on other raw materials.

All except two items are on canaanean flint; one edge was partially fine retouched (Plate 27:1,3). The term that best describes them is “canaanean blades” (Rosen 1997:46). The other two items are on semi-translucent flint, resembling (large) microliths.

Stage 3: Shaping/Retouching stage – retouching of tools

Canaanean blades are fossile directeurs of EB assemblages. Therefore, their appearance in a clear Chalcolithic context has to be explained. At this point, suffice it to say that the presence of waste products made on the same raw materials, and the almost total absence of EB pottery reduces the possibility of intrusion within the Chalcolithic layers.

The tool assemblage (N=39) typologically resembles other Chalcolithic assemblages (Table 3). However, in relative quantity, there are some major differences, which resemble EB assemblages (Rosen 1997). Thus, sickleblades and retouched blades represent the majority, while scrapers and retouched flakes appear in insignificant numbers.

Sickle-Blades (N=10; 26%)

Scrapers (N=3; 8%)

Sickle-blades are the most common tool-type (Table 3). All items were discarded after an apparent extensive use, as evidenced by the abraded and broken conditions of these items.

The scrapers (Plate 25:3), on pebble grey flint and a rounded working edge obtained by semi-abrupt retouch, can be defined as Be’er-Sheva scrapers.

Sickle-blades may be divided in two groups, their unifying attribute being the appearance of gloss along their edges. Half resemble (Chalcolithic) Be’er-Sheva sickle-blades: on banded flint, they are backed and truncated and with a denticulated working edge (Plate 35:4-6). The second group consists of two typical (EB) canaanean blades and one, on canaanean flint, with a partial back and a denticulated working edge (Plate 27:2). The “Chalcolithic” characteristics of this item are the partial back, straight truncation and denticulated working edge. The “canaanean” characteristics are the raw material and size.

Tabular Scrapers (N=3; 8%) Three items were found: two have similar shapes (Plate 26:2), the third one being too fragmentary to evaluate its original shape and attributes (Plate 26:1). The first two have an elliptic shape, obtained by parallel retouch. The bulb of percussion was removed by ventral invasive retouch. No striation marks, found on EB tabular scrapers, were noted. The typological differentiation of tabular scrapers was already debated (Rosen 1997), and the recognition of possible characteristics that would differentiate Chalcolithic from EB fan scrapers was discussed (Marder and Hermon in prep.). The tabular scrapers of Nahal Komem cannot serve as a cultural hallmark for neither Chalcolithic or EB periods.

The main differences between the groups are the use of different raw materials (banded versus “canaanean”), faceted butt (in the second group), size and the presence of truncation and backing in the first group. From a strict knapping method point of view, there is little difference between the groups, which may be placed along a continuum of a single knapping tradition. Since there is no documentation demonstrating the functional or

Burins (N=1; 3%) Only one burin, on a natural break was found. The raw material is banded flint, a possible non-utilizable blank prepared for the production of retouched/sickle blades.

65

CHAPTER IV – CENTRAL ISRAEL SITES

This excavation, together with the former one (Perrot 1961), exposed only pits of various size and shapes, with no related features, such as walls or installations, which would indicate a more permanent settlement. The absence of built structures can be explained either by the fact that the location of the excavated areas are in the site's periphery and its centre was not yet exposed, or families of mobile herders that used the open space for their domestic activities, using mainly pits, inhabited the site. Identical layouts were reported from several Chalcolithic sites in Northern Negev (Gilead 1986; 1995). Another indication that seasonal groups inhabited the site is the nature of these pits, which cut each other and were filled with diverse sediments. Their size, generally small and narrow, hint at their use as refuse pits and not subterranean structures.

technological efficiency of one type over the other, the differences may be regarded as stylistic, apparently denoting two different cultural traditions. However, the presence of items with “mixed” attributes (such as sickle blades on “canaanean” flint, wide and with a trapezoidal cross section, but backed and truncated) are indicative of a connecting link between these two apparently distinct traditions. Consequently, the presence of these apparently distinct tool-types may be interpreted as an experiment in the exploitation of a new kind of raw material, applying a similar knapping technique. Assuming a relative late date within the Chalcolithic period for the site, the first use of a new type of raw material, which later became dominant, should not be surprising. Microliths (N=2; 5%) Only one micrograttoir (Plate 27:7) and one broken retouched bladelet (Plate 27:8) were found, both on semitranslucent flint. These artefacts are typical of Chalcolithic flint assemblages, but appear in EB contexts as well (cf. Rosen 1997:65).

Khirbeth Aliya East (ca. 146/123) The flint assemblage, collected during salvage excavations at Ramat Beth Shemesh project from two adjacent (100 m.) sites – 84/16 and 84/20 (Map 2), was analysed by I. Zbenovich (in prep.). At the former site, no architectural features were revealed. Apparently, Chalcolithic settlers used cavities and hollows in the exposed bedrock for various activities. The later site displayed a variety of structures and installations; the Chalcolithic layer was later disturbed by intrusions and was partially washed away.

Celts (N=1; 3%) A single bifacial was found, a fragment of the polished working edge of a chisel. Bifacials are rare in EB assemblages (Rosen 1997). Stage 5: Discard

Even though a large amount of flint items were collected (among the largest ever analysed from Chalcolithic sites – 27045 items) and they underwent a systematic analysis, some methodological problems limited the integration of results into a broader synthesis. There is an ambiguity regarding the provenance of more than 2/3 of the lithic material from each site, since only a brief report regarding the context of only a 1/3 of the assemblage is reported. Moreover, their location is mostly surface or fill from apparently disturbed layers, for which postdepositional processes are not fully understood.

Almost half of the assemblage was collected from the surface. A quarter was found in an archaeological layer below the surface, the remaining items originating from six shallow pits (19% of the assemblage) and from two bell-shaped pits (10% of the assemblage). There is little difference in the distribution of waste products and tools among features (surface, living level or pits). The large amount of waste products collected from well-defined stratigraphic contexts, such as shallow pits and bellshaped pits, reduces the possibility of intrusion and mixture of items belonging to two potentially distinct cultural entities (Chalcolithic and EB). Therefore, the archaeological context of the flint artefacts can be used as an anchor for relating the assemblage to one cultural entity and to treat it as a homogeneous group.

The assemblage was not analysed according to conventional standards employed in the analysis of Levantine proto-historic sites; for example, the micrograttoir, a well-defined Chalcolithic tool-type (Gilead 1984) was described as “…fashioned mostly on small flakes…but also on large/medium size flakes …” (Zbenovich, in prep.).

The similarity between the relative amounts of flint groups among the various features indicate that these were not related to the production of flint tools, but rather reflect discarding activities, possibly by different knappers and/or in different episodes.

Another problematic issue is the presence of a nonneglectable number of tools belonging to different pre and proto-historic periods (N = 253, according to the text, or 247, according to Table 9 of the mentioned report). They were attributed to the Epipalaeolithic, Neolithic and EB periods. The probable influence of these tools on the general flint assemblage was almost entirely ignored and the waste products collected were analysed as a homogeneous assemblage.

Summary Apparently, the site was inhabited towards the end of the Chalcolithic period. The superposition of pits suggest that it was occupied in several occasions, possibly by the same human unit, or others of the same culture group.

66

CHAPTER IV – CENTRAL ISRAEL SITES

Disregarding the provenance of flint artefacts (mostly a surface collection) and their probable exposure to postdepositional agents and the adoption of ambiguous typological criteria make a comparative analysis an impossible task. Therefore, only descriptive aspects of selected tool-types were integrated in this work. Worth mentioning is the presence of sickle-blades (Table 3) with typical Be’er-Sheva Ghassulian attributes: raw material (banded flint), size (average of 12.5 mm. width) and morphology (backing, truncations and finely denticulated working edge). They constitute only about 10% of the collected sickle-blades, the majority of which were made on local, translucent to semi-translucent Senonian flint. A characteristic of this group is the presence of sickle-blades on bladelets, an uncommon feature of Chalcolithic sites of other areas. In general terms, the tool assemblage contains all tools typical of proto-historic assemblages: scrapers, borers, bifacials and a high number of retouched flakes and denticulates. Thus, the sites of Ramat Beth Shemesh may be related to the Be’er-Sheva Chalcolithic. However, the nature of this affiliation is unknown, as well as the nature of the Chalcolithic settlement in the area and its characteristic lithic material culture.

67

CHAPTER V – NORTHERN ISRAEL SITES

CHAPTER V

Beit Netofa valley sites (ca. 170/210)

NORTHERN ISRAEL SITES

The following paragraph describes a small amount of flint tools originating from surface collection at two Chalcolithic sites located in the hilly, western flanks of Beit Netofa valley, in the Lower Galilee (Gilead 1989). Even though no architecture was discernible on the surface, the high concentration of artefacts and the ashy sediments are evidence for extensive occupation at these sites. Among the flint tools collected, discernible are sickle-blades, bi-truncated, backed and with a finely denticulated working edge, tabular scrapers, chisels, some polished, with the working edge shaped mostly by several longitudinal removals, and scrapers. Outstanding is the presence of perforated discs, common in Chalcolithic assemblages of the north.

Tel Teo (ca. 210/285) The site, located on the western slopes of the Huleh valley, near the spring of Ein Teo, was excavated as part of a salvage project in 1986 (Eisenberg 1989). Thirteen occupation layers, ranging from the Neolithic to historic periods, were uncovered. The Chalcolithic settlement (Layers VI and VII), consists of three dwelling units arranged around a central, enclosed court. Additional features are circular, sunken silos lined with stone slabs. The architecture and ceramic types associate Tel Teo to Tel Turmus in particular, and the Golan sites in general. Description of the tool assemblage Seven cores were found in clear Chalcolithic context; 6 are irregular amorphous cores, the 7th being a discoidal rounded core. The ratio flakes/blades is 1.3, probably a bias of collection methods. Primary elements and core trimming elements were found as well. The collection method, oriented towards preserving large pieces, mostly tools, limited the study to typological aspects.

Einot Kaukab (174.45/248.45) The site was excavated for three weeks in October 2000, as part of a salvage excavation project directed by N. Getzov, of Israel Antiquities Authority. The material presented below was analysed with his kind permission. Einot Kaukab is located on the eastern slope of a small valley, on top of Wadi Avlaim, on the foothills of Mount Atzmon, ca. 700 m. north of Kaukab springs (Map 2). The scatter of material culture on the surface hints at an extended site, from which only two squares, 5 x 5 m. were excavated, on its western edge.

The tool assemblage (Gopher and Rosen in prep) consists of 148 tools, 70 of stratum VII, and the rest of stratum VI (Table 3). The assemblage of stratum VII is dominated by bifacials, sickle-blades and perforated discs (N=7). Retouched flakes and blades, notches and denticulates are few. Adzes and axes, some polished, have various sizes. Sickle-blades are backed and truncated, some with a finely denticulated working edge. Among retouched blades, one is prismatic, resembling EB canaanean blades. The tool assemblage is completed by 8 knives (5 naturally backed, and 3 abruptly backed), 2 endscrapers, one tabular scraper and 3 awls. The varia group consists of flakes with polish signs (N=4) and broken retouched pieces.

Excavation and recording methods All sediments were sieved and the material kept. Therefore, the material available for study reflects the assemblage as discarded by its inhabitants. The smallest excavation unit was the basket, while locus denotes any distinct human activity. Description of main features

The tool assemblage of stratum VI (N=78) is dominated by retouched flakes and blades (1/3 of tools), sickleblades (18%) and bifacial tools, including perforated discs (16%). Sickle-blades are truncated and backed. Adzes are trapezoidal, mostly polished. All discs are broken. Among retouched flakes (N=22), some, modified by simple to semi-abrupt retouch, are large: length 12.8 to 15 cm. and width 7 to 8.4 cm. Their bulb of percussion was sometimes removed.

Four main occupation layers were identified, with evidence of five occupation stages. The lowermost level (IV), exposed ca. 2.5 m. below surface, consists of three pits dug into the virgin soil. They contained ashy sediments, ceramic fragments and flint artefacts. The pottery was identified by the excavator to a preChalcolithic stage. Since only five tools were found in these pits, which are not characteristic of any period / culture, it was assumed that the flint content of these pits (less than 4% of the entire assemblage) did not influence the nature of the flint assemblage.

Summary The generally small size of the assemblage is due to selective collection. However, certain trends are recognizable. Raw material is mostly light brown and grey. The presence of waste products hints at a local manufacture of flint tools. Several tool-types, such as sickle-blades and perforated discs, are diagnostic of the Chalcolithic period, with some affiliation to Northern assemblages, such as the Golan sites.

Two buildings were exposed, with two occupational stages in the second occupation layer (III). One building was built directly on the virgin soil, with stone foundations and an ashy floor with a flat stone pavement. A later modification of this building included the raising of the floor by 50 cm. It was made of chopped chalk (plaster) and carefully paved with flat stones. An external

68

CHAPTER V – NORTHERN ISRAEL SITES

production of microliths, while tabular flint and grey flint with microfossils appear solely on tabular scrapers (one item) and adzes (two items) respectively (Table 14).

wall, delimiting probably a courtyard, irregularly paved with stones, was annexed to the walls of the building. A second building was found a few metres north of the first one. Adjacent thereto, a small structure, built of plaster and small stones, was discovered.

The strategy of raw material exploitation is best characterized by the use of locally available sources for the production of most tools (light brown and grey flints). The use of semi-translucent flint for the production of some microliths may reflect a meaning beyond functional needs, since light brown flint, suitable for the production of bladelets, was extensively and successfully used as well. Therefore, the possibility that the use of semitranslucent flint for the production of microliths reflects traditional/cultural aspects cannot be rejected.

Layer II consists of several buildings, with walls, severely damaged by modern human activity, were exposed. This layer is overlaid by a fill of ca. 80 cm (the surface – layer I), containing Middle Bronze Age pottery. Only a few flint artefacts were collected from this layer (2.6% of the entire assemblage). Description of the flint assemblage Even though only 2 squares were dug, up to 2 m depth, a large amount of flint artefacts was collected (N = 7932). Since all sediments were sieved and all flint items were collected, a large amount of debris was expected (70% of the assemblage). Apparently, the flint industry was locally produced, from various raw materials. The high exploitation of cores is reflected by their small size and exhausted appearance. Tools represent 3.34% of the assemblage (N=264). Among them, tools on blade, in particular sickle-blades, dominate. However, most other tool-types appear, in various amounts. Important to note is the absence of discs, typical of Northern assemblages.

Almost all raw materials were processed within the site, as evidenced by the waste products collected. Exceptions are tabular and grey with micro-fossils flints, found only on end-products. These items may reflect small-scale exchange networks of flint tools. Stage 1: Production Most cores were used to produce flakes and bladelets. Given the similarity of raw materials of cores and tools, the possibility that cores were used for blade removals at their first stages of exploitation cannot be excluded. The extensive exploitation of cores is well reflected by the high number of removals from each one (approximately 30 blanks from each core). This suggestion is supported by the high rejuvenation rate of cores, testified by the number of core trimming elements (Table 13).

Stage 0: Acquisition of raw materials Nine types of raw materials were identified on tools and waste products. Most common is a light brown nodule flint, used for the production of 50% of all tools, appearing on each tool-type (Table 14). Next common is pebble grey flint, used for the production of a ¼ of all tools. Both raw material types were on-site processed, as evidenced by cores and waste products on these raw materials. Other types of flint (Table 1) had a limited and particular use, being exploited for the production of selected tool-types (see below). At least 5 sources of flint can be pointed out: wadi pebble flint (grey and brown flint), 3 different flint outcrops (light brown and striped brown flint, cream and possibly pink flint) and flint not processed on-site (tabular and possibly grey with microfossils). The provenance of semi-translucent flint is unknown; the presence of a few cores and microliths made on this type of flint indicate its local exploitation.

Cores on light brown flint and grey flint are most common (50% and 20% respectively). Another four types of flint were recognized on cores, among them brown, cream and pink (ca. 10% each) and one bladelet core on a semi-translucent flint. Thus, excluding tabular scrapers and two adzes on grey flint with micro-fossils, it can be assumed that most other tools were manufactured on-site. An average of 6 removals of primary elements from each core was sufficient to clean its débitage surface from cortex and prepare it for further exploitation, most cores having none or a few cortex left. Only a 1/5 of cores retained cortex on their surface, while 3% have most of their débitage surface cortex covered. Thus, even if high quality raw material was available close to the site, its knappers preferred to exhaust cores (mean size are 37 ± 11 by 41± 13 mm).

Light brown and grey flint were noted on most tool-types, but grey flint on microliths, particularly for the production of borers and sickle-blades (Table 14). Apparently, the quality of these raw materials (suitable for the production of any tool type) and their possible abundance enabled their extensive exploitation. Cream flint, used mostly to produce sickle-blades and borers, was less extensively exploited, appearing on only 6 toolclasses. The remaining raw material was found in too small quantities to formulate any firm affirmations; however, several tendencies were noted: semi-translucent flint was exclusively used (except one scraper) for the

Two-thirds of cores were exploited only from one striking platform, the remaining being bipolar (1/4 of cores) or amorphous (Plate 28:4), with three striking platforms (8% of cores). Faceting appears on 1/3 of cores, often on cores with more than one striking platform, as a method to increase their exploitation (among faceted cores, almost half have two or three striking platforms, opposed to non-faceted cores, 75% of them having one striking platform). Discoidal cores

69

CHAPTER V – NORTHERN ISRAEL SITES

Two main knapping techniques were applied for the production of blades, as evidenced by butt types: flat and pointed, both in similar percentages. A comparison between flat and pointed butt blades revealed that:

(Plate 28:1) were occasionally used (4 items), while choppers are practically absent (1 item was found). The common blank removed from cores was the flake: 77% of all cores have flake scars (Plate 28:1,4); bladelets (Plate 28:2) were extensively produced (63% of cores have bladelet scars). A third of cores were used to produce blades. As shown by the observation of scar shapes on cores, few were used for the production of a single blank type; cores changed their aim, according to demand or feasibility. Thus, in strict terms, only 23% of cores were exclusively used to knapp flakes, 7% for blades and 13% for bladelets. A common combination was a core used for the production of flakes and bladelets (Plate 28:3). Only 3% of cores were exclusively used to produce blades and bladelets.

• Pointed butt blades are longer, their length distribution shows a tri-modality around 90, 60 and 40 mm; flat butt blades have a bi-modal distribution, with averages of 50 and 75 mm. • Pointed butt blades are more elongated, displaying a bi-modal distribution with two peaks at 3 (as flat butt blades) and at 5. • Pointed butt blades were knapped mostly from unipolar cores, while flat butt blades were produced from other types of cores in a higher frequency. • Pointed butt blades were produced mainly on light brown flint, as opposed to the higher diversity of raw materials observed on flat butt blades.

Stage 2: Production of blanks The high number of waste products, of all types, indicate a local production of most tools. Even though all cores bear negatives of mostly flake removals at their last stages of exploitation, and among débitage, flakes dominate (Table 13), blades were chosen as the preferred blank (45% of all tools are on blades). The value of the ratio blades/(tools on blade) is 2, while a similar index, related to flake blanks exploitation is 17, even though the production of flakes was more intense than that of blades (for each blade, 6 flakes were knapped).

It may be concluded that pointed butt blades were produced mainly from unipolar light brown cores, this method being mostly used for the production of either small or long and elongated blades. Flat butt blades were produced from cores of various shapes and from a higher variety of raw materials, with less apparent care of size. Therefore, the possibility that a knapping technique resulting in a pointed butt was a determined method to produce blades of desired shapes cannot be rejected.

Flakes

A comparison between light brown blades and all the others revealed a similarity between blades on various raw materials. Therefore, the conclusion is that raw materials and blades production are independent variables, similar knapping methods and techniques being applied on all exploited raw material.

Flakes were produced on all processed raw materials at the site, in percentages following those of cores, light brown and grey being most common. Half of the flakes were clean of cortex, suggesting an extensive exploitation of cores, also reflected in the small size of flakes (38x29x8 mm).

The hypothesis that bladelets are products of a different reduction scheme than blades was tested. A comparison of technological attributes showed no marked difference between them. The only difference between flat butt bladelets and pointed butt bladelets was that the later tends to be longer and thinner. Therefore, it may be concluded that blades and bladelets were products of a similar reduction scheme along a continuum sequence, blades being produced at first stages and, as cores became smaller, bladelets were obtained.

Three main knapping methods and techniques were employed in the production of flakes, all in similar percentages: a direct percussion with a hard hammer, from a non-faceted core resulting in a flat butt, an indirect percussion, probably with a soft hammer, yielding a punctiform butt and the removal of flakes from a faceted core, apparently at later stages of removals. Blades and Bladelets

The last comparison was made between sickle-blades and blank blades/bladelets. Given the similarity of raw materials from which blank blades, bladelets and sickleblades were produced, it was assumed that at least some of the later were locally manufactured. A main difference regards their width: sickle-blades have a mean width of 13.6 mm., blank blades a mean width of 18.9 mm, while bladelets average around 9 mm. The distribution of the width values of these groups shows a preference for blanks between 12 and 19 mm to be modified into sickleblades. Consequently, blanks, either blades or bladelets,

A sample of 50 blades and 50 bladelets was analysed (Table 28). Their common raw material is light brown flint (on 2/3 of items). Other types of raw materials are grey, brown, cream and pink flint. The presence of blank blades and sickle-blades on similar raw materials suggests a local production of the later (see below). Blades were removed mainly from unipolar cores, after cleaning their débitage surface from cortex and performing a few knapping stages.

70

CHAPTER V – NORTHERN ISRAEL SITES

with widths of this range were separately examined. The result shows that they do not markedly differ from other blades or bladelets. It was concluded that the only criterion applied when choosing blanks to be modified into sickle-blades was their width, and there was not a particular production intended to obtain blades suitable for the production of sickle-blades.

The variety of borers' raw materials suggests that there was no preference for a particular type of flint for their production. This behaviour is further evidenced by the recognition of six types of blanks used for the production of borers: flakes (35%) , blades (27%), ridge blades and primary elements (15% each) and a few items on chunks or bifacial spalls.

A local production of sickle-blades can be described as applying a pre-determined blade knapping method, using unipolar light brown cores and a soft hammer. Blanks with a desired width (between 12 and 19 mm) were chosen and modified into sickle-blades.

Except for drills, their point extending along their length, the length of awls’ points vary between 15 mm. (4 items) to 3 mm. (all others). The high variability of borers reflect a vast range of activities performed with them. Retouched flakes (N = 19; 7%) Truncations (N = 12; 5%) Notches (N = 25; 9%)

Stage 3: Shaping/Retouching stage – retouching of tools

Retouched flakes, truncations and notches are on various types of raw materials and were modified by different retouches. The possibility that some were modified by natural agents cannot be excluded. Their small amount suggest that these agents had a minimal impact on the entire assemblage. The possibility that the industry was less oriented towards a mass production of ad hoc tools, a common phenomenon in Chalcolithic assemblages (cf. Table 3), Kaukab flint knappers preferring to produce more non ad hoc tools cannot be neglected.

The tool assemblage is dominated by sickle-blades and borers (22 and 21% respectively). Noticeable is the small amount of ad hoc tools, such as retouched flakes, truncations and notches/denticulated. The absence of discs, a common tool-type in Northern assemblages, is conspicuous. Most tools were locally made, exceptions being two tabular scrapers, possibly some microliths on semi-translucent flint and a few bifacials. The preferred blank was the blade (45% of tools); a 1/3 of all tools are on flake. The presence of tools on C.T.E. (N=9), and bifacial spalls (N=3) suggest a production of tools based on the principle of using any available blank, the initial shape and size of blank being of little importance.

Retouched blades (N = 14; 5%) Retouched blades are on raw materials similar to those of sickle-blades, and in similar frequencies, which suggests a common reduction sequence for both groups. The retouch morphology vary, including almost all possibilities: from fine to scaled, partial or complete, dorsal or alternate to bifacial. This high variability suggests a non-standardized production of retouched blades, each item being modified according to the blank’s shape and the function it fulfilled (the variability of retouch may reflect different hafting methods and/or different functions).

Scrapers (N=29; 11%) Scrapers (Plate 29:1,2), were made on most raw materials, with a preference for light-brown flint (half of scrapers). All blanks are flakes, primary elements (8 items) or simple flakes of various sizes, from small (1-2 cm.) to massive (5-7 cm.). The position of retouch varies from distal end, to edges and the circumference. The possibility that the extent of retouch reflects resharpening cannot be neglected. This may explain the presence of more rounded scrapers (N = 8), items at the end of their exploitation. Two broken tabular scrapers (Plate 29:3) complete the group.

Sickle-blades (N = 58; 22%) Sickle blades dominate the tool assemblage. They include 12 without gloss, but with all other technological and morphological characteristics.

Burins (N = 8; 3%)

Three types of raw materials were used for their production: light brown (half of the group), grey (17%) and cream (10%). The remaining items were burnt. Given the fact that cores of similar raw material were found at the site, the possibility that most sickle-blades were onsite produced cannot be rejected. This suggestion is supported by the fact that all items without gloss (unused sickle-blades) are made on light-brown flint, from which a large quantity of cores was collected at the site. The limited variety of raw materials used for the production of sickle-blades (3 of the 11) suggest a deliberate selection of raw materials for the manufacture of sickleblades.

Burins are on flakes or retouched flakes; two items are on blades. Most are on break (N = 5), the rest on truncations or are dihedral. Borers (N = 55; 21%) Borers are the next common tool-type after sickle-blades. Drills, all on narrow blades, are bilaterally backed, with the point extending along most, or all of their edges. They represent 15% of borers, the remaining being awls, with one or more short points (Plate 29:4); they were subdivided into simple (Plate 29:5), massive (on large chunks or flakes) or double (with two points).

71

CHAPTER V – NORTHERN ISRAEL SITES

30:6). The possibility that broken retouched bladelets are micro-endscrapers missing the tip, cannot be excluded.

The morphology of a typical sickle-blade (Table 15) is a narrow blade, with a denticulated working edge and two straight truncated ends (Plate 30:1-3). An oblique scar on the distal end of 4 items testifies for intentional snapping, reminding microburin technique (Plate 29:1). The back, unipolarly retouched, is in all but three cases straight, suggesting a straight haft of the sickle.

Bifacials (N = 26; 10%) The bifacial group comprises two sub-types: chisels (N = 17) and adzes (N = 9). Among the former, 2 items have two pointed ends and may represent a sub-group of chisels (Plate 31:1). Both chisels (Plate 29:6) and adzes (Plate 31:2) were manufactured in a similar mode: a pebble grey chunk was bifacially flaked from two striking platforms, producing an elongated tool with a planoconvex cross-section. The working edge was obtained by small bifacial removals. Only a few items were polished. The main difference between the two types is their plan view and shape of working edge: an elongated rectangle and narrow body and a narrow working edge for chisels, trapezoidal body and a wide working edge for adzes. These differences may denote different hafting methods and function.

The high standardization of sickle-blades is indicated by several observations: all items are backed, truncation and denticulation appear constantly, and their measurements show a low standard deviation. All items, regardless raw materials, have similar sizes and share the same morphological characteristics. The sickle blades production involved several stages: 1. knapping off a blade – probably from a blade core, as suggested by the parallel scars on the dorsal face of sickle-blades. Blanks were knapped at first removal stages from cores, after the core was cleaned from cortex. 2. snapping the ends of the blank by a micro-burin technique – this operation was performed in order to control the length of the item. 3. truncating the ends of the blank, by straight fine retouch, probably to stabilize the hafting process. 4. backing the blank by abrupt, straight and unipolar retouch, probably improving their immobility within the hafted sickle. 5. preparation of a fine denticulated working edge, possibly by pressure retouch.

Two additional raw materials were used: cream flint (five chisels), light brown flint (two chisels and one adze) and striped brown flint (one chisel). The small amount of items made on these raw material, all extensively used for the production of most other tools, may denote the original small size of raw materials chunks, unsuitable for the production of large tools. Two adzes, both complete pieces, were prepared from grey flint with micro-fossils, a raw material not used for the production of any other tool-types and absent from the waste group as well. The possibility that these adzes were brought on-site as end-products cannot be neglected.

Interesting to note is the presence of many burnt sickleblades (a quarter of all groups). Among all tool-types, only three (sickle-blades, microliths and scrapers) have burnt items (14, 3 and 1 respectively). Since sickle-blades were found in similar places with non-burnt items, this phenomenon needs an explanation, beyond post-discard processes. One possibility is that sickle-blades were burnt before being discarded, during a “sterilization” process of cleaning sickle-blades after cutting, from remains of plants. Another possibility is that sickle-blades were burnt during a process of retouching or hafting, being “errors” of extended exposure to high temperature.

Six chisels were shaped from three striking platforms, having a triangular cross-section: two along their edges and one at the mid-distance from edges, along the longitudinal axis of items. One adze, with a trapeze crosssection, was shaped from four striking platforms, corresponding to the trapeze’s corners. The different shaping methods were probably dictated by the original blanks’ shape these tools were prepared from. Four chisels have a working edge prepared by longitudinal removals. Three of them were collected from the same locus and the fourth one was found in a locus below it. The possibility that these items were made by one knapper, and the shaping method of the working edge reflect his/her personal preference cannot be neglected.

Microliths (N = 18; 7%) Microliths were prepared either on light-brown flint (2/3 of them), or on semi-translucent flint (1/4 of items), chosen probably for their quality, suitable for the production of bladelets. The presence of microliths (and bladelet cores) on semi-translucent flint denote a deliberate choice for this type of flint, a choice that may have been dictated by cultural/traditional factors.

Five bifacials (3 chisels and 2 adzes) were polished, all on the working edge solely. However, the presence of 12 bifacial spalls, all bearing polish (three of them being further modified into other tools – a notch, a retouched flake and an awl), may suggest that polish was more extensively used than it appeared on the collected bifacials.

Two main sub-types of microliths were found: microendscrapers (72%, Plate 30:4,5,7) and fine retouched bladelets (among them, one ventrally retouched, Plate

72

CHAPTER V – NORTHERN ISRAEL SITES

scrapers and few bifacials. Among the later, adzes and chisels dominate; axes and discs (common in northern assemblages) are absent.

A local production, use, re-use and discard of bifacials is evidenced by: 3 unfinished items, 12 bifacial spalls and the damaged working edges of celts. Only five of them are not broken. Probably complete pieces were either stored in an un-found location within the site, or taken by the inhabitants when they abandoned the site.

Most flint artefacts were found on floors or fills; a few were found in pits, while installations were void of flints. This picture suggests that inhabitants did not dispose of their waste in dumping zones prior to their abandonment of the site, leaving most artefacts probably where they were last used. This phenomenon may advance a short site occupation, its inhabitants avoiding repeated cleaning of their living areas.

Stage 5: Discard The spatial distribution of flint artefacts shows that most items were collected from fills or living floors, regardless their type; pits were not rarely used as dumping zones (they contained only 10% of flints) and the installations were not related to the lithic industry. Only 2% of all flints are a surface collection. It might be suggested that flint artefacts were not discarded into dumping zones, but were left behind by their users at the place of their last use. This may explain the high amount of tools, particularly sickle-blades, when compared to Northern Negev sites, where flints were discarded in dumping zones, whereas usable tools were hid or carried away when sites were abandoned (this behaviour may also explain the low amount of sickle-blades in northern or western Negev assemblages).

Tel Turmus (210.68/290.89) Description of the site The lithic assemblage originates from a salvage excavation conducted by the Israel Antiquities Authority during 1997. The site is in the Hula valley of Northern Israel (Map 2). Deep sections revealed the existence of several building stages, all of the Chalcolithic period. Rooms of different houses were excavated, some with ashy tightened floors. The original plan of rooms is similar to that of the Golan sites (Epstein 1998). Their walls were built on stone foundations, consisting of two rows of large boulders and small stone fillings. Installations, in the form of small depressions paved with basalt plates, were found on some floors. A few refuse pits were uncovered. Most of the household’s daily activities were probably performed in courtyards, found adjacent to buildings. Artefacts were collected using conventional methods of basket as the smallest excavation unit and locus for every human made feature.

Summary Despite the small excavation area, a large amount of lithic material was collected. The variety of raw materials used hint at several sources of procurement, most being used for the production of all tool-types. Tabular flint, flint with microfossils and most of the semi-translucent flint were not processed at the site; they appear only on tools: tabular scrapers, adzes and microliths, respectively.

Description of the flint assemblage

One-platform cores dominate, most being used for the production of nearly all flakes, though bladelets were also produced in high percentages. Apparently, knappers changed aims of cores during their exploitation, applying different techniques and methods, depending on the stage of removals and the size of cores: at later stages, cores were faceted and a soft hammer was used for the production of blades/bladelets. Thus, a few cores were exclusively devoted for the production of a single blank type, most retaining scars of flakes and bladelets, flakes and blades and a few blades and bladelets. Blade, bladelet or flake cores are rare.

The results of the analysis of the flint assemblage were published elsewhere (Marder and Hermon, in prep.). 4252 flint artefacts were collected and analysed. Debris (mostly chips) represent more than half of all flints. Débitage, half of which are flakes, represents 1/4 of flints (Table 5). All other waste products, such as cores, primary elements and blades are present, suggesting that, at least to some extent, the flint industry was produced on-site. Tools (Table 2) represent ca. 15 % of the flint assemblage. Given the fact that most tools were discarded in a broken condition, and the apparent organized abandonment of the site, the original number of tools produced and used within the site was probably higher.

Even though flakes dominate the waste products, the preferred blank type was the blade: sickle-blades and borers on blades dominate the tool assemblage. Sickleblades retain most of the morphological characteristics of Northern Negev items: narrow, double truncated, backed and with a fine denticulated working edge. Among the gamma of exploited raw materials, only a few were used for sickle-blades production. Their high standardization suggests that sickle-blades were produced in a workshop, even though no remains of such were noted. Most tools were probably produced in each household, excluding some microliths (on semi-translucent flint), tabular

Tools on blades represent a 1/3 of the group. Both non and ad hoc tools appear in similar amounts. Non ad hoc tools (excluding bifacials, made on chunks) are on blades or bladelets, while ad hoc tools are on flakes. Sickleblades, retouched blades, bifacials and retouched flakes form the majority of tools, each in similar percentages (ca. 14% each). Perforated discs, typical of northern assemblages (Noy 1998) are well represented (Table 2).

73

CHAPTER V – NORTHERN ISRAEL SITES

bladelets. This difference is related to various modes of cores exploitation, related to raw material type: light brown cores were reduced using a blade scheme of removals, while grey flint cores were used at an earlier stage of blades production and later for the production of flakes. It is also possible that the deliberate selection of light brown flint and semi-translucent flint influenced the reduction scheme, these types being used mainly for the production of blades. Another factor may be the size of nodules, which enabled an extensive exploitation of grey cores, as opposed to light brown and semi-translucent cores. Another possibility, though less plausible, is that blanks of semi-translucent flint were taken off-site and used elsewhere.

A particular feature of the flint assemblage, is the pièce esquille (Bar-Yosef (1970); several of them were retouched and modified to various tool-types, while others were left unmodified. The possibility that the ventral modification which defines pièce esquille is intentional cannot be ignored. Stage 0: Acquisition – Extraction and testing of nodules Several raw material types were exploited: pebble grey flint, brown nodule of various shades of flint, dark brown flint with white micro-fossils, semi-translucent small nodule flint and “canaanean” flint, typical of the EB (Figure 3). Grey, brown and light brown flint were extensively used, each on ca. 1/4 of the flints (Table 1). The use pattern of each raw material type is different: most tools are on dark brown and grey flint, while most waste products are on light brown flint. This difference may reflect the characteristics of light brown flint nodules, requiring more preparation before knapping off a desired blank, while dark brown and grey flints were more easily knapped. The possibility that some tools were brought to the site as end-products cannot be excluded. The source of most raw materials is unknown, though several are expected: outcrops in the northern mountain regions or adjacent wadis and low quality flint pebbles dislocated from the western parts of the Golan plateau and rolled over a long distance. Other flint sources, located at farther distances, such as the Galilee mountains, were exploited for “canaanean” or semitranslucent flint.

Stage 1: Production 93 cores were collected, (2% of the assemblage, 9% of débitage) (Table 1). The common blank was the nodule; a high number of flakes were used as cores as well (16% of cores), which suggest an intensive exploitation of available blanks. This is further supported by the fact that most cores have multiple platforms, oriented in various directions; ca. a ¼ of them are faceted, which probably increased the efficiency of their exploitation. The débitage surface of cores is usually clean of cortex. This observation supports the previous assumption regarding the intensity of cores’ exploitation. The number of scars varies from 4 to 23, common being 11, reflecting a high degree of cores exploitation (Table 16). The low quality of flint, which required an intensive elaboration prior to a systematic exploitation from one hand, and the scarcity of other types of flint, brought from farther locations, imposed this intensive core exploitation.

Most cores were prepared from grey, light brown and brown flint. A main difference between raw materials of flakes and blades is the presence of ca. 5% of blades on semi-translucent and “canaanean” flint, while no flakes on these types of flint were found (Figure 3). This picture is well reflected also in the raw materials of tool types: both types of flint were used almost exclusively for the production of tools on blades (Figure 4).

More than half of the cores were used for the production of several types of blanks, each blank type corresponding to a different removal stage. Among the specific oriented cores, ca. 1/3 were used to produce flakes (Plate 32:1,2), 14% bladelets (Plate 32:4,5) and only 3% blades (Plate 32:3). However, when considering the possibility that the aim of cores changed according to exploitation stages, a different picture emerges: while 80% of cores were used to produce flakes, a similar amount of cores was used to knapp off blades and bladelets (Table 16). This estimation was obtained by counting the number of cores with flakes, blades and bladelets removals, respectively. Thus, it seems that most tools were produced on-site.

More than half of the bifacials are on grey pebble flint; other raw materials were used in similar, low amounts of less than 10%, except brown flint (20%). Semitranslucent flint and “canaanean” flint were not used to produce bifacials. Discs are on all but semi-translucent and “canaanean” flint (Figure 3). Almost half of the borers are on grey flint; ca. a ¼ of them are on brown flint. The remaining raw material types were used in insignificant numbers. Scrapers were mainly made on brown and grey flint (Figure 4).

The descriptive statistics of cores are presented in Table 16. Their mean size is similar to that of flakes; the ratio Length/Width and the shape of removal scars indicate that only three cores are typical blade cores, the rest having a globular to cylindrical shape. This shape is also reflected in the normal distribution of the ratio maximum width / diameter of main striking platform. The mean size of cores was roughly calculated ((Length + Width)/2) – its distribution reflects two groups of cores, with average

Blades, retouched blades and sickle-blades were made on three main types of flint: brown, grey and light brown (Figure 4). Most retouched blades and sickle-blades are on grey flint (40% and 28% respectively), while blank blades are on light brown flint. Semi-translucent flint was mainly used for the production of retouched blades and

74

CHAPTER V – NORTHERN ISRAEL SITES

analysis of cores as well. Almost half of flakes have a centripetal scar pattern, a 1/3 unipolar pattern, and the remaining flakes a bipolar scar pattern. These observations accord with the analysis of cores, which shows that multi-platform cores were mainly used for flake production.

sizes of 29 and 40 (Figure 5). The smaller group of cores was mainly used for the production of bladelets. A high amount of core trimming elements (CTE) was collected (Table 13). Among them, all types appear in similar amounts. The ratio number of cores / number of CTE shows that more than two thirds of the cores were renewed. The presence of several sub-types of CTE supports the previous assumptions regarding the high degree of exploitation of cores and the change of cores’ aim at different removal stages.

The descriptive statistics of flakes size is presented in Table 27. The distribution of their length and width is normal, (mean length = 34 mm, mean width = 27 mm). The distribution of the ratio Length/Width shows two groups (Figure 5): the smaller are transversal flakes, probably removed from centripetal cores, and a larger ones, with a square shape. Since most flakes have a square shape, the distribution of the ratio (Length + Width)/2 was calculated. The result displays a normal distribution (Figure 5), in which a 1/3 of items fall within 30% of the average of 28. The size of flakes corresponds to the length of cores.

The decortication of nodules was performed probably onsite; one platform was prepared at a preliminary stage of removals. Further, cores were renewed by CTE, and changed their aim, knappers applying various reduction schemes. Most cores retained little or no cortex on their surface, most of them being exhausted. The high intensity of cores exploitation, improved by faceting striking platforms, is reflected also in the small amount of cores, related to other waste products and tools.

Blades Blades represent a 1/5 of débitage (Table 13). The ratio flakes/blades reflects the tendency of the flint industry to be flake oriented, also exemplified by the low amount of cores that were exclusively used to produce blades. However, the ratio tools on flakes/tools on blades shows a preference for blade blanks for the production of tools. The low amount of blank blades can be explained in terms of their high exploitation, and not necessarily the result of low scale production. Moreover, 2/3 of cores display negatives of blades and bladelets removals on their débitage surface. Therefore, the possibility that most tools on blades were produced on-site cannot be excluded.

Stage 2: Production of blanks The ratio primary elements/cores (0.86) is small, suggesting that only a few primary elements were discarded, without being further modified into tools, while many were used as blanks for ad hoc tools. This concords with the high exploitation of blanks and the small amount of primary elements collected (N=80), representing 7% of the débitage and 2% of all items (Table 13). The amount of blanks (primary elements, flakes and blades) related to the number of cores (9.81) reflects the high exploitation of cores, well reflected in other characteristics of the flint assemblage. Adding tools to this ratio, an estimated number of 17 blanks were knapped off each core. This number may reflect the exploitation degree of cores only if assuming that all tools were produced and discarded within the area of the site, and collected during its excavation. The high exploitation of flint products is also reflected in the ratio (primary elements + flakes + blades + tools)/tools (2.36), suggesting that each second blank was modified and used as a tool. This high exploitation is also reflected in the high amount of rejuvenated tools, and tools that changed aim. Another example is the case of broken bifacials, used as cores.

More than half of the blades have a trapeze cross-section, with parallel, unipolar scars of previous removals on their dorsal face (Table 28). Other cross-sections are triangular (a 1/3 of blades) or rectangular. Several butt-types were observed, common being flat, on more than 1/2 of blades. A fifth have a faceted butt. Occasionally, punctiform, linear or dihedral butt was used. Most blades have 5 scars or more on their dorsal face (Table 28). This high number suggest an extensive exploitation of cores, a conclusion supported by observations on flakes and cores as well. The Length/Width distribution of blades shows that a 1/3 of them are at the limit of their definition as such, with a ratio of 2.3. The possibility that some are elongated flakes cannot be excluded. Another group consist of blades with a Length/Width ratio of 3.5. Probably most of them resulted from a blade-oriented production scheme. The width distribution of blades shows two groups: a 1/5 of blades, with a mean of 15mm, the other one with a mean of 23 mm (a ¼ of the sample). Probably these two groups reflect two different reduction schemes, according to the end-products required, the first being mainly used for the production of microliths.

Flakes A sample of 100 flakes was analysed (Table 27). Several butt types were observed: flat (1/2 of flakes), faceted (1/3 of flakes), and dihedral, linear or punctiform. This picture reflects several knapping techniques, using either a soft or a hard hammer. The number of scars on the dorsal face of flakes varies between 1 to 12, most common being 5 scars on each flake. This high number is probably the result of an extensive use of cores, as indicated by the

75

CHAPTER V – NORTHERN ISRAEL SITES

scrapers corresponds to that of blank flakes, which supports the suggestion that scrapers were locally produced, from suitable flakes.

Stage 3: Shaping/Retouching stage – retouching of tools 667 tools were collected (Table 3). They are equally divided between non and ad hoc tools.

Borers (N=60; 9%) Borers represent 1/5 of the ad hoc tools. Ca. 1/3 are drills (Plate 33:5). One awl (Plate 33:6) is on burin spall; 3 are on CTE. The flakes chosen for the manufacture of borers are similar to those of scrapers and blank flakes, indicating no particular choice in their selection.

Retouched flakes (N=96; 14%) and notches (N=78; 12%) Retouched flakes were simply modified by a partial fine retouch along one of their edges, while notches are simple flakes with two or more retouched notches along one edge. These are the most common tool-type (Table 3). Their high amount may reflect some daily, or frequent and repetitive activities, but also their frequent rate of preparation, use and discard. It should be also remembered that an unknown amount of tools were taken when the site was abandoned, since most of the left behind tools were found broken. The high variability of blank morphology indicate that there was not a predetermined scheme of production of retouched flakes, but rather an opportunistic one, available flakes being modified into tools, upon necessity.

Most borers are on blanks without cortex; 3 items are on primary elements. Drills were modified by backing along their edges; awls were modified by fine retouch (40%) or by the retouch of two notches (12 items). Two awls have two points. More than 2/3 of borers were used for drilling holes with a mean diameter of 7.5 mm; another group (12%), was used to drill holes with a mean diameter of 22 mm. Since these sizes do not match with the average diameter of perforated discs' holes and no other holed remains of material culture have been found, the use of borers is unknown, being probably used on perishable material, such as wood or leather.

Burins (N=26; 4%) Burins on truncation (Plate 38:5) dominate (N=11), followed by burins on break (N=9), and dihedral burins. All but 4 are on flakes. Burins are rare in Chalcolithic assemblages of the Negev (Gilead 1995). An important aspect of their production is that nine are on discarded non ad hoc tools, such as discs (N = 6), a chisel (Plate 38:6) and a sickle-blade. The number of burin spalls (N = 17), and their nature indicate their on-site production.

Retouched Blades and Microliths Retouched blades and bladelets represent almost 1/3 of all tools (Table 3). Among them, artefacts with gloss were defined as sickle-blades. Given their high morphological variability, a characteristic sickle-blade could not be defined. Retouched blades have a regular retouch along one or both edges. Only a few microliths (N=7) were found. This suggests a limited use within the excavated area. Given the high amount of cores used for the production of bladelets (see above), the possibility that a certain number of bladelets, either as blanks or as microliths, was taken off-site cannot be excluded.

Truncations (N=24; 4%) 2/3 of truncations are on flakes, the rest on blades. Their distal end was modified by fine retouch, forming a concave, convex, oblique or straight end. Truncations represent 7% of the ad hoc tools.

Retouched Blades (N=90; 14%)

Scrapers (N=37; 6%)

Three groups were defined: fine retouched (60%), retouched and backed (15%, Plate 34:1) and backed (Plate 34:2-5), occasionally truncated. The possibility that some backed items are non-used sickle-blades cannot be ruled out. The assumption that backing improves hafting implies that backed blades were involved in a cutting activity on a hard material, items being hafted together. If such is the case, the presence/absence of the back reflects different activities performed with retouched / backed blades. However, the possibility that backing is the result of knappers’ choice cannot be excluded. This hypothesis is supported by the fact that gloss was found on the working edges of both backed and non-backed blades.

Scrapers (Plate 33:2,3) represent 12% of the ad hoc tools. Given their low number, any statistical conclusions have to be taken with the required restrictions. This low relative amount of scrapers is worth noticing, especially when comparing with Southern Chalcolithic assemblages. The preferred blank was the flake; a few are on blade. Most have no, or few cortex on their dorsal face, which suggests that blanks were knapped from cores at a later removal stage. 3% of scrapers are on primary elements. Ca. half are end scrapers, 1/4 are rounded and less than 1/5 are side-scrapers. The remaining scrapers are retouched along one edge and the distal end. It is possible that the shapes of scrapers reflect different modes of their use and/or degree of re-sharpening. The mean size of

The delineation of retouch is partial on more than 2/3 of items. Almost 1/2 were modified along the left edge, a 1/3 along the right edge, the rest being double retouched.

76

CHAPTER V – NORTHERN ISRAEL SITES

On 2/3 of sickle-blades, the gloss is parallel to the working edge, extending along their entire length. The remaining items have a partial gloss. The delineation and extension of gloss is influenced by the position of items within the sickle, and its shape.

A few items were truncated, by one straight truncation (a ¼ of items, Plate 34:2,3) or double truncation (16%). Apparently, the length of items was not important, given the fact that only a few were truncated. The mean number of scars on the dorsal face of retouched blades is 4 (Table 17). Their pattern is usually unipolar. A few others were have a bipolar pattern and only 2% an amorphous one. This observation suggests that blades were knapped at first stages of cores exploitation, since most cores have multiple platforms (see above). Moreover, a deliberate scheme of reduction was applied for the production of blank blades.

These observations suggest that sickles used at Tel Turmus were curved, the cutting edge consisting of probably no more than 3 sickle-blades. If this reconstruction is correct, the number of sickle-blades collected would fit into 30 sickles, which may indicate the number of households living at the site. The mean size of sickle-blades are presented in Table 18). The comparison between the mean size of blank blades, sickle-blades and retouched blades shows a their similarity, indicating that on-site produced blanks were further modified into retouched blades and sickle-blades. The variability in sizes and shapes of sickle-blades suggest their non-standardized mode of manufacture.

The descriptive statistics of retouched blades’ size is presented in Table 17. Their mean size is similar to that of blank blades and sickle-blades. Moreover, the width distribution of blades, sickle-blades and retouched blades shows that all belong to one population, with a mean width of 17 mm. It may be suggested that most blade tools were produced on-site, and that retouched blades and sickle-blades are products of the same scheme of reduction.

Bifacials (N=106; 16%) These are adzes (N = 39, Plate 36:2; 37:1), chisels (N = 32, Plate 36:1), perforated discs (N = 32, Plate 37:2) and axes (N = 3).

Sickle-Blades (N=89; 14%) Sickle-blades represent almost half of the retouched blades. They are bi-truncated and backed, with one working edge finely denticulated (Plate 35:6,7). However, other shapes appear as well. Almost half the sickle-blades have a damaged/abraded working edge. Apparently, denticulation improves the qualities of a working edge; the presence of sickle blades with and without a denticulated working edge suggests a nonstandardized production of these items (different knappers, without a crystallized tradition). This suggestion is supported by variations observed in sickleblades’ shapes (Plate 35). A particular group consists of sickle-blades on apparently “canaanean” blades, with faceted butts (Plate 35:2-4).

Adzes and Chisels The main difference between adzes and chisels is the shape of working edge (straight for adzes and curved for chisels) and plan views (straight rectangular for chisels and elongated trapeze for adzes). This difference is probably functional, which influenced their shape and hafting. The amount of cortex on both types is minimal (Table 19). While up to 90% of chisels are void of cortex, half of the adzes retain some cortex. It is possible that this difference denotes a higher investment in the production of chisels.

Most sickle-blades were truncated (75%). Moreover, most complete items were bi-truncated, which emphasize the importance of a required length, part of a composite tool.

The blank type chosen for the preparation of these bifacials was a chunk or a flake. Here again, a difference is observed: most adzes are on chunks (more than 90%), while only 2/3 of chisels are on chunks. Apparently chunks were more adequate for the production of adzes, a massive tool-type.

More than 1/2 of sickle-blades were backed, probably to improve their hafting in the handle of the sickle. Artefacts that were not backed were probably hafted in a different way, or their width did not allow backing.

The cross-section shape of adzes and chisels depend on the number of striking platforms from which they were produced, and their orientation. Items with isoscelesangle triangle cross-section or a straight-angle triangle cross-section were shaped from 3 striking platforms, each point of the triangle representing a striking platform. Items with a trapeze cross-section were shaped from two striking platforms, located along their edges. More than 2/3 of adzes have a trapeze cross-section, the rest being an isosceles-angle triangle cross-section, (a ¼ of items) or a straight-angle triangle cross-section.

The mean number of scars on the dorsal face of sickleblades varies from 1 to 6, the average being 2 to 3 scars on each. Probably, as in the case of retouched blades, the blanks that were used for the production of sickle-blades were produced at the first stages of removals from cores. The pattern of scars is usually unipolar, suggesting that these artefacts were removed from unipolar cores.

77

CHAPTER V – NORTHERN ISRAEL SITES

Perforated tools were probably produced on large flakes, even though the bifacial retouch removed any signs of bulb of percussion and all the artefacts were found broken. No cores with a débitage surface large enough to produce blanks for perforated discs was found. Their similar raw materials with other blanks raises the possibility that some cores were first used to produce discs, and later on modified, in order to produce other blank types. So far, no site for the production of perforated discs was reported.

The analysis of chisels presents a different picture: 2/3 of items have an isosceles-angle triangle cross-section, and only a 1/3 a trapeze cross-section. This difference can be explained in technological terms, chisels being mostly shaped from three striking platforms, while adzes were mainly modified from two striking platforms (Table 19). A possible explanation for this difference is the type of hafting, which was probably dictated by users’ gestures using these artefacts. Two main shapes of working edges were noted, probably reflecting different activities performed with these tools. Adzes have a straight working edge, while chisels have mostly a curved working edge (Table 19). Occasionally, the working edge of adzes was renewed. The presence of 8bifacial spalls with polish remains further supports the suggestion that bifacials were resharpened.

The diameter of discs was estimated using a table for diameter measurements of pottery, assuming that their original shape was rounded. Its distribution shows two distinct groups of discs: one smaller, with a mean diameter of ca. 10 cm, and one larger than 25 cm (Figure 6). The same division is valid for thickness: 1/2 have a mean thickness of 7.4 mm, while the rest are thicker (ca. 11 mm). The same is true for the distribution of the central hole: 13 mm, against 16 mm. It might be concluded that discs were produced in two groups, one smaller than the other. While this difference reflects different uses is unknown, given the fact that no use wear analysis was performed on this type of tool.

Since more adzes than chisels (2/3 adzes against 1/3 chisels) were polished, it is plausible to assume that mostly adzes were renewed, given their probable higher rate of deterioration, as a result of a rough work. More adzes than chisels were found with a damaged working edge, suggesting that adzes were used on a harder material than chisels. The presence of more adzes with polish, and their higher rate of deterioration may suggest that polish was applied to improve the qualities of the working edge, increasing its durability.

Tabular Scrapers (N=22; 3%) Half of the tabular scrapers are broken. Among the complete ones, 6 are rectangular (Plate 33:1), 3 have an oval shape (Plate 33:2) and only one can be described as a typical “fan” scraper. Two items have a faceted butt, in one case it was thinned. Since these items were produced on a characteristic “tabular” flint, is absent from the waste products, it was assumed that they were brought on-site as end products.

A common preparation of the working edge was knapping several longitudinal removals (2/3 of adzes and 3/4 of chisels), or one large removal. Occasionally, the working edge of adzes was obtained by a transversal removal. The mean size of chisels and adzes are presented in Table 19. It seems that blanks of similar size were chosen for the manufacture of both types. Apparently tool size was less important than size of the working edge, as reflected by the standard deviation of various measurements (Table 19).

Pièces Esquilles (N=32; 5%) One of the characteristics of the flint industry is the presence of pièces esquilles (Table 3). According to BarYosef (1970), these should be considered as waste products, although, apparently, the ventral modification could be post removal of items from the core. Bar-Yosef suggested that this modification can be the result of a particular way of knapping, and thus non-intentional (Bar-Yosef 1970). However, the presence of several items with multiple scars of removals from various directions on their ventral face suggest that these modifications are intentional, and thus these artefacts should be regarded as tools (Plate 38:1,2).

Perforated discs Perforated discs (Table 3) were found only in Chalcolithic assemblages, restricted mostly to sites in the Northern Israel (see below). A few discs were found in Teleilat Ghassul (Lee 1973), two items were reported from the Be’er-Sheva site of Safadi and one from the Besor area. The importance of this tool-type as a cultural / regional and/or chronological hallmark should be highlighted. Other important aspects are their complex mode of manufacture and unknown use.

60 items with ventral removals were found, half of them being further modified into other tools, such as various ad hoc tools or sickle-blades. Except bifacials, pièces esquilles appear on every tool-type, in similar amounts of up to 5 items for each type. Half of them were modified to other tool types. In this case, the ventral modifications should not be regarded as the results of a particular method of knapping, or an intentional thinning/removal of the bulb of percussion, but rather as the modification

All discs are broken; a 1/3 were resharpened into: 6 burins, 2 borers, 2 scrapers (Plate 38:4) and one notch. This observation exemplifies the blanks' exploitation strategy and the “opportunistic” approach to ad hoc tool production.

78

CHAPTER V – NORTHERN ISRAEL SITES

bladelets. Several are double tools: end-scrapers on retouched blades, burins on awls or burins on retouched blades. Most tools were discarded after they went out of use, after being used as blanks to produce other tools, or being completely abraded. This suggestion is supported by the fact that most tools were found broken or with damaged working edges.

that defined them as pièces esquilles. Therefore, the presence of these items should be explained as a characteristic of the flint industry of Tel Turmus. Summary More than half of the assemblage consists of debris. The débitage (1/4 of the assemblage), is distinguished by the presence of all waste products characteristic of an on-site production of artefacts. Several types of raw materials were used, some being deliberately chosen for the production of distinct tool-types, such as semi-translucent flint for retouched blades or grey flint for bifacials. Other flint types were used in a similar mode, variations being probably influenced by the original size of nodules, their quality and source location.

The high degree of raw material exploitation, well reflected by the exhausted cores, used for various aims, and by the high amount of resharpened and double tools may be influenced by several factors: 1. 2. 3. 4.

The flint industry was oriented toward the production of flakes. However, there are more blade tools than flake tools, indicating that a high amount of flakes does not necessarily reflect the need for flake blanks. Probably some flake cores, as defined at their last stages of exploitation, were first used to produce blades/bladelets. This is supported by the various aims of cores, as reflected by negatives of removals on their débitage surface. Cores were extensively used, being in most cases exhausted. This high degree of exploitation is best reflected by their multiple striking platforms, the lack of cortex on their surface and their small size. Other aspects that further support this suggestion are the presence of a relatively high amount of cores on flakes and the high degree of resharpening among tools.

restrictions imposed by a scarcity of raw materials in the vicinity of the site a distinct strategy of raw material procurement a limited number of knappers traditional knapping behaviour

Since the site is located in an environment where raw materials are available in the surroundings of the site, and same exploitation strategies were applied on all types of raw materials, apparently factors 1 and 2 cannot explain the observed phenomenon. The low standardization level of tools production suggests that it was performed in each household, and specialization, if existed, was limited to particular tools types. Therefore, the economic mode of the flint industry may have been influenced by cultural factors, such as maintaining a particular tradition of knapping, using any blanks available and re-utilization of tools, even though raw materials were accessible in the vicinity of the site. Various features were defined during the excavation. The analysis of the spatial distribution of flint artefacts revealed no distinguishable concentration of flint artefacts, which could be related to any particular feature, most artefacts being disposed in dumping zones.

Tools represent 16% of the flint assemblage; among them, ad hoc tools form half the assemblage, retouched blades (including sickle-blades) a third of the assemblage, the rest being bifacials (16%) and a small amount of pièces esquilles. One of the characteristics of this assemblage is the high amount of tools on blades, especially retouched blades and sickle-blades. Among bifacials, adzes and chisels dominate, while axes and picks are almost or totally absent. The presence of perforated discs is notable, being a characteristic of Northern flint assemblages. Another tool-type worth noticing is the pièce esquille, represented by items with a ventral modification, apparently intentional, and not the result of a particular striking method. Scrapers appear in low quantities, if compared to Northern Negev assemblages.

Neve Ur (ca. 202/232) The site was discovered in the Jordan valley, 14 km south of Tiberias Lake (Perrot et al. 1967). According to the surface distribution of artefacts, apparently displaying a cultural homogeneity, the site extended over an area of 23 ha. (Perrot et al. 1967:201). Due to security reasons, the site was not excavated systematically, and only surface collections were available for study. The flint industry is compounded by bifacials (picks, resembling Lower Palaeolithic trihedrals, axes, adzes and chisels), scrapers, borers and a special group of tools, perforated tools (pièces perforées). This group can be divided into four sub-types: discs, denticulated, side-scrapers and unfinished pieces (Perrot et al. 1967:203). The former type was probably made on flakes obtained by a predetermined, Levallois – like technique, with faceted striking platform and an occasionally thinned bulb of percussion. The items were perforated by percussion, rectifications by flaking and polish (Perrot et al. 1967:204). Similar pieces were found in small amounts at Ghassul, Be’er-Sheva and some Wadi Ghazzeh sites, and in higher quantities at sites in the Jordanian plateau

The knappers of Tel Turmus applied an apparently economic blank exploitation, using any available support for the production of tools, waste products and discarded tools as well. This exploitation is reflected in the high number of resharpened tools: ca. 14% of tools were made on worn/damaged/out-of-use tools. The tendency was to use non ad hoc tools, such as sickle-blades, adzes and perforated discs to produce ad hoc tools, such as scrapers, burins or awls (Plate 38:3). Occasionally, broken adzes and chisels were used as cores, mainly to produce

79

CHAPTER V – NORTHERN ISRAEL SITES

Noy, “…some 5000 tools…” are mentioned (Noy 1998:297).

(Perrot et al. 1967:205). The similarity in the ceramic and flint components of the material culture with Ghassul and Be’er-Sheva sites suggest a contemporaneity between Neve Ur and Teleilat Ghassul and the Be’er-Sheva sites, representing a “nouvel aspect du Ghassoulien” (Perrot et al. 1967:227).

Stage 0: Acquisition – extraction and testing of nodules Different raw materials were used for the production of tools; the majority are pebbles with dark cortex, mostly dark brown flint, sometimes a grey-veined flint and a few light-colour flint. Pebbles are worn by friction; some are smooth and shiny. Their origin is unknown, though a conglomerate stratum is susceptible. A different type of flint is a pebble with light-colour cortex and light brown texture, of a higher quality than the former. Occasionally, a semi-translucent flint was used as well. It is probably of Eocene strata, locally found under basalt cover. Most flake tools are on dark brown flint. Sickle-blades are on grey-veined or light-coloured flint, but not on dark brown flint. Bifacials were made mostly on Eocene flint with a chalky orange cortex. Perforated tools and tabular scrapers are on high quality Eocene flint, of an unknown origin. Perforated discs were also made on flint with small nomolithic fossil accretions. It was not identified in the Golan area, but in the Eocene formations of Upper Galilee, such as Zor’a and Adullam.

Golan sites (ca. 221.55/253.35) Following is a summary of Epstein's (1998) work, concerning sites discovered, excavated and published by the afore mentioned author, with emphasis on the lithic assemblages, analysed and published by Noy (1998). The sites are located close to the north-eastern shores of the Tiberias Lake, on the southern parts of the Golan height (Map 2), at an elevation of 400 to 600 masl. 25 sites were discovered and ten excavated. Sites are concentrations of dwelling structures, such as Site 12, farmsteads, such as Site 23, or isolated houses, found in close vicinity to the large sites. Sites were clustered around water sources and agricultural and grazing fields. The number of dwellings varies between a few (Sites 6, 15 and 19), hamlets up to 15 houses (Sites 14, 18, 21 and 22) and larger concentrations of structures, up to 37 houses at one site (Sites 12 and 20).

Stage 1: Production

All dwellings are of the broad-house type: rectangular, the long walls are oriented along the downward incline of the slopes in an east-west direction. The entrance was in the long south wall and the floors were paved with basalt stones. The dry-stone walls were made from un-worked basalt stones. Houses were occasionally grouped in a chain of several houses in a row, with joining walls. There is no evidence of planning; moreover, the various orientation of structures indicate different occupation periods. Silos, built of basalt stones, were found close to some houses. No evidence of metallurgical or ivory activities were found. Characteristic artefacts of the Golan sites are basalt pillar figures and rope-decorated reddish-brown pottery.

A large amount of primary elements were retrieved from houses at all sites. The presence of isolated items on nomolithic flint, characteristic of perforated discs, hints at their local production, or modification / rejuvenation. Flakes dominate the débitage; their mean size are 3-4 cm length and 2.6-3.7 cm width. A distinct characteristic is their large flat butt, sometimes 2 cm wide. Only a few blades were found, in concordance with the almost total absence of blade cores. Their raw material differs from that of sickle-blades. Bladelets are rare; given the small amount of microliths, their low number was expected. Cores Cores of conglomerate origin have a globular, pebble-like shape. They are similar at all sites, originating probably from a common source. They were divided into three groups: small (< 5 cm), medium (5 to 7 cm) and large (> 11 cm). Apparently, their size depended on their exploitation degree, most having two striking platforms, while large cores have usually only one. A high degree of core exploitation is suggested by the number of core trimming elements, of all types, collected at all sites.

Description of the flint assemblage The lithic assemblage, consisting of ca. 8500 artefacts, 1/10 of them being tools, was collected from 20 sites, mostly from excavated structures. A local manufacture of tools is evidenced by the high amount of waste products, of all production stages. Some problems were encountered while summarizing the report on the Golan sites (Noy 1998),such as lack of primary data concerning absolute numbers of flint types and inventory lists of each house/site. The only quantitative data was within tables, with the occasional obscure note of “sample”, without any reference to what was sampled. Not all tool-types were represented in these tables; moreover, a sum of all published accounts of tools reached the number of 1024, while in the summary of

According to published drawings, bladelets were knapped from small cores, larger ones being used to produce flakes. The possibility that small cores were used for the production of flakes and/or blades at preliminary stages of removal cannot be excluded. When cores from completely excavated houses are compared, small ones dominate (60%), a ¼ are of medium size and 15% are large cores. Apparently, core sizes are independent to

80

CHAPTER V – NORTHERN ISRAEL SITES

their amount at each site, their distribution within houses, or site size and nature.

Tabular scrapers A few tabular scrapers were found; most are transversal flakes. Width is 4 to 18 cm., length 3 to 9 cm and thickness 0.4 to 1.3 cm. Most items retained cortex; one item has irregular striation marks on it. The common shape is elliptic; retouch is parallel, sometimes scaled, extending over one edge. The bulb of percussion was thinned. Usually, each house contained one item; a few had more (5 at house 8, site 18). A distinct feature of a ¼ of tabular scrapers is a small perforation at their centre, ca. 0.25 cm in diameter. Some items have sheen along edges, others have a smoothed side. The variability of shapes indicates that tabular scrapers were used for different tasks.

The picture presented above is not easy to interpret, given the high variation that appeared, almost each site “behaving” in a different way. Assuming that the length criterion defines true types of cores, reflecting a technological meaning, and/or the intensity of exploitation, then the differences observed may reflect non-standardized, non-specialized production of tools, knappers using various knapping methods and using cores at different exploitation intensities. The apparent diversity of the lithic industry discords the uniformity of the material culture noticed. Therefore, the possibility that the picture presented above is influenced by a weak definition of core-types, by collection methods and extent of excavation, cannot be excluded.

Borers Borers (ca. a 1/5 of tools) are on dark brown pebble flakes and occasionally chalky flint. Some are on primary elements. Almost a ¼ are on rejuvenated tools: 14 on retouched flakes, 7 on scrapers and 1 on sickleblade. Given the high amount of flakes, the intensity of rejuvenation is surprising. However, the high exploitation of cores, suggesting a scarcity of raw materials, accords with the apparent economy of end-products and their high degree of use and re-use. The point of borers was obtained by the retouch of two notches or by double abrupt retouch along edges, in the case of drills. Awls dominate; drills, made on blade, are few, while microdrills are virtually absent. Usually one borer was found at each house. Exception is site 12, where 9 and 11 borers were found in houses 7 and 15 respectively.

The ratio tools/cores is 1.75 (each core was used for the production of 2 tools (!)). Excluding a single case, house 10 at site 12, where this ratio is 6.25, all others have a ratio of 1 (53% of houses), 2 (42%) or 3 (5%). This picture is surprising, since most cores were exhausted, and ca. a third of tools (bifacials, perforated tools and sickle-blades) were apparently off-site produced (Noy 1998). This picture may be distorted by two factors: one concerns behavioural aspects of Chalcolithic inhabitants, who possibly left their sites with usable tools, or discarded tools outside the excavated area; the other factor concerns collection methods, extent of excavations and analysis of lithic material. Core-choppers

Sickle-blades

A high amount of choppers was collected, from each excavated site, varying between 1 to 10 in each house, most common being 3 choppers per house, without any particular concentration. They are mainly on dark brown flint. Few have battered butts; some have a damaged working edge. Usually, a few removals were bifacially knapped, either for obtaining a cutting edge, or to obtain blanks for the production of flake tools.

Sickle-blades were found in most houses. Most are on Eocene flint; a distinct group consists of sickle-blades on brown-orange flint – 2 found in Houses 12 and R at site 12, and 10 at site 6. A blade core of identical raw material was found at site 12, House 10. Some items were made on primary elements, others on ridges, but mostly on blades free of cortex. According to drawings, most blanks were knapped from blade, unipolar or bipolar cores.

Stage 3: Shaping/Retouching stage – retouching of tools

The width of sickle-blades is 1.1 to 1.8 cm. Their shape is rectangular, with truncated ends and back. Their working edge is denticulated or retouched. Few have a triangular shape, designed to be end-pieces in a curved sickle.

The amount of tools uncovered is unknown, since the publication does not give a report on the inventory per site; quantitative and comparative studies could not be performed (.

The variety of raw materials, retouch and size (according to drawings, one sickle-blade, classified as “knife”, is 15 cm long and 3 cm wide), suggest that the production of sickle-blades was not standardized.

Scrapers Scrapers are the dominant tool-type: they are end or sidescrapers, on cortical or non-cortical flakes; a few are on blades or transversal flakes, others are circular.

The distribution of sickle-blades in the excavated houses (Noy 1998) shows that in 3/4 of houses, 1 to 10 sickleblades were collected, while almost a 1/5 of houses were void of sickle-blades. Site 6 represents an anomaly: 43% of all sickle-blades were found in two of its houses. A

81

CHAPTER V – NORTHERN ISRAEL SITES

occasionally cleaned from cortex by a large removal. The bulb of percussion was thinned, the retouch covering sometimes most of the ventral face.

recalculation of the distribution of sickle-blades, excluding Site 6 returns a similar distribution, Site 12 being exceptional, with up to 10 items in five of its houses. These high amounts (especially Site 6) may reflect workshops specialized in their production. However, since there are no details about the débitage at these spots, this suggestion cannot be validated. Another possibility is that the inhabitants of these houses left behind most of their sickle-blades, while others took them away when the site was abandoned.

The hole has an average diameter of 1.5 cm. The technique of drilling is unknown; a possible method was pecking a small cavity on one side of the item with a single blow, followed by shaping the perforation with a sharp tool, its marks, small dots, still visible on most of pieces. The perforation could have been also made by rubbing, which would account for the special colour of the flint, the scars on the opposite side being caused by the removal of the cone and additional hammering. On some perforations there are signs of extensive rubbing. Most tools have a standard size, probably the result of a required size and a specialized production.

Backed blades were regarded as sickle-blades without gloss. The ratio sickle-blades / backed blades is 3. At site 6, there are more backed blades than in any other house indicating a possible location for a workshop for sickleblades production.

Most star-shaped have 8 to 10 points located along edges at similar distances. Points vary in size, gradually increasing in length from the base towards the highest point of the distal curve of the edge's item, reaching 8 cm at the top. The concave base varies between slight to deep. They were constantly found in the excavated houses.

Retouched blades and bladelets Only a few items were found. Retouch varies from fine to semi-abrupt, irregular or regular, continuous or discontinuous; a few are ventrally retouched. Retouched bladelets are fewer; two are micro-end scrapers. Bifacials

The second type, with a concave base and without points, was found in various houses. Items were made on large oval or trapezoidal flakes, with a width varying between 8.3 and 16.5 cm and length between 7 and 14.5 cm. Edges are retouched in most cases. The hole has standard size (average 1.65 cm.).

A common bifacials is the adze (78% of the 87 bifacials recovered); axes (N=12), chisels (N=3) and picks (N=4) are present as well. Bifacials were made on dark and light brown flint with light brown to orange cortex and sometimes on veined flint, different from perforated tools or discs, and from the conglomerate flint. Bifacials are made mostly on flakes, though some are on chunks. Polish was frequently used. There is no evidence of a local production of bifacials.

The third type is similar in all its attributes to the previous one, but its straight base.

Axes have a triangular or trapezoidal cross-section. The working edge, straight or rounded, was shaped by longitudinal removals and polished (8 items). Eleven axes show signs of resharpening at the working edge. Ten items have a damaged working edge, evidenced by deep scarring. Chisels have a rectangular shape, a lentoid cross-section and the working edge is narrow. Chisels and adzes were produced by the same technique. Only one chisel was polished. Picks were made from large pebbles; one has battering signs on its butt.

The fourth sub-group consists of perforated discs. It was widely distributed, found in relatively large quantities (almost half of the perforated tools), and in most houses (5 in one house at site 12). They were shaped by bifacial retouch, covering both sides of the items, giving them a circular shape (mean diameter is 11 cm.). Exceptions are two big items (up to 18 cm in diameter). Their biggest thickness is at the centre, where the hole is located (0.8 cm). Edges are thinned by fine retouch. The mean diameter of holes (1.15 cm) is smaller than that of other perforated tools.

Perforated tools

Retouched flakes, truncations, notches and other tools

Perforated tools were found at all sites in the Golan, being a characteristic of assemblages of this region (Epstein and Noy 1988). Their distinctness is expressed by the perforated hole, the large blank they were made of, the high quality flint, shape and technique of manufacture. Four sub-types were defined: star-shaped with concave base, circular retouched with concave base, circular retouched with straight base and discs. Most items were made on a light-to-dark brown or veined flint, found in Eocene strata. Blanks measure as much as 18 cm in width and 15 cm length. The dorsal face was

Apart the note that many tools of this group were collected, neither their amount nor their provenance has been reported(cf. Noy 1998). Among the “other tools”, worth noting are pièces esquilles, a few modified into retouched blades and flakes, and burins, usually made on rejuvenated tools. Stage 5: Discard Even though an inventory list was not published (cf. Noy 1998), some conclusions can be derived from the

82

CHAPTER V – NORTHERN ISRAEL SITES

published material. Any comparison at site level may be inaccurate, since sites were not fully excavated and in several occasions houses were only partially excavated. Therefore, the comparison between sites was done at house level, regardless of its geographical position.

made between them. The comparison excluded partially excavated houses and houses with less than 10 tools. Thus, the total amount of houses that were compared is 15: nine houses from site 12, four from site 20 and two from site 6.

The sample used for the comparison of tool distribution was 39 houses: ca. 60% have less than 18 tools; a 1/5 have up to 52 tools and the rest more than 60 tools. All houses of the last group are the fully excavated ones at site 6 (N=2) and 6 out of the 16 of fully excavated houses at site 12. In order to understand if the differences are functional (flint tool production, activities related to particular tool types, etc.), a comparison was made between houses, at type level.

At site 12, a comparison was performed between houses, at (selected) type level: 3 groups of houses were noted: single houses with no artefacts, houses with amounts of up to 15% of the type (most houses) and a few houses with a higher than 20% frequency of the analysed type. Scrapers appear in similar frequencies (up to 15%), with a higher presence at site 15. Tabular scrapers were too few for statistical analysis; at house 7 there were more than in other houses (3, opposite to 1). A concentration of perforated tools and discs was found at house 7 (1/3 of all items of those tool-types from site 12). Adzes were found in similar amounts, except houses 13 and 15, each comprising ca. 17% of adzes found at the site. A concentration of sickle-blades was found in house 15 (a 1/5 of all of site 12); other houses had 13% to 17% of sickle-blades, or less than 10% (4 houses). Backed-blades and borers have a similar concentration. The distribution of tools, by their types, did not reveal any particular concentration. Therefore, the possibility that similar activities were performed in each house cannot be excluded. The high concentration of tool-types at house 15 may be the result of discarding methods, its inhabitants leaving behind their toolkits when they abandoned the site.

Scrapers and tabular scrapers were similarly dispersed between houses, in relative frequencies of 1 to 10% (each type). A concentration of perforated tools was found at sites 6, house B (20% of items) and house 7 at site 12 (11%). The rest of the items were found in very low quantities in other houses. The same is true for discs (3 to 10%), except house 7 at site 12, where 16% of the discs were recovered. Adzes have similar low frequencies at most sites (less than 5%), but house B at site 6 (18%) and houses 13 and 15 at site 12 (14% each). A cluster of sickle blades was found at site 6, house D (more than ¼ of all sickle-blades); moreover, almost half of them were found at this site. The remaining the sickle-blades were equally divided between various sites, in percentages of up to 5%. The distribution of backed blades follows that of sickle-blades: most were found at site 6; however, their concentration was in house B, and not house D, as the sickle-blades. A high frequency of awls was noted at house 15, site 12.

Three out of the four houses analysed at site 20 had a similar amount (ca. 30 tools each). The fourth had only 14 tools. Scrapers are more common in house 20 (44% of scrapers). Tabular scrapers, perforated tools and discs were too few for a meaningful comparison. Noticeable is the absence of bifacials (only one adze found at house 22). Almost half of the sickle-blades and backed blades were found in house 1. Borers originate mostly from houses 20 and 22. Retouched flakes were found in similar amounts in all houses. Apparently, activities were performed with different intensities at each house. If this is the case, a division of labour can be reconstructed: activities involving scrapers and borers took place at house 20, while cereal reaping was more common to the inhabitants of house 1. This difference may be also related to seasonal changes of activities, or seasonal occupations of houses.

This comparison revealed some “anomalies”, reflected in concentrations of particular tool-types in certain houses. These are the two fully excavated houses at site 6 and three houses at site 12 (7, 13 and 15).Houses at site 6 are the “isolated farms” type, located ca. 100 m from each other. Houses 13 and 15 of site 12 belong to the same chain-house; house 7 belongs to another chain-house, located some 20 m. from the previous one. Their similar orientation indicate their contemporaneity. Apparently, particular activities were performed therein. House B of site 6 outstands, with a high concentration of adzes, perforated tools and sickle-blades, and only a few retouched flakes. The possibility that a specific activity, concerning all three types of tools, or three different activities, took place in this house cannot be rejected. The high amount of cores of site 6 house B suggests an intensive knapping activity. Therefore, the possibility that it served as a workshop for the production of perforated tools, adzes or sickle-blades (or all three tool types) cannot be rejected.

Notable is the picture of site 6, houses B and D. Both are single houses, located at ca. 100 metres one from the other. When their tool assemblages are compared, a difference was noted at type level. Given the high number of tools recovered from each house, the possibility that the differences in the relative amounts of tool-types are the results of different activities cannot be excluded. Consequently, at house B, perforated tools, bifacials and backed blades dominate, while at site D, scrapers, sickleblades and retouched flakes form the majority. Other

Assuming that the excavated houses were contemporaneous within each site, a comparison was

83

CHAPTER V – NORTHERN ISRAEL SITES

and of perforated tools. Bifacials (excluding picks), were shaped from two striking platforms located along edges (according to drawings). Most are on large flakes, some with cortex. The working edge, occasionally polished, was shaped by longitudinal retouch. Resharpening occurred, as well as rejuvenation; some bifacials were used as cores. Picks were made from large pebbles; their butt being sometimes battered.

tool-types were found in insignificant quantity. Notable is the low amount of borers, in both houses (cf. Noy 1998). Summary Different raw materials were exploited, regardless the geographic location of sites. Dark brown pebble flint, most probably originating from conglomerate strata in the Golan area, was used for the production of most tooltypes, such as scrapers, borers, retouched flakes and notches. Grey-veined flint and light-colour flint, originating also from conglomerate strata, were used mostly for the production of blade tools, such as sickleblades and retouched blades. Semi-translucent flint, of Eocene origin and locally found under the basalt layer, was used in small quantities. A different type of pebble flint, with a light colour cortex and light brown texture, of higher quality, was used for the production of bifacials, tabular scrapers and perforated tools. Perforated discs were mostly made on high quality Eocene flint, with small nomolithic fossil accretions, found in several formations in the Upper Galilee.

Sickle-blades are on Eocene flint of various colours; a distinct group consists of brown-orange flint. The presence of cores of this type of flint (with blade removals, according to drawings) suggests their local production. Moreover, the presence of sickle-blades of this type of flint at site 6, houses 12 and R of site 12, and the core found in house 10, site 12, may reflect their contemporaneity. The presence of backed blades, probably unused sickle-blades, hint at a local procurement of these items. Despite their high morphological variability, a general shape can be drawn: rectangular, bi-truncated ends, back and a denticulated working edge. Some items have a triangular shape, apparently coinciding with the curved shape of the sickle.

The flint industry is flake oriented, evidenced by the large amount of blank flakes, flake cores and tools on flakes. Most tool-types were on-site produced. Even though only a few blade cores were retrieved, the possibility that some exhausted cores were used at first removal stages for blades and bladelets production of cannot be ignored. Some 2/3 of cores are small, exploited from at least two striking platforms. The intensity of their exploitation vary between sites, reaching the highest peak at site 6, where 90% of cores are small and exhausted. The apparent economy of raw materials is exemplified by a high amount of rejuvenated tools. Core-choppers were found at all sites. The possibility that some were tools should be considered, given their damaged working edge and battered butts of some of them.

Borers are dominated by awls, made on dark brown pebble flakes. Almost a quarter of them were made on resharpened tools, either retouched flakes or scrapers. Other tools are retouched blades, bladelets and flakes, and notches. Important to note is the presence of pièces esquilles, some modified into retouched flakes or blades. Any analysis of the spatial distribution of flints is limited to tools found in fully excavated houses and restricted to selected tool-types (Noy 1998). Among the 39 compared houses, a fifth have more than 60 tools; these are 5 houses in site 12 and the 2 fully excavated houses of site 6. The possibility that the higher concentration of tools in these houses is the result of the abandonment strategy of houses by their inhabitants, who left behind most of their toolkit, cannot be neglected. The high concentration of sickle-blades, adzes and perforated tools at site 6, house B, may indicate the presence of a workshop.

Almost half of the tools are retouched flakes and notches (such as at site 12). Scrapers are the next major group (13%), followed by sickle-blades and backed blades (10%). Other tool-types appear in less than 10% of all. Scrapers, made on various types of raw materials and blanks, have different shapes. Tabular scrapers, some with a small hole in the middle, are mostly on transversal flakes, have an elliptic shape, parallel retouch and a thinned bulb of percussion. Almost all items retained cortex on the dorsal side. One item has straight, crossing striation marks on its cortex.

The comparison of tools found in houses revealed various concentrations of tool types, and the absence of others in different houses. The possibility that certain houses were involved in particular activities, lacking in others, cannot be ruled out. However, this difference is independent to house type (isolated farm or chain-house). Thus, it seems that the subsistence economy at each site was basically similar and independent, with possible small-scale, partial specializations, such as at house B at site 6.

Perforated tools (star-shaped, circular and discs) were found at each site. Their characteristics are a drilled hole in the middle (1.5 cm mean), size (mean of 16 cm diameter), a high quality of raw material and a unique mode of manufacture (see above).

The burial cave in Peqi’in (181.35/264.40) The site was discovered in the Upper Galilee mountains (Gal et al. 1997). Main archaeological remains were 2 Neolithic occupation phases, followed by the principal, and most impressive use of the cave as a burial site

The bifacial group is dominated by adzes; axes, chisels and picks appear in small amounts. All are on high quality flint, different from the pebble-conglomerate flint

84

CHAPTER V – NORTHERN ISRAEL SITES

during the Chalcolithic period. Its main entrance has been blocked by natural activity and is indiscernible today. The cave consists of three units, on three levels connected with corridors. The upper unit (4.5 x 6 x 4m), was levelled, raised and partially paved; a stone platform was built on its Southern side. Ossuaries were placed on the paved surface. A niche in the Northern wall was filled in bones related to secondary burial. The middle unit (3 x 4.5 x 1.2 m) contained stone built terraces which levelled the floor; ossuaries and burial jars were found here as well. The lowermost, and largest unit (6.5 x 5.5 x 2 m), was found in the inner parts of the cave. A depression in the middle of the floor was intentionally filled with soil and stones, in order to level the floor. Stone-built platforms were found here as well. The cave was robbed during the Chalcolithic period, as evidenced by the disorder in which most artefacts were discovered. The most impressive remains of material culture are the numerous ossuaries, with various decorations and shapes, burial jars, high-footed bowls, decorated with human faces and sometimes containing human skulls, small vessels, a few copper objects, most notably standards, an ivory sculpted human head and violin-shaped stone figurines. Among these impressive remains, flint objects were found as well: a few perforated discs and chisels, mostly polished (ca. two dozens each type). Their symbolic implication is suggested by their archaeological context and their mint nature. Preliminary C14 dating support an apparent prolonged use of the cave as a burial cave, extending over a few centuries, used by inhabitants of various zones of Israel, as evidenced by remains of material culture, characteristic of both southern and northern, eastern and western sites of Chalcolithic Southern Levant (cf. Gal et al. 1997). The cluster of material culture from these zones indicate a similar cultural origin, or at least a common set of beliefs and burial customs for the Chalcolithic inhabitants of Southern Levant. The nature of the site, accidentally found, hint that Peqi’in should not be viewed as a unique phenomenon in the Chalcolithic universe, and more sites as this are expected to be found (such as Kissufim in the south, Goren and Fabian 2002).

85

CHAPTER VI –SITES IN JORDAN

that time, variety of material culture, including new forms of pottery, stone, flint and tools, stamp-seals, human and animal figurines and decoration objects on exotic material (Red Sea shells and alabaster). Burials of infants in large jars were found on floors of houses, some with “imported” objects: a hematite sceptre, an obsidian knife and mother-of-pearl components. The fourth season revealed one of the most impressive finds of the Southern Levant’s proto-history, a series of murals in several houses, depicting various scenes, such as a procession of gift-bearers, a bird figure and a mask-wearing human figure, related to a spectacle amulet found at the site, both resembling the eye-idol cult of Tell Brak in Mesopotamia (Lee 1973: 340). One of the painted-walls room contained also the six animal horns (Koeppel 1938).

CHAPTER VI SITES IN JORDAN Most Chalcolithic research was concentrated on the study of the settlement in the Jordan valley, and especially the site of Tuleilat Ghassul, while vast regions of Northern and Southern Jordan are still under primary investigation (Bourke 2001). A general overview of the Chalcolithic settlement in Jordan presents a picture of a dense occupation of the Jordan valley and a more restricted one in the foothills and the western margins of the upland plateau; marginal zones to the east and south were only sporadically occupied. There are more than 100 sites in the eastern part of the Jordan valley, several of them very large (more than 10 ha), while most are smaller than 2 ha. The occupation of the Jordanian uplands is much less dense, most sites being smaller than 1 ha. North plateau sites are a mixture of small settled villages or pastoral encampments, with a ground stone repertoire resembling the ones of the Golan sites, while the material culture of sites in the central plateau, such as Sahab, link them to Teleilat Ghassul. In the south, Tall Magass stands out as a unique site of the coastal region as a large-size village, while other sites are of a smaller scale (Bourke 2001).

The second excavation season (North 1961) concentrated on stratigraphic problems, concluding that the 4 main layers were occupied for short periods, without evidence of destruction and a homogeneity of material culture from bottom to top. The third excavations season started in 1967, conducted by J.B. Hennessy, of the British School of Archaeology (Hennessy 1969, 1982). This season was concentrated as well on solving stratigraphic questions, which remained open since the previous work. The excavation revealed 9 building phases and over 100 floor levels and evidences of an earthquake, which presumably forced the abandonment of the site.

The analysis of flint artefacts from Jordanian sites must take into consideration the problem of a poorly defined prehistoric period, both in general terms of definition and in terms of periodization and cultural material (Kerner 1997). Moreover, there is a lack of correspondence of the Chalcolithic period’s definition and cultural characteristics between various researchers working on both sides of the Jordan river, even though similarities in terms of material culture have been noted in the past (cf. Gilead 1995). The situation is even less clear in Southern Jordan, where the term Late Chalcolithic is usually employed for the periodization of apparently different periods (Chalcolithic and EB). Even though the possibility that the area was inhabited during the Chalcolithic period cannot be excluded, its nature is hard to be evaluated in light of current research (Genz 1997).

Renewed excavations at Ghassul started in the nineties, under the auspices of Sydney University (Bourke 1997); no final results are yet available, though a few conclusions were recently articulated by Bourke (2001). Apparently, the housing units (representing dwellings of nuclear families) were arranged in irregular clumps around open air public spaces and were linked together by alleyways, resembling a large but undifferentiated farming village, with a considerable variability in size and elaboration of dwellings (Bourke 2001:120).

Teleilat Ghassul (ca. 208/135)

Despite the large and impressive amount of material culture collected during all excavation seasons, and the fact that Teleilat Ghassul and its findings serve as a hallmark of the Levantine Chalcolithic, there is a lack of a systematic publication, synthesizing all excavation seasons and devoted to all aspects of material culture, especially the lithic assemblage. Following is a summary of Lee’s publication of flint tools and personal comments derived from observation of the published plates.

The site is located in the Northern part of the Jordan valley, a few km northeast of the Dead Sea, at the edge of a plateau facing the valley, in close vicinity to Wadi Djarafa. An early survey estimated the extent of the site over an area of 300x200 metres, without clear boundaries (North 1961) and no apparent planning. The depth of the archaeological layers were ca. 5.5 m, divided into four main occupation strata. The first two excavation seasons were conducted on behalf of the Pontifical Biblical Institute. Another season followed, directed by the British School of Archaeology in Jerusalem and in recent years, renewed excavations of Sydney university are ongoing.

Several raw materials were exploited by the Ghassul’s knappers, among them tabular flint, grey and brown pebble flint and semi-translucent flint are discernible. Most of the material originates from Senonian layers of the Moab mountains, located a few km east of the site. Other sources are wadis of Ghor’s valley, found in the immediate vicinity of the site.

First excavation seasons (Mallon 1930, Koeppel et al. 1940) exposed rectangular house foundations, mud-brick walls, pits, installations and a large, mostly unique for

86

CHAPTER VI –SITES IN JORDAN

Worth mentioning are choppers made on pebble flint, some with blunted edges and a broken star, originally referred to as a multiple tool (cf. Malllon et al. 1934: planche 31:4). Its presence is important to note, since this tool-type was originally reported exclusively from the Golan area. Thus, the tool assemblage combines traits common in Northern Negev assemblages, as sickleblades on banded flint, occasionally shortened by microburin technique, bifacials with a transversal blow on the working edge and picks, and Northern Chalcolithic assemblages, such as stars, a high amount of chisels and a non-standardized group of sickle-blades.

The tool assemblage is dominated by ad hoc tools, such as scrapers, retouched flakes and blades, notches, denticulates and borers, being similar to those of Be’erSheva assemblages. Tabular scrapers are common, some with thinned bulbs of percussion and a faceted butt; they have different shapes and were made on various raw materials, hinting at several sources of procurement. Sickle-blades are backed, have bi-truncated ends and finely denticulated working edges. Lee (1973:251) quotes Neuville’s finding of a blade with probable remains of asphalt along its edge; this find, together with a lump of asphalt collected at the site suggests that sickle-blades were hafted using asphalt as an adhesive material. Personal observations on sickle-blades stored at the Pontifical Biblical Institute in Jerusalem suggest a higher variability of shapes and raw materials than those of Northern Negev assemblages, items on banded flint appearing along sickle-blades on semi-translucent flint and brown or grey pebble flint. Sizes of sickle-blades also vary considerably, as well as the presence/absence of other typical attributes, such as truncations, denticulation or backing. Thus, it is suggested that either sickle-blades were fabricated in several workshops, or by each household, or there was no tradition related to their production. One item has a curved back (Malllon, Koeppel and Neuville 1934: planche 29:14), hinting at the curved shape of the sickle. Intentional snapping of the ends of the sickle-blades, probably by the application of the micro-burin technique, was observed on several items (Malllon et al.1934: 59, planche 29:15), a method found so far in Northern Negev assemblages (see above).

East Jordan valley sites Following a survey along the eastern side of the Jordan Valley (Ibrahim, Sauer and Yassine 1975), 15 sites were dated back to the Neolithic/Chalcolithic periods, located along side-wadis of the east foothills, in the flat floor of the valley, on the edge hills overlooking the valley and near the Yarmouk and Jordan rivers. Some are low, oneperiod tells, while others are large, sprawling areas, with a high quantity of cultural material, but lacking discernable architecture. Among prominent sites are Tell Shuna North, Tell Abu Habil and Abu Hamid. Tell Shuna North (ca. 205/225) The site is located in the Northern Jordan Valley. It was excavated in the 50’s (de Contenson 1961) and later, during the 80’s, a salvage excavation was conducted (Gustavson-Gaube 1987). Early excavations (de Contenson 1961) revealed 3 main levels: level I – pits dug in the virgin soil, filled with cultural material resembling the Pottery Neolithic B phase, a sterile level of 40 cm, followed by level II – a thin level of 20 cm, and within it the remains of a living surface and stone structures. Level III was dated back to the EB and it is not discussed here. The lithic and pottery assemblages of Shuna II were compared to those of Ghassul.

Microliths were found in all levels, usually in small concentrations of a few items. Though not specified, the dominant type was probably the micro-end scraper, other microliths being finely retouched bladelets. The presence of a bladelet on obsidian (Lee 1973:260) may serve as evidence of long-distance contacts, the closest source of obsidian being in modern Turkey. The bladelet cores and blanks found denote a local production of microliths, all on typical semi-translucent flint Lee (1973:251), some possibly originating from a workshop (Bourke 2001:139).

The last excavations, 75 sqm., exposed a sequence of 109 occupation strata, from the Neolithic through Chalcolithic to EB periods. General characteristics are residential areas, characterized by a series of multi-phases, multiroom dwellings with associated courtyards and working areas of beaten earth and cobble pavement and various installations, pits, dumping zones and hearths. Fragments of walls delimited dwelling complexes.

Among bifacials, all sub-types appear: axes, adzes and chisels (dominating the group). Apparently, the working edge was obtained by several longitudinal blows and was occasionally polished. Several items have a working edge shaped by a lateral removal, a common phenomenon of Be’er-Sheva basin bifacials. Bifacials were shaped from two striking platforms, situated along the edges of the items. The presence of roughouts suggests a local production of bifacials. Picks, another bifacial sub-type common in the Be’er-Sheva basin assemblages, appear as well, probably in a lower amount than chisels or adzes. They were made on large chunks, with one sharpened point, obtained by the bifacial removal of flakes.

Abu Hamid (ca. 210/220) The assemblage was published by Dolfus, et al. (1986). The site is located in the Jordan Valley, on a terrace comprised of marls deposited in the Pleistocene Lake Lisan (Map 2). The site is situated at 240 mbsl, delimited on its southern and northern edges by two small valleys in which perennial springs can be found; the Jordan River lies only half a km west thereof. A survey and systematic collection of surface finds was conducted over an area of

87

CHAPTER VI –SITES IN JORDAN

from the wadis were usually of mediocre quality. Despite their large presence in the wadi beds and terrace near the site, many are unsuitable for knapping; this could explain the observed heavy nature of some of the débitage.

20 ha at the site and its surroundings. The main concentration of artefacts, which delimits also the borders of the site in which structures were found, is ca. 2.5 ha. In addition to some small soundings, 300 sqm were opened in three main areas. The occupational deposits are thin, varying from 0.3/0.5 m to 1/1.2 metres in depth. All the exposures suggest a short occupation of the site.

Stage 1: Production The flint industry is dominated by debris, (almost 1/2 the assemblage), suggesting a local production of tools, from first manufacture stages. The industry is clearly flake oriented, flakes being ten times more frequent than blades and bladelets. Among wadi pebble cores, most were exhausted; large cores are extremely rare. The poor quality and scarcity of this flint might explain why raw material was quarried from in situ beds in the mountains. A complementary explanation derives from the calculation of the ratio tools/cores (1.6) and the large amount of flakes. In order to obtain suitable blanks, cores were extensively prepared, even though a limited number of blanks were further modified into tools. This suggestion is supported by the ratio flakes/tools on flakes – for each flake modified into a tool, almost ten flakes were discarded. Moreover, the large amount of flakes hints at the local production of most tools.

Description of the site In the first two areas, two occupation levels were distinguished. The basal level, overlying the sterile soil, is composed of plano-convex mud-brick structures. Occasionally, a layer of stones marks the foundation of walls. Gravel and pebble surfaces were found related to these walls, as well as packed-clay living floors, on which material culture was found. These structures were cut by later, cylindrical pits, of the following occupation, badly eroded. It consists of patchy remains of burnt clayey floors, mud-brick walls, plastered storage bins, fire-pits and hearths, and large refuse pits. Open areas, sometimes delimited by rows of stones were found, filled with organic material and remains of structures for burning. In the third area, associated with mud-bricks and small stone-lined rectangular structures, a series of large depressions was noticed. In one of them, a circular structure of 1.5 m in diameter, built from a double row of bun-shaped finger-impressed bricks, four courses height, was found. The mud-bricks formed a protecting ring surrounding the rim of a huge jar, a pithos of 1.5 m in height and 1 m in diameter, placed on a stone slab, in a pit dug into the marls.

Stage 2: Production of blanks From a technological point of view, there is a huge mass of crude flakes with flat butts and prominent bulbs, and some artefacts that witness a careful flint knapping, such as flakes and blades showing preparation of butt and predetermined shape. A comparison of the relative frequencies of different butt types on flakes and blades shows that smoothed (plain or flat) butts are most common (65%); cortical butts appear on 7% of items, suggesting a local production, from first stages of cores exploitation. Punctiform and linear butts, common to blades and bladelets, appear on 8% of items.

No complete houses or rooms were excavated. However, the interpretation of the results of these excavations shows that structures were rectangular, the settlement opened directly to the fields and the occupation period of the site was short.

Stage 3: Shaping/Retouching stage – retouching of tools

Description of the flint assemblage The first season yielded over 5000 flint artefacts, among them almost 300 tools.

The inventory of the tool assemblage shows that the industry is dominated by retouched flakes and notches (1/3 of tools). Scrapers and sickle-blades are the next important category, each comprising 1/5 of tools. The later are backed, bi-truncated and with a denticulated working edge.

Stage 0: Acquisition – Extraction and testing of nodules Several types of raw materials were used, including flint with fine-grained texture, coarser flint and chert, and siliceous limestone. The main source of raw materials were pebbles, found in the lower levels of the marls as well as in the beds of ancient wadis. They are recognizable by their rolled nature and altered cortex, being washed down the streams from the Ajlun Mountains. Other flint, with less altered cortex, was collected from in situ layers, noted in sections cut into the slopes of the nearby wadis in the Ajlun Mountains; its closest source is 15-20 km from the site. It consists of a variety of medium grey materials with a homogeneous structure and a thick cortex. The flint pebbles collected

Noteworthy is the presence of two perforated discs and two crescent-shaped scrapers, with gloss along their working edge. The two discs were found together, one above the other. They are bifacially modified, with a thinned bulb of percussion. Their raw material (greybeige ochre-veined with cortex) was found to be used on other tools as well. Its origin is probably in the beds of Ajlun Mountains. The presence of a massive core on this kind of flint, found on the surface of the site, hint at a local production of perforated discs.

88

CHAPTER VI –SITES IN JORDAN

Cores (N=69) are of various shapes and reduction schemes: unipolar, bipolar, prismatic or amorphous. Most are exhausted. Their original size was probably large and thus their intensive exploitation, as exemplified by the presence of large primary elements and large CTE, knapped at an early reduction stage. Several types of CTE were found; among them, ridges and tablets indicate a technical knowledge that apparently contradicts the “simple” knapping method employed for the production of some tools. This phenomenon may be explained as a deliberate choice to apply “simple” knapping techniques, and not as a lack of knapping knowledge.

Summary A large amount of flint items was collected, both from surface and from excavated features. Apparently, all artefacts belong to a single cultural phase, comparable to sites in Northern Israel, in the Golan area, with Ghassul, senso lato, and with Northern Negev sites. The flint assemblage is dominated by debris, which indicates a local production of most tools. The flint industry is flake oriented, flakes dominating the débitage group; they are also the most common blank type used for the production of tools. Among tools, retouched flakes and notches dominate. Noteworthy is the relatively high amount of sickle-blades (20%), resembling Ghassulian type – backed, bi-truncated and with a denticulated working edge. The presence of perforated discs may relate the site to others in the Northern parts of Southern Levant.

Most débitage are flakes (80%); primary elements, blades and bladelets appear in similar amounts (6% each). The lack of faceting on cores’ striking platforms hints at a non-systematic method of obtaining blanks. Most butts are flat (72%), with a pronounced bulb of percussion, resulting from a direct blow with a hard hammer. The presence of other butt types, such as punctiform, linear and dihedral represent various knapping methods. The presence of flakes with polish, found at workshop 574, is evidence that the workshop was also used for rejuvenating bifacials and not only for their production.

Description of three flint workshops Three workshops were exposed in phase III, belonging to the first half of the 6th millennium BP (Navarro i Barberan 1997). They were found in two areas: two in area A and one in area B. Several occupational layers were identified in both areas, without any signs of abandonment or destruction. The main archaeological features are rectangular structures with stone foundations and mud-brick walls and later additions of small rooms, open courtyards, containing paved areas, small basins and depressions, probably used for various domestic activities and storage installations. Each workshops was found in a different occupation layer, at various locations within the site. Lithic remains contained cores, tools, flakes, blades and a large amount of debris. All artefacts were found in primary location, horizontally laid and in mint condition.

Among tools, those on flakes dominate (scrapers, borers, notches and burins); tools on blades are backed sickleblades. Bifacials were found only as fragments. The workshops were used both for the production of tools and for rejuvenation of bifacials. The presence of several knapping methods and the high variability of products’ morphology should be related to an adjustment to needs and not a lack of technical knowledge. The high degree of cores’ exploitation may be related to the scarcity of good-quality flint in the immediate vicinity of the site.

The first workshop (574) was found on a surface without clear association to any habitation structure. It contained the largest amount of flint artefacts (Table 6). The second one (105) was found outside a house, directly on the surface. It was heavily affected by erosion, which probably caused the removal of an unknown quantity of flint artefacts. The third one (157) was found outside a storage complex, containing a large jar, described above. The flint artefacts were concentrated in an area of 2m2.

WZ121: Tubna – Wadi Ziqlab The site is located at an elevation of about 500 MASL, on a limestone terrace facing one of the main tributaries of Wadi Ziqlab, in Northern Jordan (Banning et al. 1998). It has been disturbed by modern agricultural activity and by ancient Byzantine levelling of the terrace. Its stratigraphy includes the natural bedrock, on which Late Neolithic material was sporadically found, above it, a Chalcolithic occupational layer, with a varying thickness of 0.2 to 1 m, covered by Chalcolithic structures and pits. Thus, it seems that the site was used in several episodes during the Chalcolithic period. Two large walls were uncovered, probably parts of a dwelling complex that included courtyards. Several pits of various natures were uncovered in their vicinity. Two pits, with a diameter of 2 m, were stone-lined, while others are simple pits of different capacities. All pits contained similar fill, with ashy sediments, pottery fragments and flint artefacts, serving probably as dumping zones. Among the remains, worth mentioning are two broken hematite maceheads and fenestrated basalt vessels.

The flint inventory of the three workshops is presented in Table 6. Several raw materials were used, varying from fine granulated to coarse grained flint of different textures. All types are available in the immediate vicinity of the site, in wadis descending from the Ajlun mountains. Even though flint was abundant, its poor quality may explain the high intensity of cores’ exploitation, within the area of the site as well as in the workshops. Moreover, this may explain the production of large tools, such as bifacials, from raw materials brought in from a farther distance, such as the in situ banks found in the mountainous area. The possibility that the raw materials used for the production of micrograttoirs originate from a farther source cannot be excluded.

89

CHAPTER VI –SITES IN JORDAN

Only preliminary reports are available, and the lithic material was briefly analysed. It consists of Ghassulian tools, such as sickle-blades and bifacials, made on local material, found as nodules in the limy sediment on which the site is located. There is an apparent deliberate choice for a chalky, softer flint for the production of bifacials.

Tell el-Hibr The site was excavated in 1989 and its findings were published by A. Betts (Betts 1992). Located in the eastern semi-arid zones of Jordan, the site, a rock shelter on the slopes of Tell el-Hibr, was discovered during a survey of the eastern volcanic rocks, on a basalt-capped hill about 10 km east of the main volcanic massif. Within the rock shelter, an in situ deposit of 1 m of human occupation was excavated. Main features are four over-imposed pavements, made of flattish slabs of limestone and basalt, suggesting several occupation layers. Material culture remains include pottery fragments, jars resembling Be’erSheva Chalcolithic ones (Betts 1992:11), a stone ring and flint implements. Faunal remains included equid, gazelle, hare and small mammals, but no caprine.

Pella The site is a large tell in the lower foothills of the Jordanian plateau, about five km. east of the Jordan River (McNicollt et al. 1992). Two Chalcolithic occupational spots were located, one on the southern slope of the tell (area XXV) and the other one in area XIV, defined as the settlement of Jebel Sartaba, located near the tell. However, surface findings and material collected from various areas suggests that the Chalcolithic settlement extended over an area of at least 30 ha.

The flint assemblage consists of more than 700 items, among them 181 tools. The raw material was quarried from adjacent hill slopes; chunks were modified into cores, with few flake removals. The knapping method was basic, most flakes and blades having broad and plain butts. A few flakes with a wide faceted butt suggest the use of discoidal cores. Tools are dominated by ad hoc items, mostly retouched flakes and notches. Scrapers, burins, truncations and borers are present as well. Important to note is the presence of three transversal arrowheads and one possible tabular scraper.

The settlement of area XXV was found on a gravel layer, in close vicinity to the great spring, on a high place. It consists of ca. 75 cm. of occupational accumulations and material culture collected from pits of various size and shapes. Fragments of stone-built structures were noted. The internal stratigraphy of this area suggests that the Chalcolithic occupation was multi-seasonal, given the overlapping of some pits and walls. The remains in area XIV were found on a levelled platform, on which buildings were erected from dry rubble with stones and mud-bricks. There is no planning, structures seem to follow the lines of the bedrock. Floors were plastered or made of beaten earth. Pits of various sizes and function were dug from these floors. Adjacent to structures are open courtyards with installations. The settlement covered the steep hillside with small clusters of rooms and adjacent courtyards, reminding Chalcolithic sites of the Golan (cf. Epstein 1998).

A major conclusion is that “…the people using the rock shelter…were probably an indigenous North Arabian group without close connection to steppic populations in Northern Sinai” (Betts 1992:16), despite a previous conclusion regarding the ceramics of the site: “…the best typological parallels…occur in Chalcolithic assemblages of…Be’er-Sheba area…relations…point to an east-west zone in the Southern area of Transjordan and Southern Palestine” (Betts 1992:11). It is therefore suggested to relate the lithic assemblage of Tell el-Hibr to similar ones in the Arabah and Sinai arid zones, in assemblages often referred to as Timnian or Eilatian (see below), given its general characteristics: a crude industry exploiting local flint sources, the high percentage of ad hoc tools, absence of “indicative” tools and transversal arrowheads, found in Sinai assemblages (Bar-Yosef et al. 1997)

The flint assemblage was probably locally produced, from raw materials available in the nearby outcrops. Cores are small, used for the manufacture of small flakes and bladelets. The tool-kit is dominated by blade tools; among them backed blades with either a triangular or a trapezoidal cross-section are notable. The presence of perforated discs links this assemblage to the Golan ones.

Southern Jordan sites (ca. 210/1)

Kerak Plateau sites

The assemblages were published by Henry (1995), who associated them to the Timnian complex – small, ephemeral hamlets and seasonal camps found in the desert of Sinai (Kozloff 1974) and Southern Jordan (Henry 1992). The architecture of hamlets is curvilinear and includes un-hewn stone pit houses, corrals, retaining walls and storage pits (Henry 1995:354). The settlement pattern resembles the one of historic Bedouins inhabiting the region (Henry 1992). An annual transhumance cycle consisted of the occupation of large, long-term seasonal camps during the winter in the piedmont zone, followed by 9-10 months of dispersal into smaller, more mobile

A small number of sites containing Chalcolithic pottery were recorded during a survey of the Kerak plateau, covering an area of 875 sqm. east of the Dead Sea (Miller 1991). They also contained a large number of pottery belonging to different other archaeological periods. Moreover, in none of these spots could a certain Chalcolithic site be identified. Therefore, the only assumption that can be articulated is that some Chalcolithic occupation, of unknown nature and possibly of low intensity occurred in the area of Kerak plateau.

90

CHAPTER VI –SITES IN JORDAN

camp groups during the dry season. The dispersal started with movements to spring pastures at lower elevations; as the season advanced, groups reversed the movements and progressively settled at higher elevations, eventually settling on the edge of the upland plateau. The economy emphasized the herding of caprines, but relied on gazelle hunting and lacked horticulture (Henry 1995:355).

Technology Blank production focused on two main reduction schemes, one devoted to the production of bladelets and the other to the production of large tabular flakes. Microliths were knapped from sub-pyramidal and globular cores, made exclusively on a fine-grained chalcedony. Tabular flakes, with limestone cortex, were made from in situ exposures of flint. The low proportion of primary elements found points to flakes production at the sources of the flint and not at the sites.

Description of sites Twenty-four sites were recorded (Henry 1995:356), found in each physiographic unit (lowlands, piedmont and uplands). Two types were identified: ephemeral and long-term camps. Ephemeral camps have smaller areas, thinner deposits and low density of artefacts. A wider range of artefacts was found in long-term camps, such as pottery, ground stones and bone tools, items practically absent in ephemeral camps. All long-term camps have architecture, non-existent in the ephemeral camps.

Typology Even though the assemblages contained small tool samples, the toolkits display a high diversity of classes, being dominated by scrapers, microliths and retouched pieces, but also containing other types, such as geometric microliths, burins, notches and points. Noticeable is the absence of bifacials and sickle-blades (Table 6).

Jebel el Jill (J14)

Scrapers

The site is at 1,200 masl, in a saddle between sandstone promontories at the base of Jebel el Jill. Artefacts, associated to circular stonewalls, were scattered on the southern slope (ca. 1,400 m2). Identifiable remains are a corral, ten m. in diameter and a semi-subterranean pithouse lined with stones, containing a fire-pit, grinding stones and concentrations of artefacts. Alternated ash lenses, separated by sand suggest episodes of abandonment; the compacted floor was probably formed during the winter-wet season (Henry 1995:359).

Tabular scrapers form the majority of this type, though typical end-scrapers on flakes and blades occur as well. Tabular scrapers, a hallmark of this industry, were made on large, thin and flat, cortex covered flakes. Heavy retouch is typically found along the entire perimeter of items (excluding butt); the angle of retouch and the configuration of the retouched edges show distinct patterns. The retouch along one lateral edge is invasive, creates an acute angle with the inverse surface and forms a convex profile. In contrast, the retouch along the other edge is abrupt, creates a 90º angle with the inverse surface and forms a straight profile.

Other sites Other long-term sites were found within an elevation belt of 1,100-1,200 masl along the eastern slopes of Jebel el Jill and Jebel Queisa. All display curvilinear structures. The number of structures per site is either 2-4, or 10-15. In the large sites, the walls of houses are connected in a “nested” pattern (Henry 1995:362).

Burins Burins lack patterning, while dihedral and angle burins struck from truncations predominate. Nongeometric microliths

Description of the flint industries

Significant numbers of abruptly retouched small flakes and bladelets occur within most assemblages, but they lack patterning in terms of retouch morphology. Most, in fact, are fragments of retouched flakes or bladelets.

The lithic assemblage shows a dichotomy in blank production that includes bladelets and microlithic flakes struck from globular cores and large flakes from tabular flint (Table 13). The microlithic component served as blanks for the production of geometric and minute backed elements, whereas large flakes were mainly used for the production of tabular scrapers.

Geometric microliths Most of the toolkits contain low to moderate proportions of geometrics, composed of lunates, and more rarely, trapeze and rectangles. Lunates (common in the EB, cf. Rosen 1997) are small, with an abrupt back, either unipolar or bipolar.

A distinct characteristic is the blockish and angular form of debris and débitage, derived from an intensive use of the “block-on-block” technique and hard-hammer percussion. Cores display high angles between platforms and faces and blanks are typically thick, and when missstruck, angular and blockish. Another factor that may have contributed to this characteristic is the high incidence of thermal fracturing.

Utilized pieces These are thick flakes with a steep bilateral retouch. Occasionally, marks of bifacial battering and crushing are discernible.

91

CHAPTER VI –SITES IN JORDAN

Summary The sites were dated to ca. 5,700 – 4,000 B.P. Camp groups moved seasonally, from long-term, large camps in the lowlands during the winter, to smaller and ephemeral camps, located at higher altitudes, as the season became hotter and drier. These groups were not involved in agriculture, as testified by the lack of sickle-blades and the absence of cereal pollen and cereal phytoliths. The subsistence economy was based on herding and foraging sheep, gathering wild plants and occasional hunting. A possible wool exploitation was seen in the presence of tabular scrapers, personal experiments and their apparent suitable morphology suggesting they were “…ideal for use as a sheep shear…” (Henry 1995:372). Southeast Arabah sites (ca. 180/1) The sites were discovered during a survey in the southeastern parts of Wadi Arabah, ca. 70 km. north – northeast from the Gulf of Aqaba (Smith II, Stevens and Niemi 1997), spanning over the Late Neolithic, through the Chalcolithic to the EB. They are located near Aqaba (sites 14, 66 and 70-76), along the north bank of Wadi Nuweiba (site 107), in Wadi Aheimar (sites 110-116 – oval enclosures, rock alignments and pits and standing basalt monoliths at sites 115 and 116) and Wadi Abu Barka (site 142 – rock alignments and a concentration of pottery sherds, flint artefacts and copper slag). Two sites, probably related to each other (sites 93 and 94) are located in Wadi Saminiya. The first one consists of two oval rings of stones (possible burials) the other one being a multi-room structure. Two additional sites, located north of Aqaba, in the area of Wadi Yutm, were excavated in recent years (site 66 – Tell Maqass and site 74 – Hujeirat al-Ghuzlan). The last two mentioned sites were apparently industrial centres (site 66) (copper, shell and stone were worked) and local market centres (site 74). The rest of the mentioned sites are smaller, though the presence of architecture (domestic dwellings, sometimes with mud-brick walls) was noted at the other site. At site 70, a concentration of slag was noted in connection to a massive wall. Adjacent features were stone rings and semi-circular rock alignments. Among the flint tools collected, tabular scrapers and bifacials are worth mentioning. None of the tabular scrapers has incision marks. The presence of sickleblades suggests that agriculture was practised in these areas. Their morphology is different from Ghassulian sickle-blades, but it is neither Canaanean.

92

CHAPTER VII –DESERT SITES OF ISRAEL AND SINAI

The technology employed in manufacturing tools is homogenous throughout the assemblages: there is a clear dominance of flakes at all sites (Table 13). The same is true regarding tools: almost 90% of them are on flakes.

CHAPTER VII DESERT SITES OF ISRAEL AND SINAI Nahal Nekarot (174.5/105) The site was discovered in 1976 in the cultivated area of Moshav Ein-Yahav, in the Northern Arabah valley, at the joining points of Nahal Nekarot and Wadi Arabah (Map 2), in close vicinity to water sources (two springs). It was reported by Beit Arieh and Gophna (1977). Finds were found scattered over an area of 500 by 200 metres along the Northern bank of Nahal Nekarot. Additional features were hearths (pits). No architecture was discovered. The pits were dug 60 cm. in the ground; their average diameter is 1.2 m. Their bottom preserved a heavy layer of ash, while their fill contained pottery sherds, flint artefacts, stones and grinding stones (found at the top of each pit). The pottery sherds and the flint tools resemble typologically and morphologically Ghassulian assemblages. The flint tool assemblage consists of scrapers, tabular scrapers, bifacials (axes) and sickleblades, some backed and truncated and others macrolunate like.

Stage 1: Production

Eastern Sinai sites (ca. 8.39/79.77)

Despite the appearance of flint hammerstones at several sites, there is no direct evidence of flint processing at these sites (a low amount of débitage was collected). The different nature of sites where hammerstones were found (near masseboth and in domestic areas), raises questions as to whether hammerstones had a double nature functional (flint processing or plant grinding), or symbolic (near masseboth sites).

Eocene flake cores are globular, with one non-faceted platform and partially cortex covered (Plate 39:1). Mean sizes are 70 x 50 mm. Only 3 blade/bladelet cores were found; one (Plate 39:2) resembles Chalcolithic blade cores: cylindrical, with one striking platform. Nearly half of its débitage surface was used for blade removals; the rest is cortex covered. Two others are bladelet cores, on Senonian black flint. They are smaller than flake cores (25 x 40 mm), with a pyramidal shape and two opposite platforms. Although blade cores were found only in one area, the flakes/blade ratio of all assemblages is similar (10 to 1). Senonian cores tend to be small and exhausted. Their high exploitation degree can be caused by the high quality of the raw material, the scarcity of its sources or the original small size of the pebbles from which they were manufactured. Tabular cores were not found.

During the months of November 1981 through April 1982 an archaeological survey was conducted by Avner in SE Sinai (Avner, in prep.), its main goal being the record of as many as possible proto-historic cultic sites. These were roughly dated to the 6th to 3rd millennia BC (Late Neolithic to EB), only few of them radiocarbon dated (Avner et.al. 1994). Flint items dominate the collected material; they underwent a detailed study, its results being in the course of publication (Avner in prep).

Stage 3: Shaping/Retouching stage – retouching of tools The description of the flint assemblages follows the main geographic areas where they were found and is reported according to their site provenance.

Collecting methods Most assemblages are surface collections from a limited area surrounding various features of sites. All discernible artefacts were collected. Only a few sites were excavated and their sediments sieved. The description of the assemblages is therefore limited to typological analysis and general technological aspects of their production.

Wadi Radadi The flint assemblage of Wadi Radadi sites (Table 20) consists of 12 tools and 4 waste pieces: 2 flake cores, 1 C.T.E. and 1 hammerstone. The presence of cores and the hammerstone reflect a local production (even though blanks were not found on the surface). Typologically, the tools resemble those of Wadi Zalaqa sites (see below).

Description of the flint assemblages Stage 0: Acquisition – Extraction and testing of nodules

Gebel Hashem Taref

All flint assemblages (regardless of the nature of sites or their geographic position) were made on two types of flint: Eocene brown nodular flint, used for the production of most tools, and Senonian black, yellow veined flint, used exclusively for the production of small flake tools and blades. The closest sources for Eocene and Senonian flint are located in the Tih-Egma plateau (central Sinai) and in the outlet of Wadi Feiran (south-west Sinai) (Geological map of Sinai 1973).

Several sites were identified, most being labelled as open-air sanctuaries (Avner 1984). Nine provided a poor flint assemblage, dominated by tools (Table 21). The scarcity of waste products may be the result of: tools were manufactured elsewhere (few to hundreds of metres) and brought to the sites, or the nature of the sites (open-air sanctuaries) dictated the type of flint artefacts.

93

CHAPTER VII –DESERT SITES OF ISRAEL AND SINAI

The raw material is similar to all other flint assemblages analysed. There is a preference for Eocene flint for most flint artefacts. Senonian flint was used exclusively for the manufacture of small flake tools (see below).

A considerable typo-technological similarity exists between tools of Wadi Zalaqa sites and those of this region. The preferred raw material is Eocene brown flint. Bir Sawane

Scrapers (N = 17)

The flint assemblages of Bir Sawane sites (Table 22) are similar to Wadi Zalaqa assemblages. Although the collection of flint artefacts was tool oriented, same basic characteristic of the tools assemblage are discernable: crude and highly non-standardized, dominated by notches/denticulated and retouched flakes (Table 3).

Most scrapers are steep retouched (Plate 43:4), made on large, cortical flakes (60 x 50 x 14 mm). Retouch is rounded, obtained by irregular flake removals. Their nonstandardized shape, the amount of cortex on their dorsal face and the irregular retouch give the general impression of a low investment in their process of production. This conclusion suggests a possible function of this tool: a simple, primary operation in a domestic household.

The scraper group is dominated by steep retouched items (Plate 40:1). All tabular scrapers (N=5) were broken (Plate 40:2). Borers were divided in small and massive awls (Plate 40:4,5). Notches/denticulated are also small or massive. Retouched flakes (N=21) have various shapes and size, modified by fine retouch along one edge (Plate 40:3). Two adzes were collected. Two flake cores and six hammerstones are evidence of a local production of tools.

Tabular scrapers (N = 15) Most tabular scrapers are made on Eocene flint. There is a high variability of shapes (oval to rectangular, Plate 43:1-3). In only one case the butt was faceted; in all others the butt is pointed or flat. A few were made on transversal flakes and the bulb of percussion was thinned. All other blanks are elongated, with a hinged distal end. The common retouch is parallel to sub-parallel, its angle abrupt to semi-abrupt, differing almost from one tool to another. Most tools have a rounded retouch; two pieces are retouched along the edges and one at its distal end. Most items (80%) are complete. No evidence for the manufacture of tabular scrapers was noted.

Wadi Watir The flint industry of Wadi Watir sites is similar to those described above. The collection focused on tools, cores and hammerstones (Table 23). The presence of 4 flake cores (Plate 41:1) and some twenty hammerstones indicate a local production of tools, except tabular scrapers. The high quantity of hammerstones can be the result of different activities performed with these, in addition to the manufacture of flint tools.

Three tabular scrapers, made on transversal flakes, were found in situ at site M301, in close connection with masseboth. They are on light grey flint, with a similar fan shape. Retouch is parallel semi-abrupt, covering ca. 10 mm of the dorsal face. Their location, close to masseboth, their mint condition and morphological similarity indicate their role – as apparent offerings brought in one episode.

The main tool type is the tabular scraper (N=12), made on Eocene flint. Most are elongated flakes, with parallel rounded retouch. One item has a faceted butt and its bulb of percussion thinned. The high amount of these items can be explained by: (1) closeness to a tabular scraper workshop, sites of this area serving as a temporary station in a possible exchange route of these tools, (2) functional explanation or (3) the result of the collection method, focusing on indicative tools. A noticeable feature is the presence of three bifacials: an axe, an adze and a chisel (Plate 43:1,2,3 respectively).

Burins (N = 1) Only one burin, on natural break, was collected; it seems that burins, if not intrusive, had no significant role in the tool assemblage. Borers (N = 9)

The varia group consists of four Levallois-like flakes and one unusual, unique tool (Plate 41:3), made on a large pebble, with ventral flake removals and bifacial retouch of half the tool, forming a kind of handle. Its end is rounded, steep retouched, forming a scraper-like working edge. A similar, smaller version, is drawn in Plate 41:2.

All borers are awls, evenly distributed among the different sites. They were sub-divided according to size: massive and small, with a mean size of 40x30x6 mm (Plate 43:5). Apparently, this difference reflects functional aspects. A difference also exists in their mode of manufacture: massive awls, made on brown primary flakes, were obtained by large, lateral removals and sharpening of natural breaks; small awls were made on Senonian flint, the point being obtained by two fine retouched notches. One awl is made on tabular flint, on a broken tabular scraper (Plate 43:2).

Wadi Zalaqa Several sites from the Wadi Zalaqa area provided poor (quantitatively) flint assemblages (Table 24). Most flint artefacts originate from domestic sites, the largest being collected from site M308-L1. A high homogeneity among tools and débitage was noted; thus, the collection was treated as one assemblage.

94

CHAPTER VII –DESERT SITES OF ISRAEL AND SINAI

Retouched flakes, Truncations, Denticulates (N = 30, 9, 43)

Notches

The presence of Levallois-like flakes is either due to disturbance of Middle Palaeolithic sites, or results of reduction sequences of (“accidental”) Levallois-like blanks, which can be produced from a variety of techniques and methods. There are no differences among tools found at Wadi Zalaka sites, despite their different nature. In general terms, the flint industry resembles the Eilatian industry (see above).

and

Retouched flakes are the second largest tool group. The common retouch is fine, its place and the shape of tools differ between the items. Truncations were made on small Senonian flint flakes. The common retouch is semi-abrupt.

Wadi Sa’al

Notches dominate the flint assemblages. Two sub-types were observed: notches on small to medium size flakes (60 x 50 mm) and large, massive denticulated pieces, on primary flakes (Plate 43:3).

The flint industry of Wadi Sa’al sites differs from all others described above, being a smaller size flake industry, without Levallois-like elements, made on Eocene grey flint. The waste group is dominated by flakes and chunks. The four cores collected are globular, with multi-directional flake removals (Table 25).

Retouched Blades (N = 9) Common retouch is fine, partially along one edge. The scars patterns hint at the possibility that they are in fact elongated flakes, and are not products of a blade industry.

The tool assemblage is dominated by notches/denticulates (Plate 44:7) and tabular scrapers. All (but two) tabular scrapers are broken. Two different raw materials were used for their production: grey, like the rest of the flint assemblage, and tabular brown flint, similar to the tabular scrapers described above. The grey tabular scrapers are thin (less than 6 mm), with partially fine retouch (Plate 44:2), made on rounded flakes. The rest (N=5), made on brown flint, are thicker (ca. 10 mm), with parallel retouch along their edges, being morphologically similar to other tabular scrapers collected during the survey (Plate 44:5). Consequently, at least two sources, with different traditions of tabular scraper production, can be discerned. Although their geographic position is unknown, one is related to Wadi Sa’al sites, the other with the sites described above. Scrapers are made on rounded thick flakes (60x60x20 mm mean), commonly modified by scaled retouch (Plate 44:1). All borers (N=4) were found at one site; they are awls on small flakes; the point was obtained by fine retouch (Plate 44:6). Massive awls, common in assemblages described above, are absent. The tool assemblage is completed by a few retouched blades (Plate 44:3), among them noticeable is an item with a curved back (Plate 44:4, see also Rosen 1997:62).

Bifacials (N = 1) One adze was recovered, at the surface near masseboth. It is unclear whether it was deliberately placed in relation to the masseboth, or was accidentally found there. Varia The varia group includes a bifacially retouched Eocene pebble, with scraper-like retouch at one end, obtained by large flake removals, which created a triangular crosssection to the tool. Other items are Levallois-like flakes (Plate 39:3,4). Their presence was important to note, since they represent characteristics of the Eilatian industry. One item (Plate 39:3) differs from the others in terms of raw material (grey, opposite to the common Eocene), with a heavy patina and a faceted butt. It displays all the characteristics of a Middle Palaeolithic Levallois flake. The remaining items are apparently integral parts of the assemblage. A Levallois-like blank can be obtained by different retouch sequences (Boëda 1991: Fig. 6). Therefore, they may represent regular blanks that “accidentally” resemble typical Levallois flakes, results of reduction techniques from flake (discoidal) cores.

Summary The flint assemblages can be divided into two groups: the first represented at Wadi Sa’al sites and the second group by the rest of the flint assemblages. The main differences are the absence (in the former group) or presence (in the later group) of Levallois-like items and the size of tools: smaller, with a lack of massive tools in the first group, opposed to larger and cruder with more tools with cortex in the second group. Another difference is the presence of thin tabular scrapers, partially fine retouched in the first group and their absence in the second group. Possible relations, direct or indirect (sharing common sources for tabular scrapers) can be pointed: the presence of tabular brown scrapers in the first group (representative of the second group) and the appearance of grey tabular flint scrapers in the second group (typical of the first group).

Summary The flint assemblage of Wadi Zalaqa sites is the product of a flake industry with a low level of standardization. Two distinct types of raw materials were used: Eocene flint for most tools and Senonian flint for small flake tools. Dominant tool-types are notches/denticulates and retouched flakes. Tabular scrapers, an important element in the tool-kit, display a vast gamma of shapes, retouches and raw materials. Given their great variability and the fact that no core for their production was found, different sources for their origins are proposed.

95

CHAPTER VII –DESERT SITES OF ISRAEL AND SINAI

The absence of arrowheads from these flint assemblages suggests a discrepancy between them and Late Neolithic flint assemblages of Southern Levant or nawamis sites. The same is true regarding parallelism with Ghassulian or EB flint assemblages: no similarities were found between these assemblages and those described herein.

Domestic areas Tools from domestic sites are surface collections. Their characteristics are the presence in varying frequencies of all tool-types. Differences exist between Wadi Watir sites and all others: the high frequency of tabular scrapers, the presence (contrasting to absence) of bifacial tools (axes and adzes) and the high number of hammerstones found at Wadi Watir sites, probably relatable to a particular activity performed there.

Apparently the first group of sites is a variant of the Timnian, while the second one belongs to the Eilatian. Some contacts existed between them, in terms of mutual exchanges (tabular scrapers), and thus, their coexistence. The differences between the two groups of lithic assemblages are not very pronounced; the appearance (or not) of the Levallois-like component may reflect only a different approach to a flake-core exploitation. The size of tools may reflect raw materials limitations, and the differences in the treatment of tabular scrapers can be related to different knappers. It is not clear how these flint industries are correlated with social groups of people: two (different) ethnic groups, populating the same area, occupying the same habitats and leaving behind similar cultic remains (Avner 1984, 1993), or a single ethnic group, which produced the two groups of flint industries.

Summary The flint assemblages described above were divided into two groups, corresponding to two flint industries: the Timnian (Kozloff 1974) and the Eilatian (Ronen 1970). The Timnian was identified only in Wadi Sa’al region, the Eilatian in all others. However, a more accurate description of theses industries is needed. Possible analogies between the nature of site (masseboth, sanctuaries or domestic sites) and their flint assemblage were verified. Apparently, the presence of flint tools in open-air sanctuaries is accidental. There is a quantitative difference between masseboth and domestic sites (fewer in the former); most tool-types appear in both kind of sites. In one case only, three tabular scrapers were the single tool-type found near a masseboth site (M-301 5), indicating a deliberate placement of these items near a masseboth. Given the “profane” nature of tabular scrapers (a common tool-type, found in domestics contexts), their presence close to masseboth can be explained by: a possible higher economic value of these items (obtained by mutual exchanges) which metamorphosed them into offerings for an unknown deity (the standing masseboth). Another possibility is that tabular scrapers were used in a ritual performed in the vicinity of the masseboth. The presence of tabular scrapers in burial contexts, further testifies to a non-profane nature of tabular scrapers. Their constant appearance in the tool-kit of Neolithic through Bronze Age households, confer a clue for their usual nature - a common tool-type of domestic use.

Discard Following is a description of the distribution of flint tools according to main features types: “masseboth” – standing slabs of stones, either singular or in groups, “open-air sanctuaries” – complex structures with a row of standing stones and a courtyard, and “domestic sites”, with architecture, pits and fireplaces. Excavated artefacts and surface collections were analysed separately. Masseboth sites The tool assemblages collected near masseboth (usually within a radius of 5 m) contained almost all tool-types. Retouched flakes and notches/denticulated dominate and appear regularly. A high number of notches/denticulated and tabular scrapers were collected from Wadi Sa’al 2. The tool assemblage from excavated masseboth is smaller than the previous one (ca. 2 tools/site). In general terms, a similarity exists between the two groups. Scrapers, borers and bifacials are absent from excavated masseboth.

Timna 39 (ca. 143/92) The site was discovered during the Arabah survey (Rothenberg 1978). It is located in the Timna valley, some 30 km north of the Gulf of Eilat on the Red Sea. Prior to the excavation, visible remains consisted of an oval enclosure, 29x23 m., with a wall of 1.5 m. width and 3 attached tumuli, apparently the remains of dwellings. Small mortars of red stone, granite pestles and hammerstones, flint implements and pottery fragments were noted on the surface, as well as copper ore nodules and small pieces of slag, indicating metallurgic activities. Another activity area consisted of a concentration of slag, flint artefacts and grinding tools, surrounding a smelting furnace. The cultural material collected and one radiocarbon date (4460 B.C.) place the site within the early phases of the Chalcolithic period (Sugar 2000:77).

Open-air sanctuaries The surface collection of tools from open-air sanctuaries displays a picture of heterogeneous groups of tool-types. The flint tools collected from excavated open-air sanctuaries exhibit a picture of a very limited tool-kit, lacking tabular scrapers, usually one tool/site (except M311), possibly the result of a deliberate cleaning by their users. Therefore, it was assumed that the presence of flint tools in these features is accidental.

96

CHAPTER VII –DESERT SITES OF ISRAEL AND SINAI

characteristic of this group: 6.9 ± 1.9 mm. The standardized nature of the cache is evidenced by the uniformity of raw materials used and size of items.

Among flint tools, scrapers dominate; one is a corescraper. Only one tabular scraper was found. Its shape is oval, without incisions on its cortex, and with the butt and bulb of percussion thinned. Important to note is the presence of bifacials, such as chisels, picks and adzes, probably involved in metallurgical activities performed at the site. Bifacials are common in proto-historic assemblages prior the EB, therefore they support a dating of the site to the end of the fifth millennium BC.

A common shape for tabular scrapers is elliptic (15 of the 25 complete pieces), from rounded to elongated oval. One item has a typical fan shape. The ratio length/width divides the tabular scrapers into three groups: items wider than longer, with a general rounded shape (N=3), items with an oval form, with L/W values of 1.1 to 1.7 and elongated items (N=4), with L/W ratios of up to 1.9.

The cemetery at Eilat (14.32/88.52)

The technological aspects of tabular scrapers also hint at the homogenous nature of the cache. All except four items have faceted butts. One item has a pointed butt, another one a flat butt. In two cases, the bulb of percussion was thinned. Hinged fractures are observable on six pieces, generally the larger ones. No incisions were seen on their dorsal face, generally completely covered by cortex. In three pieces, one flake scar (near the proximal end) is discernible. On these items, the dorsal face was smoothed. Retouch appears in most cases on the left side of tools, usually nibbled to fine. Occasionally, the retouch on the left edge is accompanied by nibbling along the right edge and the distal end. Most items have a partial retouch; the possibility that part of it is accidental cannot be excluded. Only 4 items have striation marks along the ventral side of the edge (the right edge in all cases). The general impression is that (at least) most tools were not utilized, given their mint condition.

The site, discovered by Uzi Avner in 1978, is located near the western suburbs of modern Eilat. More than 20 tombs of various shapes, and two open-air shrines were identified. The C14 of the site suggests a continuous use, covering the middle of the 6th millennium B.C. to the second half of the 5th millennium B.C. (Avner 1996). Following is a description of tabular scrapers, the main type of flint artefact collected within the circles of stones delineating the tombs. Other flint items are few flakes or chips. A polished axe made on magmatic black stone, found beneath a row of flat stones laid in front of one of the tombs complete the flint assemblage. Tabular scrapers were found in non-domestic context at a few sites: at Biq`at Uvda 6 a tabular scraper was found in an open-air shrine (Yogev 1983). A tabular scraper was the single item found associated with a feature described as an “altar”, at the site of Shiqmim (see above). Tabular scrapers were found also in burial context: a cache of several dozen in a tumulus at Ein Yarka (Rosen pers. comm.), a few in nawamis sites in Southern Sinai (BarYosef et.al. 1986), two in a burial cave at Azor (McConaughy 1979:219), two at Shiqmim cemetery (Levy and Alon 1987b) and a few in the burial cave at Peqi`in (Gal et al. 1997). Thus, the tabular scrapers cache found in Eilat tombs is not an unusual phenomenon.

The homogenous nature of the assemblage points at a single source of the items. Only one tabular scraper is distinct from the others by its raw material (darker, with a reddish cortex), the parallel retouch along its edges, the pointed butt and the thinning of the bulb. All these attributes suggest a different source for this item. Summary

All but one of the tabular scrapers are made on Eocene brown flint, with a white-yellow cortex. One item (technologically different) is made on a darker flint, with reddish cortex. Of the 33 items recovered, 8 are broken on the longitudinal axis and 4 laterally. Mean size are 112.5x76.4x6.9 mm. Length varies between 72.3 and 180 mm, most values being 90 to 110 mm. Five items form a distinct group, with length values up to 150 mm. Width is homogenous, with values ranging from 50.9 to 146 ± 21.3 mm. Two items are especially wide, one of them being different from the rest of the tabular scrapers also in other aspects (see below). Their thinness is

The tabular scrapers were discovered in a burial context. A similar cache was found in a tumulus at Ein Yarka (Rosen pers. comm.). It is important to stress the morphological and contextual similarity between these two assemblages. Although tabular scrapers are common in Chalcolithic-EB tool-kits, their appearance in burial / cultic contexts hint at another aspect of their nature, such as grave offerings, embedding a non-deciphered symbol.

97

CHAPTER VIII – DESCRIPTION OF RESULTS

several sickle-blades production centres, spatially (and perhaps temporally) located at various spots in this area (such as Safadi or Wadi Ghazzeh A), all using banded flint, support the assumption expressed above.

CHAPTER VIII DESCRIPTION OF RESULTS A major problem that arose during the interpretation of the results of this research was to identify the various factors that determined the nature of the flint assemblages, as they were available to us for analysis. Broadly, they can be divided into two types:

A different example is the use of semi-translucent flint in the production of microliths. Apparently, the choice for this type of flint is to be related to its characteristics, most suitable for the production of bladelets. In this case production centres were recognized as well, at some Wadi Ghazzeh sites and probably at Tel Sheva and Abu Matar as well. In these cases, we see a deliberate choice for a particular raw material for the production, mostly in specialized workshops, of single tool-types.

Site-excavation and collection methods of material culture (data acquisition), as exemplified at Wadi Ghazzeh sites, and methodological methods of data interpretation and presentation, such as the case of the Golan sites (to mention just a few examples), both factors making a synthesis of the material and its inclusion in a broad framework of comparison a difficult (and sometimes an almost impossible) task. Thus, for each case where these factors may have had an influence on the analysed assemblage, only a limited number of observations was performed on the lithic material.

Are these examples to be related to cultural factors influencing decision-makings of Chalcolithic knappers? A possible answer would be that both are the results of cultural factors, though their origin was different: in the case of banded flint, its choice could have been accidental, while semi-translucent flint was deliberately chosen for its characteristics. If such is the case, the intensity of appearance of these phenomena may be used as indicator for the nature of cultural ties between sites.

Archaeological factors: taking into consideration postdepositional processes that affected the nature of each assemblage and evaluating their influence, how to interpret the results of the analysis – in other words, how to isolate the mechanisms that shaped lithic assemblages: cultural, social, economic or simply the particular choices and skills of the artisan whose result we analysed.

Another variable that can be interpreted in various modes is the exploitation degree of cores, estimated according to the débitage surface that remained unused (the smaller the débitage surface unused, the higher the exploitation degree). If this phenomenon occurs in an area scarce of suitable raw materials and their availability is low, it is expected to find a high exploitation degree of cores. Another factor that may influence this variable is the standardization of production, as a result of craft specialization: a specialized industry will be more efficient and subsequently, a higher exploitation of cores will be expected (intensification of production). Theoretically, craft specialization will be also reflected in an efficient knapping and mode of manufacture, which in turn may be signs of high technical skills.

A scheme for the results’ interpretation is presented in Figure 8. The presence of tools made on exotic (non local) raw materials, may be the result of economic factors, such as the existence of an exchange network, which in turn requires a production centre, specialized in the manufacture and distribution of the traded product. However, if products have a symbolic meaning, often materialized by their context, their production on exotic raw material may be the result of this aspect, and not necessarily economic ones (for example, Rosen’s (1987) model for the economic implication of tabular scrapers, versus their appearance at the burial site of Eilat). The production of particular tool-types, by a specific knapping method and on selected raw materials may be the result of an a priori knowledge of the knapper, regarding physical characteristics of those raw materials and their availability, but also they may be the result of cultural/traditional constraints that dictated the nature of raw materials’ exploitation. For example, the shape of a typical Northern Negev Chalcolithic sickle-blade (narrow, backed, double truncated and finely denticulated along its working edge) may reflect the sickle’s shape itself and continuous adjustment towards its better exploitation (a prolonged use). However, the constant choice for banded flint as the main raw material for the production of sickle-blades in this area should be assigned not to a particular choice of the knapper, nor to physical characteristics of this type of flint (a regular pebble) nor to its abundance (at some sites, such as Grar, it is absent from the nearby wadi), but probably to cultural enforcements. Moreover, the recognition of

The problem may be clearer when discussing the case of the workshop for the production of sickle-blades, discovered at Safadi. The lithic remains, found in one concentration, clearly belonged to one or possibly more knappers, specialized in the production of sickle-blades, from their first phase of modification and until the last one, probably resharpening used pieces. Thus, the artisan probably went to the adjacent wadi, collected its suitable raw material (in our case – banded flint) and modified it using a rigid knapping sequence, probably the most fitted one for the production of blades, knapped from pyramidal cores that were exploited on half of their potential débitage surface. These blades probably required a standard size (as reflected by a low variability of sickleblades size) and shape (further on, these blades were truncated, using a microburin technique, backed and finely denticulated). The remains of the workshop contained a high amount of discarded blanks, at various

98

CHAPTER VIII – DESCRIPTION OF RESULTS

characterized by their small size, lack of architecture, few installations and a general impression of an ephemeral occupation, expressed in the scarcity of remains of material culture. In contrast, Safadi and Tel Sheva were apparently occupied for a prolonged period, given the impressive architectural remains (both above and under ground structures) and a large and diverse amount of material culture. Apparently, the inhabitants of the two groups of sites had a different life-style, implying different economic and possibly socio-politic organizations. However, the lithic assemblages of these sites are similar, in terms of applied technology, tool-type diversity and their relative amount, despite differences of other cultural remains, such as pottery.

modification stages (N=3708), and cores (N=69) that could have a more prolonged use. Thus, facing what is clearly a workshop, several aspects are discernible: • A low exploitation degree of cores (high availability and abundance of raw materials), • A high standardization of cores, blanks and tool morphology, • A high technical skill reflected in the adoption of the microburin technique for controlling the length of blades, • The efficiency of production is not necessarily high, given the high amount of discarded, unused, blanks and the low exploitation degree of cores

Two extreme explanations derive from the analysis of this phenomenon: flint assemblages are not sensible to socio-economic factors active in Chalcolithic cultures, recognizable in other aspects of material culture, or, socio-economic organization, as related to lithic assemblages, was similar in “ephemeral” sites (Ramot 3) and more complex habitation centres (Safadi).

Thus, we can see that in a clear case of concentration of production, which may reflect craft specialization, not all hypothesized phenomena occurred. The question is how to identify craft specialization, or production centres, in case the archaeological evidence is unclear? The answer may be that only in a few cases, a spatial distribution of artefacts, and not necessarily their detailed technological study, may lead us to conclude that we identified a workshop, and subsequently, a socio-economic phenomenon of craft specialization.

Finally, before presenting the results of the comparison between the assemblages described above, some general information regarding sites and their lithic material is presented. The data used for comparison originates from 48 sites, extending over a geographic area that includes modern-day Israel and Jordan, the Jordan valley and Sinai. They were excavated in various periods, most of them between the 70’s and the 90’s of the last century, while a fifth of them were collected in the last ten years. The lithic assemblages of half of these sites was directly analysed, comprising ca. a third of the data-base; the information regarding the others was collected from their publications. The amount of flint artefacts included in the data-base is 220,166 items, among them 18,975 tools, 99,393 débitage products and 101,798 debris pieces.

Another example is the lithic assemblage of Grar, a site located close to the source of raw materials, the wadi of Nahal Grar. When its blanks and cores are compared to those of Nahal Be’er-Sheva area, there is a clear dichotomy between them: Grar products are consistently smaller and cores are more exploited. This phenomenon cannot be linked to raw material characteristics or procurement strategy, since both Grar and Be’er-Sheva sites are located along wadis in which similar pebbles can be found. Moreover, it is improbable that the economy of Grar’s inhabitants was more concentrated on flint tool production (no workshops were traced) and at both sites knappers prepared morphologically and technologically similar tool kits. Therefore, the remaining explanation can be linked to the general term savoir faire, which implies different traditional (and possibly diachronic) roots of the inhabitants of these two clusters of sites.

The sites incorporated in this work comprise most of the Chalcolithic remains found in the Southern Levant, from which lithics were reported. Assemblages from sites located on the territory of modern day Syria and Lebanon were omitted, given the political restrictions that exist today and as a consequence limited access to material. It may be assumed that some of the conclusions drawn from this work are valid for certain parts of those regions as well. Assemblages included in this work but not analysed directly were usually not attainable or their publication was satisfactory for the purposes of this work.

Another factor, presumably affecting the nature of flint assemblages, was the site’s characteristics. Aspects such as site location (proximity to raw materials or to other sites), site function (habitation, workshop, cultic, etc…) or site life-span (duration of occupation) would influence the nature of their flint assemblages. An example of a clear link between site nature and its lithic assemblage is the Eilat burial site, where most items were tabular scrapers, apparently brought on-site as grave-goods.

Following the current state-of-art in Chalcolithic research (Chapter II), the analysed assemblages were divided into 5 main groups, according to their geographic location. The assemblage of the site Nahal Komem (or Wadi Zeita, according to first publications), was described separately, given its probable time location, at the end of the Chalcolithic period or the very beginning of the EB. This classification enabled the validation of the existence of

The relation between site type and its lithic assemblage is less clear in other cases, for example when comparing the flint assemblages of sites such as Ramot 3 or Meitar, to assemblages collected from larger sites in the same area, such as Safadi or Tel Sheva. The first two sites are

99

CHAPTER VIII – DESCRIPTION OF RESULTS

there is an apparent increase in the number of exploited raw materials, “canaanean flint” marking its first, probably deliberate, appearance. However, other aspects of raw material procurement strategy and manipulation are similar with other Chalcolithic industries.

these regional culture groups, to isolate each lithic characteristic and to locate it within the cultural mosaic of the Chalcolithic period. A comparison between the groups allowed us to identify possible relations/connections between them and to suggest a possible time-location for each group, based on absolute dating and lithic characteristics.

Southern sites were found in areas with scarce lithic resources; in south-eastern Sinai assemblages, two flint types were modified: Eocene pebbles to produce mostly flake tools, while the other, of a higher quality, was used to produce smaller tools. A third type, tabular flint, restricted to fan scrapers, was imported from an unknown location. The selection of the first two types of raw materials was probably forced by limitations of flint sources and movement patterns of the human groups, along which these flint types could be procured.

The characteristics of the lithic industries were divided into three main groups: strategy of raw material procurement and its use, technological aspects of tool production, and techniques and characteristics of particular tool-types. Aspects of intra-site spatial analysis were mentioned only generally, given the fact that in most cases Chalcolithic inhabitants disposed of their remains in dumping zones, pits of various size, shapes and previous functions (such as the bell-shaped pits or underground structures).

It is clear that location of sites was not related to flint sources, their inhabitants probably carrying with them blocks of flint or semi-prepared cores. Whenever habitation sites were located close to flint sources, there are more exploited raw materials, such as at some Southern Jordan sites (J408, J521), where semitranslucent flint was on-site exploited to produce bladelets. A flint exploitation pattern related to site type (and transhumation movement) was observed: knappers of long-term sites used a few raw materials (J11, J14), reachable in the nearby areas, while at ephemeral sites more flint sources were used (J521 – up to 6 types). An interesting experiment is to trace flint sources, and thus reconstruct the procurement strategy, which would lead to a drawing of the movement patterns.

The geographic groups which presumably delimitate also regional cultural entities are, from north to south: I. The Golan area, extending over Northern parts of the Jordan valley, on both sides of the river. II. Northern Israel, extending over the Upper and Lower Galilee mountains. III. A central zone, very poorly defined, mainly due to a lack of assemblages available for analysis in a clear archaeological context. IV. The Be’er-Sheva valley, extending up to the Dead Sea basin. V. the Western Negev area, including north Sinai; VI. A Southern zone, with sites in Southern and eastern Jordan, the Arava valley and south-east Sinai.

Northern and western Negev sites are mostly located on the bank of wadis, in areas abundant with raw materials. A selection of particular raw materials could be noted: banded grey ones were used as cores for the production of sickle-blades (Safadi, Tel Sheva, Grar, etc.). Since its characteristics do not differ drastically from those of other pebble flint, it is suggested to regard this phenomenon as a cultural mark; at a certain point in time, Chalcolithic knappers of the Negev chose to use banded flint for the production of sickle-blades and thus a tradition was established. Given the fact that sickleblades were produced in regional centres, as discovered at Safadi or Gazzzeh A, the establishment process of this tradition was probably short and its impact prolonged.

In the following paragraphs each of the analysed parameters will be described, highlighting factors that influenced their value, such as environmental, socioeconomic, cultural or personal influences of knappers. This presentation will validate or falsify the presence of the cultural entities proposed above. Consequently, the lithic industry of each entity is presented. Comparison results Stage 0: Acquisition – extraction and testing of nodules Types variability and Abundance

The same is true regarding semi-translucent flint, used for the production of microliths, also in production centres (Ghazzeh A, Tel Sheva). However, its choice was probably dictated by its characteristics, more suitable than pebble flint for the production of bladelets. These traditions had probably a high impact on the society, given the fact that they were found in all western and Northern Negev, including regions where banded flint is not easily obtainable, as at Grar, its nearby wadi lacking banded flint.

There is a gradual change in the number of raw material used for the production of flint tools, with a geographic trend: from 2 to 3 types of flint used in Southern, desert assemblages (NE Sinai sites), to 4-6 in the Negev (as at Safadi or Tel Sheva) and the central parts of Israel (Kh. Aliya), to 9 types of flint exploited by Northern knappers (Tel Turmus). This variability was probably imposed by environmental characteristics, mainly availability of raw materials and their accessibility. At Nahal Komem, a site occupied towards the end of the period (ca. 3500 B.C.)

100

CHAPTER VIII – DESCRIPTION OF RESULTS

Knappers of central Israel (Kh. Aliya) chose a limited range of raw materials to produce their tools, mostly a semi-translucent Senonian flint, applying a similar acquisition strategy to the one of the Negev sites. The nature of this flint influenced the production of tools, most of them on small blanks (sickle-blades on bladelets is one characteristic of this assemblage). The presence of (a small amount of) sickle blades on banded flint in these assemblages suggests a link that existed between the groups – probably an exchange of end-products, sickleblades being imported to this area. As shown below, typo-technological aspects exclude the possibility that this group belonged to the one described above, but probably existing as an independent cultural entity.

materials were modified within the site’s area. The same is true for the site of Kaukab, where most tools were onsite produced, from local raw materials. A different conclusion is proposed for the similar phenomenon of local modification of raw materials at Abu Hamid in the north and at Wadi Ghazzeh site A. Their common characteristic is the presence of workshops, specialized in the production of few tool-types. The first explanation was excluded, given the fact that in other aspects these sites do not differ from other sites of the relevant areas, the only outstanding characteristic being the presence of workshops specialized in flint tool production.

Northern assemblages (including those in the Golan plateau) are characterized by the exploitation of a high amount of flint types, between five and nine, of different sources and qualities (Kaukab, Tel Turmus). Semitranslucent flint and tabular flint were used mostly for the production of microliths and fan scrapers respectively, but not solely, as occurs in Northern Negev sites; there is no particular or exclusive preference for a particular flint type for the production of individual tool-types. A tendency to use “cannaenan-like” flint for the production of blades was noted at Tel Turmus, a possible precursor sign of the following period, the EB.

Cores represent 3% to 10% of débitage at most sites, except ephemeral sites of Southern Jordan and one basecamp (at J11 cores represent 6% of débitage). The low amount of cores at ephemeral sites may be related to the high mobility of the human groups and the activities performed there.

Stage 1: Production

Flake cores Flake cores dominate the lithic assemblages of most sites, representing 70% to 90% of cores (Table 3). Virtually only flake cores were found in desert sites, indicating a restricted number of knapping methods.

The variety of raw materials points for different sources – habitation sites not being located in close vicinity to flint sources, several quarries were exploited, such as wadi pebbles, mountain outcrops or terrace conglomerates. Occasionally, flint might have been brought to the site from distances bigger than 15-20 km. If its quality allowed, the same raw material was used for flakes, blades and bifacials. The absence of any well articulated relation between raw material and tool-types indicates a lack of a cultural tradition regarding this aspect, or rather the factor that dictated which raw material to be used was related to its availability and quality.

At most sites, flake cores have one striking platform, except at Meitar and Tel Sheva, where flake cores have mostly two striking platforms, probably the result of exploitation of a low quality of flint (Meitar) or the application of a different reduction sequence for the procurement of flakes (Tel Sheva). Flake cores with two striking platforms are common also at Tel Turmus, probably a result of the use of a poor quality of flint, which required an extensive preparation. The possibility of an extended use of suitable pebbles for the production of cores, as a result of scarcity of raw materials at the above mentioned site cannot be excluded as well.

Use diversity and Locality of modification

The raw materials used for the production of flake cores vary from a few in Southern, desert sites, to 6-8 in Northern sites. There is no particular choice of raw materials for flake core production, flakes being made on all available raw materials.

Two main factors influenced the use diversity of raw materials and their locale of modification: a “natural” one, concerning physical characteristics of flint and its source, and an “artificial” factor, i.e. anthropogenic, with two possible facies – economic, implying the presence/absence of exchange networks and workshops, or cultural/traditional, related to restrictions regarding the use of particular flint types, such as banded flint in the Be’er-Sheva sites. In most cases, regardless of geographic location, 75% of raw materials were processed within the sites, for most of the tool-types; the remaining ones were imported on end-products (as tabular scrapers). Exceptions are ephemeral sites in Southern Jordan, some located in close vicinity to flint sources and with a tradition of flint knapping that did not require the use of specific raw materials, where all raw

The exploitation degree of cores was evaluated by two variables: the ratio between blank and core size, and the amount of cortex on the circumference of the core (Table 3). There is no marked difference between the exploitation degrees of flake cores of the various assemblages; flake cores were discarded after a desired size limit of blanks (as reflected in the ratio mentioned above, with values of 0.91 to 1.2 at all sites). Furthermore, flake cores retained some cortex on their débitage surface (between 24 and 26%). The exception is Tel Turmus, where flake cores were cleaned of cortex, probably as a result of the poor quality of flint.

101

CHAPTER VIII – DESCRIPTION OF RESULTS

The flake core initialization was estimated by two ratios: tools on flake / flake cores, and amount of flakes (tools and blanks) / flake cores. The results showed various results: flake cores exploitation is more accentuated at Southern Jordan sites defined as ephemeral (with values of 2-3 as at J408 and J521), than at long-term sites (values of 0.7-1, as at J11 and J14). Apparently, knappers of ephemeral sites used more types of raw materials and exploited flake cores in a higher degree than those of long-term camps. This difference may be related to the nature of sites: at long-term camps, apparently located near flint sources, knappers exploited the locally available flint, which did not require an extensive exploitation, whereas at ephemeral sites the exploitation pattern was different: cores were extensively used.

Apparently, a non-determined flake was the aimed blank for the production of most tools at the analysed sites. Therefore, differences in core size cannot be explained in terms of differences in the desired product. An intense flake cores exploitation occurred at Grar, Tel Turmus and Kaukab and Tel Sheva, as reflected in the values of core initialization (Table 3). Moreover, the small size of pebbles utilized by knappers in the north (Tel Turmus and Kaukab), and their low quality, may have influenced the size of flake cores. A low flint quality was observed at Meitar and Ramot 3 as well, while pebbles used by Grar knappers may have been smaller than those found in the Be’er-Sheva basin.

Another difference was observed when sites of western Negev were compared to those of the Be’er-Sheva basin: knappers of western Negev sites (as at Grar and R45) exploited more intensively flake cores for the production of blanks, further modified into tools, than those of the Be’er-Sheva basin (ratios of 10-12 as opposite to 1-3). However, when the ratio flake and flake tools by flake cores is compared, the picture is less clear, practically at each site the result being different. Therefore, assuming an on-site production of flake tools, it may be suggested that the knappers of Grar and site R45 produced more tools on flake related to flake cores, than their neighbours of the Be’er-Sheva basin, despite similar conditions of raw material availability. This discrepancy may be the result of a cultural/traditional difference.

The data regarding blade cores is fragmentary, therefore only limited information is available for interpretation (Table 3). Blade cores are rare at sites in the desert, an expected situation given the fact that tools on blades are not common in these assemblages. At all other assemblages, blade cores represent between 30% and 40% of cores. Exceptions are R79, Abu Matar and Tel Sheva (blade cores represent ca. 4-10% of cores) and Mitcham C and Ghazzeh sites O and M (60% of all cores). At R79, Abu Matar and Tel Sheva, a low production of blades was observed. The situation was apparently different at Mitcham C and Ghazzeh sites O and M, where blade cores dominate the cores group. This primacy may be the result of workshops for the production of blades, the excavations at these sites focusing on collecting the remains of these “ateliers”, either due to limitations of excavated area (Mitcham C) or collecting methods (Ghazzeh sites).

Blade cores

Core initialization varies greatly at other sites; practically, at each site the result is different, therefore clustering is impossible. At some sites, such as Kh. Aliya and Kaukab, there are more flakes struck off cores and less tools on flakes, related to the amount of flake cores (ratios of 24 and 2-4 respectively). At others, such as at Tel Turmus or Nahal Komem, less flakes were struck-off cores (ratio values of 11-13), but there are more tools on flake related to flake cores (ratio of 5 at Tel Turmus).

The exploitation degree of blade cores shows an apparent rigid reduction scheme at sites of western and Northern Negev: cores of these areas are usually made on banded flint, have one striking platform, half of their circumference remained cortex covered and they were not exploited beyond a fixed length (L=40-60 mm, W=40-50 mm). Probably most of these cores were used for the production of sickle-blades. Important to note is the presence, though in small amounts, of blade cores on banded flint at sites where workshops were not localized (as Ramot 3 or Meitar), but still very similar in shape and size to those originating from workshops (as Safadi or Ghazzeh A).

A possible interpretation of the picture presented above is that core initialization is influenced by many independent factors, and therefore can be only evaluated at the site level, each case behaving in a different way, given its particular conditions. The size of flake cores may have been influenced by the original size and quality of raw material, shape and size of desired blank and intensity of core exploitation. Several differences were observed: cores of Be’er-Sheva sites Abu Matar and Safadi are the largest (L=60-62 mm, W=64-66 mm), while Grar flake cores and those of Northern assemblages (Tel Turmus and Kaukab) are the smallest (L=33-37 mm, W=34-41 mm). In a middle position are cores of Meitar, Ramot 3 Tel-Sheva and R79 (L=40-50 mm, W=56 mm).

The reduction strategy applied on blade cores by knappers of Northern sites, such as Kaukab or Tel Turmus is different than the one described above. Apparently, the reduction scheme is more flexible: more raw materials were used to produce blades, cores have one or two striking platforms and there is less cortex left on cores’ circumference (25%). Cores are also smaller than those of Negev assemblages (L=37-44 mm, W=3542 mm), probably as a result of the original size of raw materials.

102

CHAPTER VIII – DESCRIPTION OF RESULTS

Core initialization is another variable that changes geographically: while at the Negev sites there is a low amount of blade blanks and blade tools (the ratio [blades+tools on blade]/blade cores, returned values between 0.3 to 4), blade cores were more intensively exploited at sites in the north and the centre (the value of the above mentioned ratio varies between 14 to 22). This difference may be related to two main factors: one is the scarcity of high quality flint in the north, which imposed an intensive exploitation of cores; the second possible factor is the presence of workshops specialized in the production of blade blanks at sites in the Negev, such as Safadi, Ghazzeh A and possibly Mitcham C, which caused a high number of blade cores, their products being exported.

Technological indices Several ratios between waste groups were calculated, in order to characterize the nature of the flint industry (Table 26). At most sites, cores were not rejuvenated more than once, as reflected by the ratio [CTE+tools on CTE]/cores, which returned values between 1 and 2 for most of the assemblages. however, several assemblages are different: at Meitar and Ramot 3, the ratio has values of 5 and 8 respectively, probably being influenced by the raw materials used at these sites (of low quality, being mixed with chalk), which required cores rejuvenation for a better exploitation of cores. Sites Ghazzeh A and B are different as well (ratios of 10 and 30 respectively), probably the products of the workshops for the production of blades and bladelets located there, blade and bladelet cores requiring some preparation of platforms and débitage surfaces appropriate for knapping blades and bladelets.

If indeed Nahal Komem represents a late occupation, towards the end of the Chalcolithic period, several characteristics of lithic industries can be defined: blade cores represent almost half of all cores, reflecting an emphasis on the use of blades as a main support for tool production. Another aspect is the increase of core exploitation, blade cores on “cannanean” flint being characterized by two or more striking platforms. There is also a shift in the strategy of raw materials procurement and use, “cannanean” flint becoming the preferred raw material.

The decortication stage of cores was evaluated by the ratio [primary elements+tools on primary elements] /cores. At most sites, this ratio gave values spanning from 1 to 5, which means that a few primary elements were sufficient to clean the cortex from the core’s débitage surface. There are some anomalies to this rule: at site J521 the value is 5, outstanding among other sites of the area, where the values are around 1 (as J11 or J14). Probably the site was located near flint sources, all the process of core decortication and initialization being performed onsite, while at other sites cores might have been cleaned from cortex close to their quarry zones. At the opposite pole is the site of Tel Turmus, where the value of this ratio is less than 1, suggesting that the decortication of cores was performed off-site, probably near the acquisition spot of raw materials. The same is true regarding sites in Northern Sinai, as R79 and R48 (their ratio values are 0.2 and 0.6 respectively).

Bladelet cores The exploitation strategy of bladelet cores is similar at most sites: the preferred raw material is a semitranslucent flint, cores have one striking platform and a few cortex on their circumference. The assemblages of Kaukab and Tel Turmus differ again, several raw materials (6 and 7 respectively) being used to produce bladelets. About a ¼ of all cores were used to produce bladelets; exceptions are desert sites, where they were found in lesser amounts (few to none in south-eastern sites, few in the south Jordan sites) and Be’er-Sheva sites (none at Safadi and few at the others). Exceptional are the assemblages of Tel Sheva and Mitcham C, where bladelet cores represent almost a third of cores, probably remains of a workshop specialized in the production of microliths, such as at Ghazzeh site A.

The ratio [blanks+tools]/cores indicates the exploitation degree of cores. Even though a value of 20-30 tends to represent most assemblages, the values of this ratio vary greatly between sites of the different geographical areas, and within each geographic zone. Most of the desert sites of Southern Jordan have a value around 20, with two extremes: site J24 (15) and J14 (41), both long-term sites, but with an apparent different exploitation rate of cores.

At most sites, a few bladelets and microliths were collected, related to the number of bladelet cores; exceptions are two sites of Northern Sinai (R45 and R79, where the ratio [bladelets + microliths ]/bladelet cores returned a value of 10). This picture reflects a particular activity performed with microliths, absent or practised in a lesser extent at other Negev sites. The size of bladelet cores is similar at most sites (L=23-33 mm, W=24-40 mm), suggesting a common source of raw materials. Outstanding are the bladelet cores of Ramot 3 and Meitar (L=38-41 mm, W=36-45 mm), suggesting a different source of raw materials, or an abundance which did not impose an extensive exploitation of bladelet cores.

Sites in Northern Negev may be divided into two groups: those where the ratio is low (between 16 and 40 – all located in the Be’er-Sheva basin) and sites with a very high value (109-118 – Ramot 3, Meitar and Mitcham C). The high value of this ratio at Mitcham C may be explained by the fact that the flint assemblage may represent mainly remains from a workshop, where cores were extensively prepared and exploited).

103

CHAPTER VIII – DESCRIPTION OF RESULTS

and Ramot 3 (15 and 22 respectively). Sites of the western Negev are characterized by an increase in the use of blades as support for the production of tools, especially at Ghazzeh A and R79, where a local production of sickle-blades influenced the value of this ratio. Outstanding is site R45, where tools on blades are rare. Sites in the centre (Kh. Aliya) and in the north (Abu Hamid and Tel Turmus) are characterized by a more balanced production of supports for tools, their values of this ratio being between 1 and 3. Nahal Komem has among the lowest values of this ratio (0.9), emphasizing the increased use of blades as a main tool blank, a phenomenon common for EB assemblages (Rosen 1997).

Two main factors differentiate these groups: Be’er-Sheva basin sites are large, with under and above ground architecture and a varied material culture, while Ramot 3 and Meitar are apparently ephemeral occupations, without architecture and a restricted material culture. Subsequently, knappers of the Be’er-Sheva basin used wadi pebble flint, easily obtainable from the nearby wadi, while a poorer quality flint, mixed with chalk and originating from the dislocation of the exposed bedrock was used by the knappers of Ramot 3 and Meitar. These factors dictated the exploitation degree of cores: many removals were needed to obtain suitable blanks from a low quality flint, while a short occupation of a site (such as at Ramot 3) may enforce a high exploitation of cores, knappers preferring to use and re-use the collected chunks instead of searching new raw materials. The high value at Shiqmim may be the result of a different classification of waste products, which included chips into the flake group and increased the group of blanks.

Stage 2: Production of blanks Primary elements The relative amounts of primary elements change according to the geographic location of assemblages. Sites in the desert (Southern Jordan sites) have a few primary elements (1-6% of débitage), suggesting that core decortication was performed off-site, probably close to raw material sources, apart J521, where primary elements represent 10% of débitage, a result of site’s location, close to raw materials sources. Primary elements of Northern Negev and Grar assemblages form from 11 to 26% of débitage, reflecting an on-site decortication of pebbles; Ramot 3 and Meitar are exceptional, with a low amount of primary elements, probably due to the raw materials used, flint exposures in the natural bedrock. North Sinai sites are characterized by a low amount of primary elements, of 1% to 6% of all débitage, reflecting an off-site decortication of nodules. The assemblages of Kh. Aliya and Tel Turmus have similar amounts of primary elements, hinting at a decortication of raw materials close to their sources. A different behaviour was adopted by Kaukab’s knappers (primary elements represent 21% of débitage, similar to Northern Negev assemblages), the decortication of cores being performed on-site.

Sites in western Negev are also divided into three groups: Grar (value=15), R48 and R79 (values of 34 and 37) and R45 (value=96). These differences are difficult to interpret, mostly the high value of R45, which is similar in all aspects to others of the area. The very low value of Ghazzeh A (4) is probably the result of selective collection methods, which focused on cores and less on flakes or primary elements. The ratio [flakes+primary elements]/cores was calculated in order to identify the orientation of the flint industry. All industries, regardless of their geographic position, are clearly oriented towards flakes production. The ratio has values of ca. 5 at most sites, with a few exceptions: 2 ephemeral camps in Southern Jordan, (J408 and J521, where there are more microliths than at other sites) have values of 1 and site Ramot 3 in Northern Negev, with a value of 22, probably as a result of the ephemeral nature of the site, with less emphasis on the production of blades. In general terms, the sites of the Be’er-Sheva basin have the highest values of this ratio (12-15 at Safadi and Abu Matar), while sites in the western Negev (Grar, R79) and in the north (Abu Hamid and Tel Turmus – value = 4) produced relatively more blades. The low value at R79 is explained by a local production of blades, probably used as blanks for sickle-blades.

Flakes Flakes dominate the waste products of all sites, representing in most cases more than 50% of débitage. Exceptions are two sites of Southern Jordan, where the lithic production was oriented towards the manufacture of microliths (J408 and J521). The scarcity of flakes at these sites may be influenced by their nature, ephemeral stations during the transhumance cycle – the inhabitants practised a limited number of knapping activities, focusing on the production of microliths. This variation is viewed as a marker of a wide range of activities performed during the movement of human groups, rather than a cultural difference, since other aspects of material culture relate these sites to others of the area. It may be concluded that the Chalcolithic flint industries of Southern Levant were oriented towards the production of flakes, which were the main blank for tool production.

The values of the ratio tools on flakes/tools on blades vary according to geographic zones. At the desert sites of Southern Jordan, flakes and blades were almost equally used as supports for the production of tools, especially at ephemeral sites (as J408 and J521), where microliths were extensively used. The other group of desert sites, in eastern Sinai, is distinguished by a low amount of tools on blades, while microliths are absent for all except one site (see below). Sites in the Northern Negev are characterized by a low amount of tools on blades, related to those on flake (values of 5-9 at the Be’er-Sheva basin sites) and extremely low amounts at sites such as Meitar

104

CHAPTER VIII – DESCRIPTION OF RESULTS

The knapping technique of blades was similar in most assemblages: a direct blow with a hard hammer, resulting in a flat butt. A second technique, indirect blow with a soft hammer, was applied for the production of ca. a 1/3 of blades at most sites. Exceptions are Ghazzeh sites A and B, where pointed butt dominate, as a result of a massive production of bladelets in workshops. The low amount of cortical butts suggests that the removal of blades commenced after preliminary stages of core decortication. Blades with a faceted butt are rare at most sites, except Tel Turmus (23% of blades) and Ghazzeh A (26%). While the first case may be explained by an attempt of knappers to facilitate the removal of blades from low quality flint, the second case may be related to a standardization of blades and bladelets production.

The sizes of flakes are similar at most sites (length 31 – 34 mm; width 29 – 31 mm), exceptional being Northern Negev sites, where larger flakes were knapped (length 50 – 60 mm; width 44 – 50 mm). This difference derives probably from the reduction pattern of wadi pebbles, used for the production of flake tools – knappers of Northern Negev chose large pebbles to produce flakes, whereas in other areas, smaller pebbles were modified into flake cores. This difference may reflect a cultural-traditional divergence between Northern and western Negev sites, given the fact that sites of these areas are located along wadis of similar drainage system, where pebbles of similar size are expected to be found. Concerning sites of other areas, the size of their flakes was probably influenced by the shape and size of raw materials, usually small and rolled pebbles, as at Northern sites.

The scar patterns on the dorsal face of blades shows that most were struck-off unipolar cores (50-100%), bipolar cores being more common in Northern assemblages (Kaukab and Tel Turmus – 20% of blades). This dichotomy may be related to a difference in the raw materials as well and the degree of systematization of blades production (see above). The number of scars shows that blades were knapped at later removal stages from cores, most having 3 to 6 scars on their dorsal faces.

The variability of butt types reflects an inconsistency in knapping methods among sites; in Northern and Western Negev sites there are more flakes with a cortical butt, suggesting that pebbles were modified on-site, without a long sequence of core preparation. Opposed to this strategy is the one adopted by Northern knappers – as a rule, flakes of Northern assemblages lack cortex on their butt, probably due to a low quality of pebbles, which required a more extensive preparation prior to their exploitation of cores, from which potential supports could be knapped. The possibility that this preparation stage was performed near the flint source cannot be excluded.

The ratio blades/tools on blade is indicative of the “efficiency” of blade production, i.e. how many blades were suitable for the production of tools and, additionally, import/export of blades or blade tools to and from the site. At most sites, this ratio has values spanning between 2 and 5. Low values were observed at Tel Turmus and north Sinai sites R48 and R45 (0.8, 0.1 and 0.3 respectively), suggesting that either most tools on blades were imported on-site as end-products, or most blade blanks were modified into tools. Given the fact that blade cores are rare to absent in the north Sinai sites, the possibility that tools on blades were imported to the site as end-products cannot be excluded. Blade cores represent 23% of all cores of Tel Turmus. Therefore, it is suggested that the low value of this ratio may be the result of an extensive use of available blade blanks, without a special requirement of size and shape. The high value of the ratio at Safadi (10) reflects the discarded blade products of the workshop for the production of sickle-blades, discovered at the site.

Blades The relative amounts of blades vary according to the geographic position of sites and their nature (Table 28). In Southern Jordan, the débitage of ephemeral sites is dominated by blades, while in long-term camps blades represent ca. 15% of the débitage. Sites of Northern Negev have a low blade production (6% to 10% of débitage), with two extremes: Ramot 3 and Meitar (3% and 5%) and Mitcham C at the opposite pole (16%). In the first case, the low amount of blades was probably influenced by the nature of the sites, while in the later, the presence of a blades production workshop may have influenced their high appearance. Northern (Kaukab and Tel Turmus) and centre (Kh. Aliya) assemblages have a higher relative number of blades than Northern Negev assemblages (13, 18 and 12 respectively), a characteristic which may be related to cultural/traditional factors.

Stage 3: Shaping/Retouching stage – retouching of tools

A regional dichotomy regarding the number of raw materials used for the production of blades was observed: while in the Negev assemblages three flint types were usually exploited, in Northern assemblages this number rose to 5 (Kaukab) or 7 (Tel Turmus). A possible explanation is related to the standardized production (usually in workshops) of sickle-blades in the Negev, while in the north the production of sickle-blades is less rigid, sickle-blades being manufactured on various raw materials and their shape varies greatly.

The number of tool-types vary between 10 and 11 at most sites; desert areas sites (Southern Jordan and eastern Sinai) have few tool-types (4 – 7), while sites in the north (Abu Hamid, Golan sites and Tel Turmus) stand out by their high amount of tool-types (13 – 14). The low variability of tool-types in the desert areas is related to the limited activities performed with flint tools in these areas, probably as a result of environmental restrictions, which imposed limitation of subsistence activities.

105

CHAPTER VIII – DESCRIPTION OF RESULTS

Two tool-types characterize northern assemblages, being virtually absent from other zones: perforated discs and pièces esquilles. While the first is a regional cultural hallmark, i.e. a local “invention” of Chalcolithic knappers, the second type may result from the use of an anvil during the process of tool manufacture, a product of a knapping technique, and not a tool-type per se.

Scrapers Scrapers are common in all assemblages, regardless their provenance; most are simple end-scrapers on flakes, the working edge being modified by semi-abrupt retouch. The relative amount of scrapers from desert assemblages is 3% to 37%. This difference was probably caused by different subsistence activities performed at each site.

Tool-types were divided into: ad hoc (flake tools, their production requiring no special knapping method – scrapers, burins, borers, truncations, notches and retouched flakes) and all others (non ad hoc tools), products of particular knapping methods (retouched blades and bladelets, sickle-blades, bifacials and tabular scrapers). The frequency of non ad hoc tools vary by geographic location of sites. Desert sites have a low amount of non ad hoc tools, a result of a limited range of activities performed with flint tools, the demand being of simple and easy to prepare tools. Exceptions are Wadi Sa’al and Wadi Wattir, where tabular scrapers were found in an apparent cultic context (masseboth and open-air shrines). Two sites in Southern Jordan (J11, a long-term camp and J408, an ephemeral site) have a high percentage of retouched blades and microliths. Given their different nature, this similarity is hard to explain. The possibility that a same activity, performed with microliths (hunting?), cannot be excluded.

The amount of scrapers in Negev assemblages is 20% 35% of tools, while in centre and Northern assemblages (Kh. Aliya, Tel Turmus, Golan sites and Kaukab), they form 10% of tools. This dichotomy may be the result of a different approach to scraper use, the frequency of their manufacture, use and re-use in Negev assemblages being higher than in the north. Another possibility is a different use of scrapers: while in the north they might have served a specific task, in Negev assemblages they were used for a multitude of functions. Yet another explanation is related to economic factors: a specific activity, performed with scrapers, was among the main ones in Negev sites, while in the north it was less practised, or practised with other tool-types. Given a general similarity of economic organization both in Northern and Negev sites, the first two suggestions may explain this dichotomy. Outstanding among scrapers are the “Be’er-Sheva” type (see Chapter II). They were only found in the Be’erSheva basin, being representative of sites of this area.

Sites in Western Negev (North Sinai sites and Grar) have a tool assemblage almost equally divided between non and ad hoc tools; microliths and blade tools dominate the former. The high amount of microliths may reflect a cultural tradition, or a specific task performed with them.

Tabular scrapers Tabular scrapers have similar amounts (1 – 3% of tools) in most assemblages, regardless of their geographic location. Exceptions are a few desert sites (W. Sa’el, W. Zalaka and W. Wattir), where they represent ca. a 1/3 of all tools. Apparently, tabular scrapers had a non-domestic use at these sites, being found in cultic/mortuary context (close to masseboth or open-air shrines). Five aspects are important to note:

Northern Negev sites are characterized by a low amount of non ad hoc tools. The high amount of ad hoc tool may be the result of a high availability of raw materials, which enabled a high frequency of tool production and discard after a short-term use. A complementary factor may be traditional, knappers preferring to produce new tools, instead of re-sharpening used ones, most tools being produced in each household.

1. the high variability of shapes and the fact that they were mostly found broken suggest re-sharpening, tools being discarded only after a prolonged use. 2. tabular scrapers were used both as domestic tools, with a steady apparition of ca. 2% of tools at most sites, and as symbolic items. Several stylistic aspects, such as perforation of some items in the Golan sites or smoothing of the dorsal cortex (Tel Sheva), may denote regional preferences, or different production centres. 3. the similarity of raw materials and manufacture modes suggest a common source for their production, probably in the Negev highlands. The presence of a workshop of secondary production at Ghazzeh A site (only tabular blanks were found at this site), implies an organized mode of manufacture and probably distribution of tabular scrapers, in a network covering most of the research area.

The relative amount of non ad hoc tools at Kh. Aliya and Abu Hamid is similar to that of Northern Negev sites (a 1/3 of tools), suggesting similar approaches to tool production. Contrasting are assemblages from northern sites, such as Kaukab, Tel Turmus and Golan sites 20 and 21, almost equally divided between non and ad hoc tools. The scarcity of raw materials in these areas influenced probably the frequency of tool production, re-sharpening and discard, tools being abandoned only after extensive use. This supposition is supported by the fact that many of the Tel Turmus tools were found broken, and the amount of double tools or re-sharpened tools is higher than at other sites. The high amount of non ad hoc tools at Golan sites 6 and 22 may denote the presence of two workshops, specialized in the production of perforated discs and sickle-blades.

106

CHAPTER VIII – DESCRIPTION OF RESULTS

was not concentrated in a workshop, nor did traditional factors impose a specific shape, but rather each household produced or achieved its own tools. The presence of backed blades on semi-translucent flint and discarded on various production stages (sickle-blade blanks) implies a local production of sickle-blades.

4. the presence of tabular scrapers in most habitation sites, regardless of their nature (in Safadi as well as Ramot 3) hint at their low economic value, these items being available for anyone. 5. their possible double nature, occasionally being found in cultic/mortuary context, as at Shiqmim and Eilat cemeteries and at Peqi’in burial cave.

Sickle-blades of northern assemblages (Tel Turmus and the Golan sites) are similar to those of the Negev: bitruncated, backed and with a fine denticulated working edge. Their morphological variability is higher: the width varies between 13 and 20 mm, not all sickle-blades are backed and denticulation was noted on only half of the items. Moreover, northern sickle-blades were made on several raw materials and not as in the Negev, where the majority of sickle-blades are made on banded flint.

Sickle-blades Sickle-blades were found in all but desert assemblages, their relative amount and shape varying in each of the defined geographic zones. Their absence in the arid zones may be related to their supposed function, as parts of sickles used for harvesting cereals, agriculture not being performed by the inhabitants of the analyzed sites of the desert (Southern Jordan and southeast Sinai).

No evidence for the production of sickle-blades in workshops was so far found, as occurred in the Negev; the high variability of raw materials and shapes of northern sickle-blades suggest that they were produced either by each household, or few, occasional workshops (a high concentration of sickle-blades at Golan site 6 may reflect the presence of a workshop at that site).

Sickle-blades of the Negev have a standardized form: made on pebble-banded flint blades, they are backed, double truncated and a fine denticulation along the working edge. The straight gloss along this edge suggests that the sickle was straight as well. The standardization of these items is best reflected in the low variability of their width: between 11 and 13 mm at sites of Northern and western Negev. The uniformity of the morphological characteristics of sickle-blades is related to: 1. 2.

Another difference between northern and Negev sickleblades is the delineation of gloss along the working edge: apparently, a Northern sickle had a curved working edge (evidenced by items with a curved back and an oblique gloss), while a southern one was straight (there are no sickle-blades with a curved back and the delineation of gloss is usually straight in Negev sickle-blades).

sickle-blades were produced in workshops that served an area beyond the borders of the site. cultural/traditional factors imposed the use of banded flint for the production of sickle-blades, chosen among other flints of the local wadis. The presence of single banded flint blade cores at most of the sites of this area (excluding of course sites were workshops were located) suggest first that sickle-blades production was not monopolized in the hands of a few specialists, and the tradition of their production (from banded flint) was well-defined.

The main difference between northern and western Negev sites is the relative amount of sickle-blades: in Northern Negev sickle-blades do not exceed 14% of all tools, while in western Negev there are twice as much sickle-blades. Whereas this difference may be interpreted in economic terms, agriculture playing a more important role in the subsistence economy of western Negev inhabitants, the possibility that it is the result of different discard patterns (Northern Negev inhabitants left the sites with most their tools, or hide them in “secret” places) cannot be ignored.

The sickle-blades of Kaukab, a northern site, are different from those of surrounding assemblages, resembling Negev ones: their mean width is 13.6 mm (narrower than typical northern sickle-blades), they all have a straight back and double truncations and denticulated working edges occur frequently. There is also a deliberate choice of a limited number of raw materials for their production: among the 9 raw materials types exploited, more than 60% of sickle-blades are on light brown flint, while the others are on grey or cream flint. The limited extension of the excavated area and the singularity of this phenomenon in Northern assemblages cuts down the potential for a coherent interpretation of the Kaukab assemblage and its apparent similarity with Negev assemblages. The possibility that a proximity to high quality raw materials enabled a standardized production of sickle-blades at Kaukab cannot be excluded.

The banded flint sickle-blades of Kh. Aliya were imported, probably from one of the workshops of the Negev, which implies a contemporaneity between these sites. The remaining sickle-blades, which form the majority, have a varied morphology, of blank type (blades or bladelets), retouch and size, the width varying between 10 and 13 mm. The main raw material used is the semi-translucent flint, common on most other tools. It might be concluded that the production of sickle-blades

The main characteristic of Nahal Komem sickle-blades is the appearance of two distinct reduction sequences: one resulting in typical Negev sickle-blades and the other producing “cannanean” sickle-blades, among them a few were backed (see above). A possible hallmark for later Chalcolithic assemblages could be a high amount of sickle-blades among tools (they represent a quarter of tools at this site) and the presence in an homogeneous assemblage of these two distinct sickle-blade types.

107

CHAPTER VIII – DESCRIPTION OF RESULTS

Picks are virtually absent in most of the research area, excluding a few sporadic appearances at Golan sites 12 and 22 and Ghazzeh sites A, B and O. Picks are absent from desert assemblages, excluding Timna 39, where they represent 2% of all tools. Their presence at this site may be related to a metallurgical activity performed at the site, picks being apparently used as heavy hammers.

Microliths There are several characteristics unifying microliths of the research area: they are on semi-translucent flint, most are micro end-scrapers and they were produced in workshops (Ghazzeh site A, and possibly Tel Sheva, Abu Matar and Mitcham C). The main dissimilarity concerns their frequencies, which vary by geographic location: 12 and 30% (Western Negev), 1 and 2% (Northern Negev, Golan and most northern assemblages) and around 6% of tools at Tel Turmus, Kh. Aliya and Nahal Komem. The high amount of microliths in Southern Jordan sites is due to the presence of geometric microliths, absent at all other sites and characteristic of these assemblages. There are a few exceptions: at Mitcham C and Tel-Sheva, microliths are 7% and 4% of all tools, respectively. At both sites a high amount of bladelet cores was found; microliths may have been produced and exported. Another exception is Kaukab in the north, where microliths are 7% of tools. One explanation is linked to the fact that blade and bladelet tools represent more than a 1/3 of tools, reflecting a local tradition of using blades/bladelets as main tool blanks. The high quality of the local flint and the high amount of bladelet cores sustain the suggestion of an intensive, local production of bladelets.

Picks are more common in Northern Negev assemblages, and especially at Tel Sheva (4% of all tools, compared to none or 1% of tools at the remaining sites of the area). A possible relation may connect picks with the copper metallurgy, being involved in this industry as heavy-duty tools, probably as hammers. Perforated discs were found almost exclusively in northern assemblages, while stars were found only in the Golan sites. Outside this area, two discs were found at Safadi and one at Ghazzeh A site in Western Negev, brought to these sites as end-products. Important to note is the fact that so far no workshop for the production of discs was found, though a northern location is susceptible. Another interesting aspect is the absence of discs at Kaukab, a site located in the north. Description of lithic industries by geographic regions

Bifacials

The following paragraphs are devoted to a description of lithic industries in each geographic zone, according to main stages of chaîne opératoire for tool production.

Several sub-types of bifacials were recognized: axes, adzes, chisels, picks and perforated discs (including stars). Their relative amount and mode of manufacture vary from each geographic unit to the other. Bifacials are absent from Southern Jordan assemblages; few were found in southeast Sinai assemblages, mainly adzes of a low quality. Notable is the high amount of bifacials at Timna 39 (9% of all tools), possibly the result of their use in metallurgical activities.

Golan sites Chalcolithic knappers of this area exploited several flint sources, located at various distances, on some occasions exceeding 25 km from habitation sites, as evidenced by the high number of flint types used (ca. 6). The low quality of local flint and its low availability was probably the main factor that imposed a search for various sources for flint. Observations on waste products indicate that more than half of these raw materials were on-site processed. A deliberate choice for a particular raw material was observed on sickle-blades, microliths and bifacials (including discs and stars), the remaining tools being made on practically any available raw material.

A general trend is for more axes and adzes in the Negev assemblages, while chisels and adzes dominate the bifacials group of Northern assemblages. This typological difference does not necessarily reflect different activities, but rather stylistic/traditional aspects. Another difference is in their mode of manufacture: bifacials of the Negev were mostly shaped from two or three striking platforms, (see Chapter III), the working edge being often shaped by a transversal blow, while Northern bifacials have usually two striking platforms and the working edge was shaped by several longitudinal removal (cf. Tel Turmus).

Several core types were found, dominated by single striking platform flake cores; a further use of cores, by renewing striking platforms and adding new ones caused the exhaustion of most cores. This behaviour may be related to a scarcity of raw materials in the vicinity of sites and to their low quality. Occasionally, pebbles were modified into choppers; the possibility that they were used both as cores and as heavy-duty tools cannot be excluded. Blade and bladelet cores are rare, the industry is clearly oriented towards the production of flakes, mainly with wide flat butts. This orientation is also reflected in the ratios between flakes and blades tools and blanks, flake blanks and tools on flakes being ca. three times more common than blades or blade tools.

Another difference is their general shape: bifacials of the north are elongated and narrower, while those of the Negev seem more “robust”, being shorter and wider. Apparently, bifacials were produced in workshops, both in the north (Abu Hamid and possibly Golan site 6) and in the south (possibly at Ghazzeh site A and Safadi, where a high amount of bifacials was collected – 16% of tools, compared to 2 – 6% at other sites of the region).

108

CHAPTER VIII – DESCRIPTION OF RESULTS

The main differences between both sites regard their tool assemblages, both from a typological and technological point of view. Outstanding is the absence of perforated discs at Kaukab, a tool-type that constantly appears in all northern assemblages of this period. Other quantitative differences are the high amount of sickle-blades, borers and microliths at Kaukab, while Tel Turmus assemblage is dominated by tools on flakes. The high amount of retouched blades at Tel Turmus may be typological: given the low standardization of sickle-blades, they were defined as such only by the appearance of gloss along their working edge. However, similar morphological and technological attributes merge these two groups into a single one. Thus, sickle-blades appear in similar, high percentages at both sites. The main difference between the two groups of sickle-blades concerns their degree of standardization: sickle-blades of Kaukab were mostly produced from a few raw materials and they are narrow, almost all bi-truncated, backed and with a fine denticulated working edge, while those of Tel Turmus are wider and other morphological variations are frequent.

The tool assemblages are dominated by ad hoc tools, produced at each household. Bifacials, mostly chisels and axes, were apparently produced in a few centres. Sickleblades are rectangular, backed, truncated and with a fine denticulated working edge. Produced on 3 to 5 types of raw materials, they are wider than southern sickle-blades and have a higher morphological variability. A few items with a curved back may represent end-parts of a curved sickle. A concentration of sickle-blades at Golan site 6 hints at the presence of a workshop at the site. Perforated tabular scrapers and stars are restricted to Golan sites. The lithic assemblages reflect an economy based on the household as the main subsistence unit, most tools being produced by need and not in a systematic, organized way. A possible workshop of sickle-blades, discs and bifacials was at Golan site 6, as indicated by their concentration, but not by any archaeological evidence. If so, the workshop probably served a local community, uniting the Golan area in a large economic (and cultural) unit. Other Northern assemblages

At both sites flint artefacts were abandoned at the location of their production or use, since no dumping zones were located. The inhabitants of both sites carried away usable tools when sites were abandoned ( most tools were found either heavily abraded or broken). A major difference was noted in the amount of rejuvenated tools, common at Tel Turmus, but absent at Kaukab.

The assemblages of Kaukab and Tel Turmus were analysed in more details, therefore additional information is available. The main visible difference between them and Golan assemblages is the presence of stars and drilled tabular scrapers in the later group. A finer comparison between Kaukab and Tel Turmus reveals further differences, some explainable in terms of raw materials qualities and exploitation mode, others apparently having a traditional/cultural nature.

Other Chalcolithic sites of Northern Israel and the Northern parts of the Jordan valley may represent occupations of human groups with similar cultural associations to those of the sites described above. Sites as Neve Ur, Beith-Shean, Beit Netofa, Abu Hamid, Pella, Tell Shuna North and Tel Ali may belong to a similar culture group to the Golan sites (sensu lato), according to flint typological similarities; however, as it was shown in the previous paragraphs, a detailed examination of lithic artefacts may reveal differences interpretable in cultural/traditional variations and not necessarily to raw material constraints, economic factors or different knappers and their skills. Thus, in general terms, it may be presumed that the northern part of the Southern Levant has several unique characteristics (mentioned above) that can separate it from other regions of the research area. Within it, fragmented regional archaeological cultures probably co-existed, as noted when the assemblages of Tel Turmus were compared to those of Kaukab in particular and those of the Golan sites in general.

Several sources of raw materials were used at both sites, acquired either by direct quarrying or through an exchange network (tabular and semi-translucent flints). However, the main difference is the quality of exploited raw materials, those of Kaukab being more suitable for tool production. Moreover, a more careful selection of raw materials for the production of tool-types was observed at Kaukab (sickle-blades), whereas at Tel Turmus most tool-types were produced on any available flint. Semi-translucent flint was used at both sites for the production of microliths. Technological observations reflect a similarity in the knapping traditions at both sites: cores were exhausted, being used to produce a variety of blanks, apparently core changing its aim according to reduction stages. Direct percussions with a hard hammer were used alternately to indirect blows with a soft hammer, according to desired blank and limitations imposed by core’s shape and size. Faceting of striking platforms and rejuvenation of cores by core trimming elements were common at both sites. The presence of pièces esquilles at Tel Turmus solely suggest a different technique adopted by knappers of this site, involving probably the use of an anvil. Even though flakes dominate the waste groups of both sites, blades were the preferred support for the production of tools at Kaukab, and to a lesser extent at Tel Turmus.

The site of Peqi’in is a unique phenomenon in the area; the presence of items of material culture characteristic of both southern and northern regions excludes its sole use by northern people, rather reflecting the similar religious (and probably cultural) background of the Chalcolithic inhabitants of Southern Levant.

109

CHAPTER VIII – DESCRIPTION OF RESULTS

The main characteristic of the tool assemblage is the dominance of sickle-blades and retouched blades (half of all tools) and the high amount of tabular scrapers. The presence of typical Northern Negev sickle-blades (on banded flint, bi-truncated, backed and with a denticulated working edge) along with typical EB sickle-blades (on canaanean flint blades with a trapezoidal cross-section) is the most striking feature of this assemblage. Bifacials, virtually absent in the EB, appear as well. The presence of sickle-blades on banded flint suggests that Nahal Komem hosted the remains of a population from the area of the Northern Negev that occupied the site at a later period than the Be’er-Sheva basin or the Western Negev. The high amount of sickle-blades may be the result of a shift in the manufacture strategy, less ad hoc tools being produced, used and discarded.

The subsistence economy was similar the Golan sites; tools were made in each household, typological and quantitative similarities hinting for similar activities performed at each site. Three workshops, identified at Abu Hamid, produced and re-sharpened mostly bifacials, but other tools as well. The presence of used tools in the same loci as the remains of the workshops suggest that the knappers were part-time specialists, their other activities including typical aspects of the subsistence economy of a Chalcolithic household. Central Israel assemblages The central part of Israel is poorly represented, a few sites being excavated and a single lithic assemblage was published – Khirbeth Aliya East. The methodological approach adopted during the analysis of this material limited its integration within a synthesis, and only a few general aspects could be retrieved from the report.

Northern Negev assemblages Northern Negev sites form a distinct cultural entity, the “Ghassul – Be’er-Sheva”. Among them, several types are discerned: with underground structures and above-ground architecture, installations, pits and open courtyards, extending over large areas, and a rich and diverse material culture, some with ateliers of various products (sickle-blades, ivory, copper). Superposition of structures, repairs of buildings or abandonment of old ones are evidence of repeated occupation stages. Main sites are Safadi, Abu Matar, Tel Sheva, Shiqmim and Nevatim. Along these, there is a group of smaller sites, lacking architecture or underground structures, and no internal stratigraphy. Main features are pits and activity areas between them. Sites of this group are Meitar, Ramot 3, Ramot Nof and Mitcham C. While Ramot 3 and Ramot Nof are apparently of an earlier period, Mitcham C is a part of an extensive occupation which extends below large parts of southern modern Be’er-Sheva.

A few raw materials were used, among them semitranslucent flint is the dominant one, its choice being related to its availability and not necessarily to traditional constraints. Its extensive use influenced the nature of the industry, blades and bladelets being common tool blanks (1/3 of tools are made on blades/bladelets). The extensive use of semi-translucent flint caused a knapping method that produced a high amount of un-usable waste; a long preparation of cores being probably necessary before suitable blanks could be produced; cores were discarded only after prolonged use, mostly being exhausted. Most tools were on-site produced; the presence of sickleblades on banded flint, typical of Negev assemblages, on one hand, and the absence of perforated discs on the other, suggest a southern orientation of relations. The presence of axes in high frequencies, comparable to Northern Negev assemblages, indicates the same. These cultural ties were weaker than among Negev sites, as evidenced by diversity of sickle-blades, some procured by an exchange network from Negev sites, while others being made from the locally available raw material.

The comparison between assemblages of both groups of sites revealed their similarities, in typo – technological terms. Knappers exploited local raw materials, changing their knapping methods according to flint’s quality. Semitranslucent flint was used for microliths production; banded flint is common for sickle-blades. Tabular scrapers were not on-site produced. It may be concluded that a similar strategy for raw materials procurement was adopted by knappers of all these sites.

Little can be said regarding socio-economic organization of inhabitants of this area, except that their subsistence strategy was similar to other Chalcolithic settlements, either of the north or of the south; workshops were rare, if existed at all. Few items were procured through a goods exchange network, and all others were manufactured in each household, from the locally available raw material.

The flake is the principal tool blank; Northern Negev flint industries are dominated by flakes, obtained from pebble flint unipolar cores, by a technique involving a direct blow with a hard hammer, resulting in a wide, flat butt and prominent bulb of percussion. Occasionally, such as at Tel Sheva, Meitar or Ramot 3, discoidal cores were used, especially when low quality flint, mixed with limestone, was exploited. In these cases, a faceted butt appears more frequently. The exploitation degree of cores was low (mean value of ratio tools/cores = 2) and the amount of cortex on the débitage surface of cores.

Nahal Komem assemblage Apparently, the lithic assemblage of this site is transitional between the Chalcolithic and the EB periods. Main innovations are the extensive use of “canaanean” flint as a main raw material for tool production and the use of blades as a main support for the production of tools, even though flakes dominate the waste products. Cores were often rejuvenated and highly exploited.

110

CHAPTER VIII – DESCRIPTION OF RESULTS

excluded. Scrapers of this region were termed “Be’erSheva” scraper, given their characteristics: on primary flakes, with a prominent bulb of percussion, a cortical or flat butt and semi-abrupt retouch along its edges. Important to note is the fact that this type was also found at all sites in the Western Negev area as well.

A typical knapping sequence would involve the acquirement of pebbles from the nearby wadi, its decortication within the site, suitable removals from this stage being used for the production of scrapers, or smaller ones for borers. The small amount of scars on the dorsal face of flakes indicates a reduced number of removal sequences, probably not more than three. Additional support for this suggestion is gained from observations on cores, most retaining cortex on their circumference and with a few striking platforms. Moreover, only a few cores were renewed (commonly each third core).

Microliths are few in Northern Negev sites, but Mitcham C (7% of tools), where this amount may be related to the high amount of bladelet cores found, all probably being remains of a microliths workshop. Interesting to note is the fact that at Tel Sheva, where a similar concentration of bladelet cores was found, there are a few microliths (4% of tools). An explanation is that at Tel Sheva most products were exported, while at Mitcham C microliths were locally used.

Blades were knapped from banded flint pyramidal cores, with one unfaceted striking platform and retaining cortex on half of their circumference. Their knapping technique, used mostly as blanks for the production of sickle-blades, involved the use of a soft hammer, resulting in a pointed butt. Further, these blanks were shortened to a desired length (depending to a required length of sickle-blades to be incorporated within the sickle) by the re-invention of the microburin method (see above at Safadi).

Bifacials represent 2% to 6% of all tools; adzes dominate, followed by axes, a few chisels and occasional picks. Only two perforated discs were found, both at Safadi, brought as end-products, through an exchange network system. Apparently, picks are related to mining activities, being tools that could have been used to dig underground structures. Important to note is that picks were not found at sites defined as ephemeral (Ramot 3, Ramot Nof or Meitar), while very few to none were found in most other research areas, where such structures are absent as well.

Blade cores from a clear context were discovered at Safadi; another concentration of blade cores was found at Mitcham C, probably reflecting the presence of another workshop at this site as well. Blade cores on banded flint, in limited numbers, were found practically in each site of the area. Thus, aside production by specialist knappers in workshops, sickle-blades were also locally manufactured in some households. The fact that there is practically no difference between products of workshops and those of part-time specialists implies that all knappers had similar skills and knowledge, regardless their locality of activity (workshop or household). It is suggested that flint knapping did not require particular specialization, or, subsequently, all flint knappers were specialists.

Adzes, axes and chisels were modified by two, sometimes three striking platforms, their working edge being shaped by a transversal blow. It is not known whether this removal served to renew the working edge or not. Other modes of preparation of the working edge are small, longitudinal bladelet removals, or a few convergent flake removals. Only a few bifacials were polished, the criteria of which one would be polished depending probably on particular choices of the knapper.

Bladelets were exclusively produced from semitranslucent flint; bladelet cores were found in small concentrations at a few sites, as at Tel Sheva and Mitcham C. Important to note is the absence of bladelet cores at Safadi and their very limited number at Abu Matar. The same knapping scheme described above for sickle-blades is true regarding microliths: apparently, these were produced mainly, but not only, in specialized workshops. Bladelet cores have usually one striking platform and a débitage surface cleaned from cortex. Faceting probably increased the exploitation degree of these cores. Their small size and frequent renewal of striking platforms suggest these cores were exhausted.

Artefacts were mainly found in dumping zones: pits, outof-use bell-shaped silos or underground structures. Concentrations of particular types reflect remains of workshops: at Safadi, a small pit contained blade cores and hundreds of their products, and at Tel Sheva, most bladelet cores were found in one locus. An important aspect of these assemblages is the similarity between those originating from large sites, as Safadi and smaller ones, as Meitar, both in typo-technological and relative quantity aspects. Another important characteristic is the presence of specialized workshops, for microliths and sickle-blades, alongside the manufacture of the same products, though in smaller quantities, in households.

The tool assemblages are dominated by flake tools, such as retouched flakes, notches and denticulates. Scrapers comprise 15 to 34% of tools; they are mostly made on primary elements, with partial semi-abrupt retouch at their distal end, but may extend along one edge or both, sometimes covering the entire circumference. The possibility that the extent of retouch denotes resharpening of the tool, and not a distinct sub-type cannot be

Western and Northern Sinai assemblages Among sites of this area, several display impressive remains of lithic material culture. These, such as Ghazzeh A and M, are located in the Nahal Besor area, their most outstanding feature is flint workshops for sickle-blades, microliths, bifacials and apparently tabular scrapers.

111

CHAPTER VIII – DESCRIPTION OF RESULTS

Unfortunately, the selective collecting method of the lithic material, focusing on “indicative” tools, limited their potential for research.

the one collected at Safadi, where no workshop for the production of tabular scrapers was reported, since it probably did not exist.

Another area, intensively investigated, is the bank of Nahal Grar, where a cluster of sites was excavated and published under the name of Grar. Its lithic remains were extensively studied and published (Gilead 1995). Another cluster was noted in Northern Sinai, following a survey in the area. Among the several dozens of Chalcolithic findspots discovered, 4 were excavated and yielded a sufficiently large assemblage to be broadly analyzed.

Respecting technological aspects, the industries of western Negev sites are similar to those of the Be’erSheva basin sites: flakes were produced mostly from single-platform pebble cores and represent the preferred support for tool production. A possible difference is the intensity of cores exploitation: assuming a similar size of pebbles from Nahal Grar and Nahal Be’er-Sheva and its tributaries, knappers of Grar sites removed more flakes from each core than those of Be’er-Sheva sites, Grar flakes and cores being smaller than those of Be’er-Sheva. Thus, a possible traditional mark can be traced, involving the exploitation degree of cores. However, the possibility that Nahal Grar pebbles are smaller than those of Nahal Be’er-Sheva cannot be excluded, since for example banded flint, a common pebble of Nahal Be’er-Sheva, is lacking from Nahal Grar. Therefore, the size of cores and flakes might have been influenced by the original size of pebbles and not by the intensity of exploitation.

A common feature of these sites is their topographic location, all along wadi banks, in proximity to water sources. Common features are dwellings built from mudbricks, occasionally with stone foundations, installations, pits and open courtyards. Isolated, small scale workshops were found within habitation areas. The presence of perforated discs at Site A testifies for some kind of relations with northern groups, probably based on a goods exchange network.

An intensive production of blades/bladelets characterize the lithic assemblages of western Negev. Even though a few blade and bladelet cores were found at Grar, the high amount of tools on blades (especially when compared to Be’er-Sheva sites), testifies for a deliberate, intensive production of blade and bladelet blanks. The possibility that some of them were produced at Ghazzeh site A or to a lesser extent at B, cannot be excluded. A concentration of blade cores was located in one of the survey sites of north-eastern Sinai. The possibility that blanks produced there served a local community, including other habitation sites of the area, cannot be excluded.

Several characteristics of the flint industry could be drawn. Knappers exploited the local pebble flint to produce most tools – pebbles were collected from wadis, brought to the site, decorticated and exploited for the production mostly of flakes. Semi-translucent flint was exclusively used to produce microliths, hundreds of bladelet cores being found at Ghazzeh site A (around 800 cores!), but at Grar sites and Northern Sinai sites as well, and not in workshop contexts. Banded pebble flint was mainly used to manufacture sickle-blades, a large amount of blade cores being found at Ghazzeh sites A and B, while only a few at other sites.

Blades and bladelets are (at least partially) products of a different knapping technique than flakes: more blades than flakes have linear or pointed butts, suggesting the use of a soft hammer, blanks being knapped off mostly from unipolar cores. Their sizes follows the tendency for lighter blanks than those of the Be’er-Sheva sites – the mean blade sizes at Grar are 39 mm length and 17 mm width, opposed to 71 mm and 24 mm at the Be’er-Sheva sites. Thus, while the method and the technique of blade and bladelet production is identical at Be’er-Sheva and western Negev sites, the size of blanks differ. Another difference is the use pattern of blanks: while in western Negev they might have been used mostly locally, the high concentration of bladelet cores and the scarcity of bladelets at Tel Sheva for example, suggest that products of this workshop were exported to farther distances.

A unique aspect is the apparent presence of a workshop for the production of tabular scrapers at Ghazzeh site A. Apparently, large tabular flakes were brought to the site and further modified into tabular scrapers. No cores on tabular flint were reported from any of these sites. The presence of a workshop for a secondary production stage of tabular scrapers at Ghazzeh site A is a unique phenomenon within the research area. However, its importance must not be overestimated, given several reservations presented below. Even though Roshwalb reports “blanks in all stages of manufacture, from totally unretouched blanks to finished tools” (Roshwalb 1981:39), she also adds that “…lacking is any débitage debris – chips, chunks, decorticage flakes, simple flakes…” on tabular flint (Roshwalb 1981:39), while in other places she writes that débitage “…tabular flint will not be discussed…” (Roshwalb 1981:58).

One of the obvious characteristics of Western Negev assemblages is the high amount of microliths (10 to 30% of tools), mostly micrograttoirs made on semi-translucent flint, apparently locally produced. Since no use-wear studies were performed on microliths of this period, their use is unknown; therefore, their high amount remains

Thus, the possibility that the blanks in fact represent broken tabular scrapers cannot be excluded, given the fact that apparently no tabular flint waste products of any type were found. Moreover, the relatively high amount of tabular flint found at site A is not markedly different from

112

CHAPTER VIII – DESCRIPTION OF RESULTS

Therefore, a few characteristics of these lithic industries could be articulated. Most industries exploited local raw materials; tabular flint appears to be the only raw material used from farther sources, being imported from several sources as end-products, (tabular scrapers), as testified by the different raw materials and production methods. An apparent deliberate choice for specific raw materials was observed in eastern Sinai assemblages, where delicate tools were made on black raw material, of a high quality, while larger and cruder tools were made on brown flint. Consequently, knappers of Southern Jordan sites chose semi-translucent flint to produce microliths, while other tools were made on the local tabular flint.

unexplainable at this stage of research. The proximity to a workshop where hundreds of bladelet cores were found (Ghazzeh A) may have had an influence on the amount of microliths found at the other sites, new supply of bladelets or microliths being locally reachable. Their production scheme is similar to that reconstructed for other Chalcolithic assemblages: mainly semi-translucent, single-platform bladelet cores were highly exploited, the knapping process being performed with a soft hammer and a possible indirect blow. Another characteristic is the high amount of sickle-blades at Grar and some north-eastern Sinai sites (14 to 33% of tools). The relative amount of sickle-blades depends on classification methods or strategy of material collection (as in some old excavations, such as at Wadi Ghazzeh sites, where a few ad hoc tools were collected), or frequency of tool production and rejuvenation (i.e. a different approach to scraper production, renewal of working edges at Grar, against the production of new and discard of worn out ones at Be’er-Sheva sites).

The flint industries are oriented towards the production of large flakes, except those of Wadi Sa’al, where the dominant blank is a small flake, and some Southern Jordan sites, where bladelets were used as blanks for microliths, a phenomenon restricted to sites of this area. Cores are mostly unipolar; the presence of a few items resembling Levallois points/flakes suggest the use of discoidal cores as well.

Thus, differences in the relative amount of tool-types cannot serve as parameters for defining cultural entities. Moreover, sickle-blades of Western and Northern Negev assemblages share morphological characteristics (denticulated working edge, back and double truncation) and size (11-13 mm width) and raw materials (banded flint). A major difference is the absence of picks at Grar and north-eastern Sinai sites. Picks were reported from Wadi Ghazzeh A, B, D and O sites. Other bifacials, such as axes, adzes and chisels resemble Northern Negev ones.

Few tools can be said to be characteristic of this zone. Common tool-types are retouched flakes, notches and denticulated. Tabular scrapers were found in relatively high amounts at a few sites. Bifacials were rarely used, except at Timna 39, where an apparent metallurgical activity was performed with bifacials of all sub-types, including picks. Retouched blades with a curved back may represent a characteristic tool-type of this region. Another possible candidate is a tool with a wide working edge, fan-shaped by semi-abrupt retouch and two notches along its edges, thinning the mesial and proximal parts, forming a handle-like part (cf. Wadi Watir, Plate 33:2,3).

The main differences between Northern and Western Negev lithic assemblages are the extensive use of blades as a preferred blank for tool production and a high amount of sickle-blades and microliths in Western Negev sites and the exploitation degree of cores, more removals being knapped from cores at Western Negev sites (such as at Grar), despite a high availability of raw materials. At this stage, it is unclear if these characteristics represent an earlier entity (the Besorian), particular choices of the local knappers (as exploitation degree of cores, frequency of tool production and their rejuvenation) or specific economic conditions (proximity to workshops).

Outstanding is the lack of contact between the desert zone and northern areas, as reflected in the lithic evidence. The only apparent relation is the tabular scraper, used both as a tool for domestic activities and in cultic contexts. The uniformity of lithic production and its restricted products is tentatively related to a high mobility of human groups inhabiting this zone, sharing cultural/religious ties with populations of other zones, but otherwise having a distinct cultural identity.

Desert assemblages (east Sinai, Arabah valley, Southern and eastern Jordan) Despite the fact that this area is the largest, covering the Arabah valley, Southern and Eastern Jordan and Eastern Sinai, little can be said regarding its lithic industries. Previous research defined two industries: Timnian and Eilatian. The present work failed to isolate their characteristics, given the fact that no distinct features were isolated in the analysed material. Moreover, the data available for study was incomplete in most cases, sites being partially excavated, data originating from selective surface collection or analysed by different methodologies (such as the Southern Jordan assemblages).

113

CHAPTER IX – CONCLUSIONS

The above presented picture may be summarized as follows:

CHAPTER IX CONCLUSIONS The work presented above aimed at four main goals: the evaluation of the potential of lithic studies for the reconstruction of socio-economic aspects of Chalcolithic societies, their description and the reconstruction of the Chalcolithic cultural mosaic of Southern Levant, each with their lithic cultural hallmarks (fossiles directeurs). Before going into details of each of the work’s goals, some general remarks, concerning the research on Chalcolithic flint industries and its results, are presented.

1. a simple operational scheme, involving the use of various, mostly local raw materials, in a flake industry, aimed at producing simple flake tools. 2. a volumetric reduction of cores reflected by discoidal cores and their products. 3. a well-defined blade and bladelet industry, pyramidal cores used for the production of sickleblades; bladelets are modified mostly into micrograttoirs. The microburin technique is applied to control the length of sickle-blades. A deliberate choice of raw materials was recognized in both cases, banded pebble flint for sickle-blade blanks and semi-translucent flint for microliths. 4. production of various types of bifacials, both on flint, limestone of basalt. 5. unique techniques and tool-types appear: perforation of tools and “invention” of new tooltypes – discs and stars, characteristic of Northern assemblages.

One of the most striking characteristics of the organization of flint production is its high flexibility and process-adaptation to constraints (as raw materials) and the variety of knapping methods, spanning from simple ones, aimed at producing flake tools, through the volumetric method of discoidal cores and the blade/bladelet production to sophisticated methods that created bifacials and perforated discs. Thus, the lithic industry is described as innovative, highly adapted to needs, knappers applying specific knowledge whenever needed, as reflected in the standardized production of particular tool-types (such as discs, bifacials or sickle-blades), otherwise changing methods according to raw materials’ quality.

The analysis of lithic assemblages from the research area allowed the recognition of five main culture-zones, with possible variation within each group: 1. A Southern desert group: Arabah, eastern Sinai, Southern and eastern Jordan. Possible variations were noted by Henry (1995) at some Southern Jordan sites. 2. A Northern Negev group, referred to as the “Be’er-Sheva – Ghassul”, limited to sites in the Be’er-Sheva valley. 3. Sites of western Negev (Besor and Grar areas, and north-eastern Sinai), some earlier but sharing most of the techno-typological characteristics. 4. The central part of Israel - a poorly represented zone, only partially studied. 5. A Northern group, extending in the Jordan valley and the Golan heights. Possible variations are in the Hula valley and Upper Galilee.

The production of tools is oriented towards the fulfilment of functional needs, esthetical requirements being observed on tools found in burial contexts (such as bifacials and tabular scrapers, with a particular morphology, see above). Traditional marks are more pronounced in Negev assemblages, where sickle-blades are mostly produced on banded flint, showing a high standardization, both of products originating from workshops and those from regular households. This phenomenon reflects a common operational mode of knappers, all sharing a same connaissance of tool production, using the same savoirfaire, regardless of locality of production. Another characteristic of the lithic industry is the strategy of raw material exploitation, three main approaches being identified: knappers chose optimal raw materials for the production of particular tools, such as semi-translucent flint to produce microliths, tabular flint for the manufacture of tabular scrapers and to some extent bifacials. Another strategy carries traditional marks, such as a deliberate choice of banded flint for the production of sickle-blades, while the third strategy responds to functional needs; locally available flint was used for the production of most tools. Moreover, within the third strategy other raw materials can be included, as limestone for the production of scrapers, bifacials, awls, etc. at Grar (Gilead and Fabian 1995), or bifacials on basalt, as evidenced at Tel Turmus (Marder, pers. comm.).

Each zone was economically independent; a basic unit was the household, supported by regional workshops, their products being exchanged with other regions: perforated discs were found at some Negev sites, while banded flint sickle-blades were found in the central zone. This network advances a contemporaneity of Negev and northern sites (supported by C14 dates). Sites in the desert areas were isolated, no typical products of other regions were found; the presence of tabular scrapers, both in domestic and cultic contexts, denotes a common symbolic background with other sites of Southern Levant. In the following paragraphs the main conclusions of this work are presented according to the aims presented in Chapter II.

114

CHAPTER IX – CONCLUSIONS

The analysis of technological aspects of waste products and tool-types enabled the reconstruction of the operational scheme along both the techno-social and techno-psychological axis. Thus, various knapping techniques and methods were identified and reduction sequences reconstructed. Among the observed variables, the size of waste products, taking into consideration raw materials characteristics, served as a valuable index for evaluating and comparing between the exploitation degree of cores, efficiency and quality of production and knapping methods of the various industries. However, a more detailed technological research, based on experimentation and refitting, would reveal further aspects, such as isolation of technological aspects influenced by decisions taken on personal levels, by each knapper, and a more detailed reconstruction of the operational scheme of each product.

Evaluation of the lithic analysis’ potential for the reconstruction of Chalcolithic societies One of the most challenging aspects of this work was to synthesize material originating from sites excavated by different methods and analyzed under various frameworks. For example, the lithic assemblages of the Besor area (Wadi Ghazzeh sites), were mostly excavated at the beginning of the thirties in the last century and analyzed fifty years later. Even though not specified, the collection method was most probably oriented towards gathering “indicative” material, frequently ad hoc tools and waste products being neglected. Thus, quantitative computations and comparisons are extremely limited. The most striking example is that of Teleilat Ghassul, a type-site of the Chalcolithic period, one of the most famous ones of this region, used as a cultural comparison unit when discussing cultural developments in the Southern Levant of the 4th – 5th millennia B.C. However, its lithic assemblage was never comprehensively published, even though a doctoral thesis described some of its aspects (Lee 1973) and reports of the thirties’ excavations were published (Malllon, Koeppel and Neuville 1934, Koeppel 1938).

The quantitative and technological research of the tool assemblages enabled the individualization of regional cultures, recognition of relationships between environment and tool-types and reconstruction of socioeconomic organization in the research area. Important to note is the evaluation of the re-sharpening degree of tools, confronted with the research on raw material abundance and availability and preservation state of tools when discarded (complete, versus broken/damaged).

Another difficulty arose when assemblages were analyzed using a different framework, such as the Golan ones. Thus, despite the large amount of material, various aspects are missing from its publication (Noy 1998).

The record of the archaeological context of tools revealed important aspects of tool-types, such as tabular scrapers, perforated tools and bifacials, found both in habitation sites, used in domestic activities concerned with the household economy, as well as in cultic contexts, near open-air shrines, masseboth and cemeteries, exposing another, symbolic aspect of these tool-types. The use of tabular scrapers in burial/cultic contexts both in desert environments, as well as in the Negev area and Northern Israel hints at a common cultural/belief background and their material manifestation, using similar symbols.

Finally, political factors restricted the accessibility to some collections, stored in neighbour states, or are still waiting for research by the archaeologists who excavated the sites and collected the material. Despite all these limitations presented above, the value of a large synthesis was (hopefully!) demonstrated in this work. However, there is a need to develop an acceptable, common methodological framework for analysis of flint artefacts of this period, in order to facilitate future research based on gathering information from various publications, or subsequently, to present, in as much detail as possible, the methodology applied when material was published and its result, the “adjustment” of results to others, obtained by applying other methods of research, being more facile.

The spatial analysis of lithic material location revealed two main discarding strategies: inhabitants of Negev sites disposed their refuse in dumping zones, pits of various size and shapes, probably out-of-use structures, available in the vicinity of their habitat. Remains from workshops were also dumped into regular pits, together with material from other activities. Thus, aspects as spatial organization of work, or activities related to gender differences could not be performed at these sites. The spatial distribution of flint in Northern assemblages presented a different picture: artefacts were abandoned in an apparent random mode, either in pits, fills or floors. Remains from workshops were found at the locality of their production. Northern sites are a better locus for performing a detailed spatial analysis of remains of material culture, when issues such as spatial organization of work are discussed.

The analysis of the lithic material using the chaîne opératoire concept proved to be efficient, each research stage contributing to achieve the goals of this work. Thus, the reconstruction of raw materials acquisition strategy, quarrying and exploitation revealed important aspects related to socio-economic and cultural organization during the discussed period. However, a detailed geologic research, aimed at identifying sources of the various raw materials and their properties is highly needed, in order to extend and increase the reliability of research on raw materials.

115

CHAPTER IX – CONCLUSIONS

Workshops were found in all but desert and centre areas. The lithic industries of the desert area are not complex, most tools being termed as ad hoc, produced by simple knapping methods. The high mobility of the human groups inhabiting these zones probably had an influence on the absence of workshops. The lack of comprehensive research in the central part of Israel is probably responsible for the fact that no workshops were found in this area. Workshops were located in the Negev: possibly a blade and bladelet one at Mitcham C, a bladelet one at Tel Sheva and a bifacial one at Safadi (all identified as small concentrations of these products), clear workshops at Safadi (sickle-blades) and at Ghazzeh A (sickle-blades and microliths). In Northern Israel, high concentrations of sickle-blades and bifacials may indicate a workshop at Golan site 6, while at Abu Hamid three workshops produced and rejuvenated different tool – types.

Chalcolithic socio-economic aspects reflected in lithic assemblages Several reduction sequences characterize lithic industries of this period: 1. a simple one, aiming at producing non-determined flakes, mostly from unipolar or amorphous cores, to be used as supports for ad hoc tools (retouched flakes, truncations, notches, borers and scrapers in some occasions). It appears in each of the culture zones. Its remains were found in households and the workshops of Abu Hamid. 2. a blade one, aimed at producing blanks for sickleblades, from banded flint pyramidal cores. It was found in the Negev area. Its waste and products were found in pits belonging to workshops (as at Safadi), but in households as well (e.g. Meitar, Tel Sheva). 3. a bladelet one; microliths, mostly micrograttoirs on semi-translucent flint were their main product. It was found in all but the desert areas (exceptions are a few sites in Southern Jordan). Bladelets and their cores were found in Negev sites as concentrations (hundreds were collected from Wadi Ghazzeh A), small clusters (Tel Sheva), or singular items (most other sites). 4. a sequence that produced bifacials, found in all areas, less in the desert. Their production and resharpening in workshops, besides their probable manufacture in domestic contexts, is evidenced at Abu Hamid and Wadi Ghazzeh site A. Another possible production centre was at Golan site 6. 5. a complex one, aimed at producing perforated tools (discs and stars). Since the process of their manufacture is unknown, (knapping off a large flake, thin, bifacial removal and hole drilling), it is unclear whether these items were produced only by specialists or not. So far, no remains from this sequence (except the end-products) were found. A northern location of its production, apparently by specialized knappers, is expected. Discs were found in the north and in the Negev, but not in the centre or desert sites. 6. a sequence devoted to tabular scrapers production, items found in all the study area. So far, the only site where stages of their production were apparently isolated is Ghazzeh site A. Tabular scrapers were apparently produced in workshops, probably located near raw material sources.

A simple calculation of the number of sickle-blades produced at Safadi returns the value of 700 (10 blanks from each core). Assuming an average number of 4 items were hafted in one sickle, the workshop of Safadi produced around 200 sickles. It may be assumed that it was active on a seasonal basis, serving a limited community, with a sphere of influence restricted to the radius of the site and its immediate neighbours. The pit of Safadi was found incidentally (it is a small depression from an ancient surface, with no relation to structures or other features). Moreover, the site, as well as others, was only partially excavated (10 to 50% at the most), which leaves little opportunity to find remains of this kind and subsequently to artificially increase their importance, whenever they are identified. So far, no clear locations for (Chalcolithic) tabular scrapers production or perforated discs were identified. Apparently, tabular scrapers were produced in the south, possibly near raw materials sources, in the Negev highlands, while perforated discs were made in the north. Discs are common in northern assemblages, (only 2 items were found at Safadi and few at Ghazzeh A); tabular scrapers appear in same percentages throughout the sites of Southern Levant. Both types were found in domestic and ritual contexts. However, their economic value should not be overestimated, since they were found in almost every household and at any site type (ephemeral, such as Ramot 3, and villages, such as Safadi). Another important aspect of workshops is their location, always found integrated within habitation sites that can be referred to as villages (such as Safadi or Abu Hamid). This may be related to the strategy of raw material procurement and exploitation: knappers collected most of their flint in the immediate vicinity of their habitation sites and modified it within the site. Tabular scrapers and perforated discs, both produced on raw materials not available in the vicinity of habitation sites, were probably produced in workshops located close to quarry localities.

To summarize, lithic production was performed both in households and workshops. Important to note is the fact that products of both location types are similar; the only change is quantitative: for example, while at Safadi a pit containing 69 banded flint blade cores and hundreds of blade fragments was found, only isolate cores were found at the neighbour site of Abu Matar, each in a different location. However, the blade cores from both sites are identical, being exploited by applying the same method. The same is true regarding the production of microliths.

116

CHAPTER IX – CONCLUSIONS

The analysis of the pottery assemblage from Abu Matar and Zoumeili led Commenge-Pellerin to conclude that the ceramic industry was oriented towards the manufacture of domestic products; the mode of manufacture was local and familial, without indication of specialist artisans or products of high social value (Commenge-Pellerin 1987:130). This conclusion is consistent to the ones presented in this work.

The presence of an exchange network of goods is testified at three levels, defined by the movement distance of items and their types: sickle-blades, microliths and possibly bifacials were traded at short distances, at radiuses of a few dozen kilometres; a longer distance exchange network, of a few hundred kilometres, testified by the presence of perforated discs, originating from the north, in Negev assemblages and probably tabular scrapers from the south in Northern assemblages. A possible third level, linking localities of a few thousand kilometres, traded with obsidian, was found at Teleilat Ghassul. The small amount of traded items reflect the low economic importance of this network, items being exchanged more as social symbols, rather than economic value. Therefore, it is suggested to view it not as an economic mechanism, but rather a social-cultural one.

An analysis of pottery assemblages from Northern Negev sites indicated that “…vessels were hardly carried or exchanged between sites … communities were settled … no market for pottery products existed…economic integration of the different parts of the Northern Negev was low or non-existent…the different sites represent autonomous villages which were not part of larger complex socio-economic systems…potters were not highly professionals…engaged in the craft on only a part time basis…” (Gilead and Goren 1995:211). When referring to a particular group of vessels, the “Cream Ware”, the same authors conclude that “… some were produced elsewhere…this exchange was not of a commercial, utilitarian nature…”, but rather they had ”…a social function…elements in regulating social relations…” (Gilead and Goren 1995:211).

Another important economic factor is the similarity of tool-types and their relative amounts between large sites, such as Safadi or Tel Sheva, and smaller ones, such as Ramot 3 and Meitar. The same is true regarding sites in the western Negev (Grar), compared to more seasonal ones of north-eastern Sinai. This similarity is explained in economic terms, the basic economic unit being the household, with similar functions and needs at ephemeral sites as well as at larger villages. Thus, larger villages may be seen as aggregations, clusters of economically independent units, gathered or dispersed according to resources in a given area. Furthermore, the similarity between products of workshops and those manufactured in households suggests that there were no special requirements from knappers working in workshops, and they were not particular specialists, their knowledge being shared by all knappers.

A different opinion was expressed by Bourke (2001), regarding pottery production at Teleilat Ghassul; an increase in the production of specialized ceramic types, such as large storage vessels (opposed to mass-produced hole-mouth and cooking pots and feasting vessels, i. e. cornets) indicates a developing social stratification. Other indicators of specialized economy are the changing manufacturing processes and expanding distribution networks over time and the presence of unusual ceramic forms, such as anthropomorphic and zoomorphic vessels at Gilat and En Gedi. A similar dichotomy was observed in the production of basalt vessels, utilitarian ones being on local material, apparently produced on-site, while basalt fine-stone vessels were made at specialized sites in North Jordanian upland, and brought in as end products. Maceheads are viewed as markers of social status, evidence of social stratification and elite activity. The presence of seals, tokens and potmarks are interpreted as a “…first tentative to move towards heightened levels of bureaucratic recording practise…” (Bourke 2001:145).

As reflected from the analysis of the lithic material culture, there is insufficient evidence to indicate marked social differences between members of any community, or between inhabitants of smaller or larger sites. Apparently, the lithic industry was oriented towards obtaining optimal solutions with minimum effort or investment; thus, esthetical requirements are limited to few items (such as cleaning off cortex from perforated tools) and traditional marks being restricted (as banded flint used for the production of sickle-blades in Negev area). Mostly local raw materials were exploited, their quality imposing shifts in knapping methods and techniques (as observed at Meitar, Kaukab or Safadi).

Despite a widespread material exchange and a regional trade in basalt, alabaster and hematite, their economic importance was low, most material originating from the vicinity of sites and locally processed. The economy was based on distinctive subsistence strategies, the storage of agricultural surplus (evidenced at Teleilat Ghassul by the presence of huge pithoi) is a sign of developing elites. The Chalcolithic society is viewed “… as one in which the “traditional” symbolic/religious ruling elite gradually came into competition with an emerging elite based on the control of agricultural surpluses…” (Bourke 2001:152).

When reconstructing socio-economic aspects of the Golan site communities, Epstein (1998) uses terms as “…egalitarian …social structure…of family or clan relationship…or joint exploitation of subsistence by extended family groups…trade and exchange contacts were limited…There was…maximum exploitation of the raw materials available in the immediate vicinity” (Epstein 1998:333). These conclusions correspond to those regarding this aspect and presented in this work.

117

CHAPTER IX – CONCLUSIONS

A different opinion was proposed by Lovell (2001), following her analysis of the Teleilat Ghassul ceramic material, where in the later phase a specialized, on-site pottery production and the import of items from the west was noted. Limited contacts existed with the Negev as well, though a general isolation of the site was noted. Moreover, “…levels of development and contact between villages was inconsistent at the best…” (Lovell 2001:51).

4. the western Negev sites, an apparent (possibly earlier) sub-facie of the Northern Negev area, in the Besor and Grar area and north-eastern Sinai. 5. the desert zone, extending along the Arabah valley to eastern Sinai and Southern and eastern Jordan. Each zones’ characteristic lithic industries were described above. Apart from defining features of each culture zone, there are some shared attributes, suggesting a common cultural background of Southern Levant’s inhabitants. This is best reflected in manifestations of religious/cultic activities, mainly the use of tabular scrapers, and to a lesser extent of bifacials as symbolic tools, found in the north as well as in the south in burials contexts (Mezad Aluf, Peqi’in cave, Eilat cemetery, but not in the coastal burial caves, Nahal Qana cave or Kissufim cemetery).

While Gilead and Goren’s conclusions accord well with conclusions of this work, Bourke’s results are markedly different and require an explanation. The apparent specialized economy proposed by Bourke is based on the recognition of “expanding distribution networks”, though the same author further concludes that “… there is little reason to posit systematic, long-distance, reciprocal trade…a low intensity … exchange is more likely, but such cannot be … important in the local economy…” (Bourke 2001:150). Moreover, the dichotomy of ceramic and basalt production is similar to the lithic production (tabular scrapers and sickle-blades, versus borers or scrapers). The lithic and ceramic evidence, as reflected by the analyses presented above (Gilead and Goren 1994 and this work) do not support a reconstruction of the Chalcolithic society as presented by Bourke.

The fact that tabular scrapers were found at most sites, despite their nature, size or geographic location, most being broken after use, suggests that their economic value was not high, these tools being accessible to all. Therefore, tabular scrapers did not indicate higher social or economic status, but rather had a symbolic meaning with a cultural implication. Since their function is unknown, their symbol is obscure as well. An activity involving cutting was related to them. Therefore, a ritual cutting is suggested as a possible symbolic meaning of tabular scrapers, their presence in burials being interpreted as offering tools to be used in this activity.

Shalev and Northover (1987) analysed the Shiqmim copper metallurgy and arrived to the conclusion that there are marked differences between copper used for prestige items, imported from a more specialized production centre and that used for large tools, locally made (Shalev and Nortover 1987). A similar dichotomy was noted for pottery, basalt and flint tool production.

Apart from a common religious/cultic background, Southern Levant communities shared few characteristics of lithic material culture. The impression is of a highly differentiated cultural mosaic, each zone being isolated, with specific characteristics, shaped by adaptations to environment, quality of raw materials and knapping choices of local artisans.

The Chalcolithic cultural mosaic of Southern Levant The lithic assemblages’ techno-typological analysis supported the previous division of the Southern Levant in five main cultural zones (for example Anati 1963):

The cultural entities of each zone evolved probably locally, movements of people being apparently testified at Ramot 3 and Ramot Nof in the Be’er-Sheva basin (related to the Besorian of Western Negev) or at Kaukab, its lithic assemblage sharing characteristics with those of Northern Negev sites.

1. the Northern zone, including the Golan heights, Hula and Northern Jordan valley and Galilee mountains, extending into the Jordanian uplands, apparently with variations in material culture, as evidenced when comparing between Kaukab in the Galilee zone, Tel Turmus and the Golan sites7. 2. a poorly represented central Israel part, with cultural ties to Northern Negev. 3. the Northern Negev area, mostly the Be’er-Sheva basin, along main wadis of the region, confined in the north by the Hebron mountains, in the east by the Judean desert and in the south by the Negev highlands.

The presence of perforated tools in some Negev sites implies a contemporaneity with northern assemblages, the probable origin of these items; sickle-blades on banded flint at Kh. Aliya East link the site to the Negev. Sites of the desert do not share material culture with other regions; they might have been isolated, both culturally and economically, or these ties are not reflected in the lithic remains. A long existence of the Timnian and Eilatian was proposed (Rothenberg and Glass 1992, Avner et. al 1994). The present work only partially succeeded in anchoring in time, or better defining, the characteristics of these industries.

7

The site of Teleilat Ghassul stands out as a not yet fully understood phenomenon, with links both to northern and Negev areas. It may be viewed as a center that had a cultural influence along a north and east axis, or, subsequently, as a center that absorbed cultural influences from the respective areas.

118

CHAPTER IX – CONCLUSIONS

were used on poorer quality raw materials (such as at Meitar); there is no reason to assume that a preferred flake was the aim of using discoidal cores – no such tools were found in the tool-kit of sites where discoidal cores were collected. The question of discoidal cores found in Timnian or Eilatian assemblages from Sinai has no single answer. Two possibilities were raised: some of the discoidal cores belong in fact to dispersed Middle Palaeolithic assemblages, while the others are regular cores, as appear also at Tel Sheva for example, without being characteristic of a region / period / culture.

When clustering sites of the Nahal Besor area, two groups were discerned by Gilead: sites A, B, D and M, belonging to an early cultural entity, the Besorian, and sites E and O, similar to Grar – Be’er-Sheva sites. Gilead and Goren 1995). The present work failed to isolate major differences between Besorian and Be’er-Sheva lithic assemblages: the huge amount of banded flint blade cores and semi-translucent bladelet cores (and their products, sickle-blades and microliths respectively), found at Ghazzeh site A are typical of Northern Negev assemblages. Be’er-Sheva scrapers are present in Besorian sites as well. The Besorian sites of the Be’erSheva valley, Ramot 3 and Ramot Nof have relatively small lithic assemblages and without any distinct characteristics. Apparently, cultural/temporal differences between Besorian and Be’er-Sheva sites are not reflected in lithic assemblages. The use of banded and semitranslucent flints and their related reduction methods may have started in the earlier Besorian phase and continued later as well.

A further technological hallmark is the multitude of techniques. Simple ones, common in the Lower Palaeolithic (such as production of choppers and handaxes - picks), appear alongside discoidal cores typical of the Middle Palaeolithic period, blade and bladelet techniques of the Upper Palaeolithic, the microburin technique typical of the Epi-Palaeolithic and elaborated techniques, unique for the Chalcolithic, producing perforated discs and other perforated tools. Thus, a flexible industry and a high adaptability to raw material constraints characterize the Chalcolithic flint tool production. Whenever aesthetical aspects were expected, as for the production of bifacials found at the Peqi’in burial cave, Chalcolithic knappers obtained high results; in other cases, little effort was invested in the production of most tool-types.

Epstein (1998) highlights contact points between the Northern Levant and the Golan sites, the later being under Mesopotamian and Anatolian spheres of influence. Aspects of material culture (such as architecture, pottery characteristics, ground stone artefacts and iconographic symbols) link the Late Neolithic layers of Byblos with the Golan sites (Epstein 1998:35). A possible link to Neolithic flint assemblages is the presence of pièces esquilles, wide sickle-blades and the high amount of chisels. The appearance of pièces esquilles may be due to a local preference of knappers to use an anvil, while the size of sickle-blades is relatable to their non-standardized production. However, the continuity from the Neolithic to the Chalcolithic cannot be rejected on lithic grounds.

Another characteristic of Chalcolithic industries is the selection of raw materials for the production of particular tool-types; tabular flint was exclusively used for the production of tabular scrapers, while microliths were made on semi-translucent flint. Banded flint, a common pebble type in Be’er-Sheva wadis, was used for the production of sickle-blades mainly in Northern and western Negev, and to a lesser extent at Kh. Aliya and probably at Ghassul as well.

A local, gradual development was observed in the ceramic repertoire of Teleilat Ghassul, its lower levels containing ceramic elements attributable to the “Besorian” and “Wadi Raba” cultural entities, while the upper layers contained “Ghassulian” pottery. Lovell (2001) recognizes 4 traditions: Be’er-Sheva (Negev), Neve Ur (north Jordan valley), Ghassul (south Jordan valley) and Golan (north Israel/Lebanon and Northern Jordan). This cultural reconstruction of Lovell matches to the one presented in this work.

Several tool-types were recognized, both in past and current research, to be characteristic of the Chalcolithic period: micrograttoirs, already identified and defined in the past (Gilead 1984) are bladelets on semi-translucent flint, with one or both ends modified by retouch, which occasionally appears along the edges as well. Microliths were found in all but the desert areas of Sinai and Southern Jordan, usually in small amounts; exceptions are some sites in western Negev, where they can reach a third of the tools. These tools were produced in workshops and also in regular households.

Definition of cultural flint hallmarks Several cultural hallmarks could be isolated. As a general rule, Chalcolithic flint industries are oriented towards the production of flakes, the commonly applied technique for their production was a direct blow, probably with a hard hammer, on an unfaceted striking platform of unipolar cores. These were in most cases unprepared, a simple blow on one pointed end of a pebble would create a flat surface from which further removals were knapped. Aside from these, discoidal cores, resembling the Levallois method were found. Apparently, discoidal cores

A recently defined tool-type is the “Be’er-Sheva” scraper: on a primary flake, with cortical or flat butt, the working edge shaped by semi-abrupt retouch on the distal end, along one or both edges, or covering its circumference (Gilead and Hermon, in press). Their absence in other areas suggests to view “Be’er-Sheva” scrapers as an indicative of Negev assemblages.

119

CHAPTER IX – CONCLUSIONS

tabular scrapers, solely found in the Golan sites. Their morphological characteristics were described in details in other publications (Noy 1998). Suffice it to say that only isolated items were found in Negev assemblages, most probably brought in within an exchange network. At this stage of research, little is known about their mode of manufacture and function. Another characteristic tool of Northern assemblages is a narrow chisel.

An additional characteristic tool-type of Negev assemblages is a sickle-blade made on banded flint, with a finely denticulated working edge, a straight back and double truncated. Technological aspects isolate it from sickle-blades of other regions: produced from unipolar cores, they are narrow (mean width of 11.4 mm.) and with a triangular or trapezoid cross-section. An important aspect, found only on sickle-blades from the Negev, is the application of a technique resembling the microburin technique of the late Upper Palaeolithic. It involves the preparation of a notch at the distal end of blades to be modified into sickle-blades and snapping off part of the blade, near the notch. This technique was applied to control the length of the components to be hafted into a sickle, the handle having probably a fixed length.

The concluding remarks of this work regard two aspects: one is the reconstruction of human life in the Chalcolithic Southern Levant and the second one concerns the methodology of research. The results of the lithic industries analysis, supported by other research fields as well, shows a society organized in several cultural clusters, corresponding to geographical zones and with an economy mainly based and adapted to the exploitation of local resources, partially augmented by an exchange network of specific products, some of which were manufactured by part-time specialists.

A technological characteristic of Negev assemblages is the preparation mode of axes, adzes and to a lesser extent chisels as well. Often, they were shaped from three striking platforms, two located along the edges, and an additional one along the longitudinal axis of the item, forming a kind of crest. The cross-section is often triangular, each point representing one striking platform. An additional characteristic is the preparation of the working edge, using a transversal blow, usually on each side of the item. There is no evidence to support that this method was solely applied to re-sharpen the working edge. Therefore, it is suggested to view it as a tradition mark on bifacial production of the Negev.

Flint tool production was organized on mainly two levels: the manufacture of basic, simple tools, such as retouched flakes, scrapers or notches/denticulates, and the other, requiring a more rigid technique, of tabular scrapers, bifacials, microliths and sickle-blades. The production of the first group of tools was flexible, applying various knapping techniques on the local raw materials, while the second group of tools was made on particular raw materials and by a rigid reduction sequence. The large variety of applied knapping methods are indicatives for highly skilled artisans; traditional aspects were observed on particular reduction schemes, such as those that produced sickle-blades in the Negev.

Another bifacial type, common in Negev assemblages, is the pick: prepared from large pebbles, one of the ends being pointed by a series of removals, usually from three directions. The butt, cortex covered, often shows battering signs. Their massive aspect, and the battering signs suggest their function as heavy-duty tools, possibly involved in digging underground structures. Another possibility is that they were used in the copper industry, as suggested by their presence at Timna 39, where evidences of metallurgical activities were identified. Picks are practically absent from other regions (single items were found at Golan sites 12 and 22).

Where the research methodology is concerned, the main conclusion is that there is a need for the standardization of research, in order to perform a fruitful comparison and synthesis of assemblages. The other remark regards the application of the chaîne opératoire concept, which proved to be an efficient method for the analysis of lithic assemblages of the Chalcolithic Southern Levant.

Core-choppers are common in Negev assemblages; they are absent from most other parts, except a few items in the Golan sites. They are simple globular pebbles, with a few removals forming a working edge and most of the circumference cortex covered. It is unclear whether these were used as tools (some have battering signs on their butt and/or damaged working edges), or they are flake cores, used to produce a few primary flakes (possibly blanks for scrapers). The presence of core-choppers both in the Negev and at Golan sites may be related to the exploitation of similar flint resources, i.e. wadi pebbles. One of the single cultural hallmarks of Northern assemblages are perforated tools: discs, found in the Golan area and further, in the Jordan and the Hula valleys (see Newe Ur and Tel Turmus), stars and perforated

120

REFERENCES

1973

Geological Map of Israel and Adjacent Countries. Geological Survey of Israel, Jerusalem.

Adams, W.Y., Adams, E.W. 1991 Archaeological Typology and Practical Cambridge University Press, Cambridge.

oriente, Berlin. Bar-Yosef, O., Belfer, A., Goren, A., Smith, P. 1977 The Nawamis near Ein Huderah (Eastern Sinai). Israel Exploration Journal 27:65-88.

Typology.

Barkai, R. 2000 Flint and Stone Axes as Cultural Markers: socio-economic Changes as Reflected in Holocene Flint Tool Industries of the Southern Levant. unpublished Ph.D. (in Hebrew), Tel Aviv University, Tel Aviv.

Albright, W.F. 1932 The Chalcolithic age in Palestine. Bulletin of the American Schools of Oriental Research 48:10-13. Alon, D. 1977 Nahal Besor - site D. Hadashot Archeologiot 61-62:42.

Baruch, U. 1986 The Late Holocene vegetational history of Lake Kinneret (Sea of Galilee), Israel. Paléorient 12:37-48.

Alon, D., Levy, T. E. 1980 Preliminary note on the distribution of Chalcolithic sites on the wadi Be’er-Sheba-Lower Besor drainage systems. Israel Exploration Journal 30:140-147. 1989

Beit Arieh, I., Gophna, R. 1977 A note on a Chalcolithic site in Wadi Arabah. Tel Aviv 4.

The archaeology of cult and the Chalcolithic sanctuary at Gilat. Journal of Mediterranean Archaeology 2:163-221.

Belfer-Cohen, A. 1988 The Natufian Settlement at Hayonim Cave. Ph.D. dissertation (unpublished), Hebrew University, Jerusalem.

Amiran, R. 1977 Pottery from the Chalcolithic site near Tell Delhamiya and some notes on the character of the Chalcolithic-Early Bronze Age I transition (Hebrew). Eretz-Israel 13:48-56.

Ben-Tor, A., Portugali, Y. 1987 Tell Qiri a village in the Jezreel valley, Jerusalem. Betts, A.V.G. 1992 Tell el-Hibr: a rock shelter occupation of the fourth millennium B.C.E. in the Jordan Badiya. Bulletin of the American Schools of Oriental Research 287:5-24.

Anati, E. 1963 Palestine Before the Hebrews. Jonathan Cape, London. Andreu, G. 1992 Images de la Vie Quotidienne en Égypte au Temps des Pharaons. Hachette, Paris.

Blackham, M. 1999 Tulaylat Ghassul: an appraisal of Robert North's excavations (1959-60). Levant XXXI:19-64.

Audouze, F. 1985 L'Apport des sols d'habitat a l'étude de l'outillage technique. In La signification culturelle des industries lithiques, edited by M. Otte, pp. 57-68. British Archaeological Reports, Oxford.

Boëda E, Geneste, J -M, Meignen, L. 1990 Identification de chaînes opératoires lithiques Paléolithique ancien et moyen. Paleo 2:43-80.

du

Bordes, F. 1950 Principe d'une méthode d'étude des techniques et de la typologie du Paléolithique ancien et moyen. L'Anthropologie 54:19-34.

Aurenche, O., Cauvin, J., Cauvin, M.-C., Copeland, L., Hours, F., Sanlaville, P. 1981. Chronologie et organisation de l’espace dans le Proche Orient de 12,000 à 5600 av J.C. in Sanalville, P. and Cauvin, J. (eds.) Préhistorire du Levant. CNRS, Paris. Pp. 571-601.

1961

Avner, U. 1996 Ancient Masseboth Stones in the Negev and Sinai and their Interpretation. M.A. (unpublished, in Hebrew), The Hebrew University, Jerusalem.

Burton, M., Levy, T.E. 2001 The Chalcolithic radiocarbon record and its use in Southern Levantine archaeology. Radiocarbon 43(3): 1223-1246.

Typologie du Paléolithique Ancien et Moyen. Delmas, Bordeaux.

Bourke, S.J. 1997 The "pre-ghassulian" sequence at Teleilat Ghassul: Sydney university excavations 1975-1995. In The Prehistory of Jordan, II. Perspectives from 1997, edited by H.G.K. Gebel, Kafafi, Z., Rollefson, G.O., pp. 395-427. ex oriente, Berlin.

in prep. Archaeological survey and excavations in Sinai. Avner, U., Carmi, I., 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 Late Quaternary Chronology and Paleoclimates of Eastern Mediterranean, edited by O. Bar-Yosef, Kra, R. S., pp. 265-300. RADIOCARBON, Tucson.

2001

Banning, E.B., Lasby, D., Blackham, M. 1998 Excavations at WZ121, a Chalcolithic site at Tubna, Wadi Ziqlab. Annual of the Department of Antiquities of Jordan 42:141-159.

The Chalcolithic period. In The Archaeology of Jordan, edited by B. MacDonald, Adams, R., Bienkowski, P., pp. 107-162. Sheffield Academic Press, Sheffield.

Boydston, R.A. 1989 A cost/benefit study of functionally similar tools. In Time, Energy and Stone Tools, edited by R. Torrence, pp. 67-77. Cambridge University Press, Cambridge.

Bar-Adon, P. 1980 The Cave of the Treasures. Israel Exploration Society, Jerusalem.

Brezillon, N. M. 1968 La dénomination des objets de pierre taillée. CNRS, Paris.

Bar-Yosef, O. 1970 The Epipalaeolithic Cultures of Palestine. Ph.D. dissertation (unpublished), Hebrew University, Jerusalem.

Carmi, I., Epstein, C., Segal, D. 1995 Radiocarbon dates from Chalcolithic sites in the Golan. Atiqot XXVII: 207-209.

1994

Clarke, D. L. 1978 Analytical Archaeology. Methuen & Co. LTD, London.

Form, function and numbers in Neolithic lithic studies. In Neolithic Chipped Stone Industries of the Fertile Crescent, edited by H. Gebel, G., Kozlowski, S., K., pp. 5-15. ex

Cohen, R.

121

REFERENCES

1988

Settlement on the Negev highlands from the fourth millennium BCE to the fourth century BCE. Qadmoniot 21:62-81 (Hebrew).

Epstein, C., Noy, T. 1988 Observations concerning perforated flint chalcolithic Palestine. Paléorient 14:133-141.

Commenge-Pellerin, C. 1987 La Poterie d’Abou Matar et de l’Ouadi Zoumeili (Beershéva) au IVe millénaire avant l'ère chrétienne. Association Paleorient, Paris. 1990

Fabian, P. in prep Excavations at Mitcham C, Be'er-Sheva. Atiqot.

La Poterie de Safadi (Beersheva) au IVe millénaire avant l'ère chrétienne. Association Paleorient, Paris.

Fabian, P., Hermon, S., Goren, Y. 2004 Ramot 3: a Chalcolithic open-air site in the Northern Be’erSheva valley. Atiqot 47:57-80. Finkelstein, I. 1996 Toward a new periodization and nomenclature Archaeology of Southern Levant. In The Study of Near East in the Twenty-First Century, edited Cooper, Schwartz, G.M., pp. 103-124. Eisenbrauns, Lake, Indiana.

Remarques sur le Chalcolithique récent de Tell ash-Shuna. Revue Biblique LXVIII:546-556.

Copeland, L., Vita-Finzi, C. 1978 Archaeological dating of geological deposits in Jordan. Levant X:10-25.

Frumkin, I., Zak, I., Carmi, M., Magaritz A. 1991 The Holocene climatic record of the salt caves of Mount Sedom, Israel. The Holocene 1:191-200.

Cziesla, E. 1990 On refitting of stone artifacts. In The Big Puzzle International Symposium on Refitting Stone Artifacts, edited by E. Cziesla, pp. 9-45. Eickhoff, N. Arts, D. Winter, Bonn.

Gal, Z., Smithline, H., Shalem, D. 1997 A Chalcolithic burial cave in Peqi'in, Upper Galilee. Israel Exploration Journal 47:145-154.

Danin, A. 1985 Paleoclimates in Israel: evidence from weathering patterns of stones in and near archaeological sites. Bulletin of the American school of Oriental Research 259:33-43.

Garanjer, J. 1992 La préhistoire dans le Monde. Presses universitaires de France, Paris. Garfinkel, Y. 1992 The Material Culture in the Central Jordan Valley in the Pottery Neolithic and Early Chalcolithic Periods. Ph.D. (unpublished), Hebrew University, Jerusalem.

Dollfus, G., Kafafi, Z., Rewerski, J., Vaillant, N., Coquegniot, E., Desse, J., Neef, R. 1988 Abu Hamid, an early fourth millennium site in the Jordan Valley. In The Prehistory of Jordan, edited by A. N. Garrard, Gebel, H. G., pp. 567-601. British Archaeological Reports International Series 369, Oxford.

1993

The Yarmukian culture in Israel. Paléorient 19:115-135.

1999a

Radiometric dates from eight millennium B.P. Israel. American Schools of Oriental Research, Bulletin of the 315:1-15.

1999b

Neolithic and Chalcolithic Pottery of the Southern Levant. The Institute of Archaeology, the Hebrew University of Jerusalem, Jerusalem.

Dunnell, R. C. 1971 Systematics in Prehistory. Free Press, New York. Eddy, F.W., Wendorf, F. and associates 1999 An Archaeological Investigation of the Central Sinai, Egypt. American Research Center in Egypt, Inc. Cairo, and University Press of Colorado, Boulder.

Garfinkel, Y., Hermon, S. in prep. Excavations at the Chalcolithic site of Meitar.

Ehrlich, A. 1973 Quaternary Diatoms of the Hula Basin (Northern Israel). Geological Survey, Jerusalem.

Garstang, J. 1935 Jericho: city and necropolis (fifth report). Liverpool Annals of Archaeology and Anthropology 22:143-184.

Eisenberg, E. 1989 The Chalcolithic and Early Bronze Age I occupations at Tel Teo. In L'Urbanisation de la Palestine a l'Age du Bronze Ancien, edited by P. de Miroschedji, pp. 29-40. British Archaeological Reports International Series 527, Oxford.

Gat, J. R. 1981 Paleo-climate conditions in the Levant as revealed by the isotopic composition of paleowaters. Israel Meteorological Research Papers III:13-28.

Eldar, I., Baumgarten, Y. 1985 Neve Noy, a Chalcolithic site of the Be’er Sheba culture. Biblical Archaeologist 48:134-139.

Genz, H. 1997 Problems in defining a Chalcolithic for Southern Jordan. In The Prehistory of Jordan, II. Perspectives from 1997, edited by H.G.K. Gebel, Kafafi, Z., Rollefson, G.O., pp. 441-448. ex oriente, Berlin.

Elliot, C. 1978 The Ghassulian culture in the Levant. Levant X:37-54.

Gero, J. M. 1989 Assessing social information in material objects: how well do lithics measure up? In Time, Energy and Stone Tools, edited by R. Torrence, pp. 92-106. Cambridge University Press, Cambridge.

Epstein, C. 1978 A new aspect of Chalcolithic culture. Bulletin of the American Association of Oriental Research 229:27-45. A Pottery Neolithic site near Tel Qatif. Israel Exploration Journal. 34 : 209-219.

1998

The Chalcolithic Culture of the Golan. Israel Antiquity Authority, Jerusalem.

of the Ancient by J.S. Winona

Fitzgerald, G.M. 1934 Excavations at Beth-Shean in 1933. Palestine Exploration Fund Quarterly 66:123-134.

Cowgill, G. L. 1967 Comments on Chang, Major aspects of the interrelationship of archaeology and ethnology. Current Anthropology 8:236237.

1984

from

Everitt, B. 1974 Cluster Analysis. John Wiley & Sons, New York.

de Contenson, H. 1960 Three soundings in the Jordan valley. Annual of the Department of Antiquities of Jordan 4-5:12-98. 1961

tools

Geyh, M. A. 1994 The paleohydrology of the Eastern Mediterranean. In Late Quaternary Chronology and Paleoclimates of the Eastern

122

REFERENCES

Mediterranean, edited by R S Kra O Bar-Yosef, pp. 131147. RADIOCARBON, Tucson.

Goodfriend, G. A., Magaritz, M., Carmi, I. 1986 A high stand of the Dead Sea level at the end of the Neolithic period: paleoclimatic and archaeological implications. Climatic Change 9:349-356.

Gilead, I. 1984 The Micro-endscaper: A new tool type of the Chalcolithic period. Tel Aviv 11:3-10. 1987

A new look at Chalcolithic Archaeologist 50:110-117.

Be’er-Sheba.

1988

The Chalcolithic period in the Levant. Journal of World Prehistory 2:397-445.

1989

Chalcolithic sites in Beit Netofa valley, lower Galilee, Israel. Paléorient 15:263-267.

1990

The Neolithic-chalcolithic transition and the Qatifian of the Northern Negev and Sinai. Levant XXII:47-63.

1994

The history of the Chalcolithic settlement in the Nahal Be’er Sheva area: the radiocarbon aspect. Bulletin of the American Schools of Oriental Research 296:1-13.

1995

Grar a Chalcolithic Site in the Northern Negev. Ben-Gurion University Press, Be’er-Sheva.

2007

The Besorian: A Pre-Ghassulian Cultural Entity. Paléorient 33(1), in press.

Gopher, A. 1993 Sixth-fifth millennia B.C. settlements in the coastal plain, Israel. Paléorient 19:55-63.

Biblical

Gopher, A., Gophna, R. 1993 Cultures of the eight and seventh millennia B.P. in the Southern Levant: a review for the 1990s. Journal of World Prehistory 7:297-352. Gopher, A., Rosen, S.A. in prep The chipped stone assemblage of Tel Teo. Gophna, R. 1979 The settlement of the North-Western Negev during the Chalcolithic period (the fourth millennium BC). In The Land of the Negev, edited by A. Shmueli, Grados, Y., pp. 203-208 (Hebrew). Minister of Defence Publishing House, Tel Aviv. Goren, Y. 1990 The "Qatifian Culture" in Southern Israel and Transjordan: additional aspects for its definition. Mitekufat Haeven 23:100*-112*.

Gilead, I., Alon, D. 1988 Excavations at proto-historic sites in the Nahal Besor and the Late Neolithic of the Northern Negev. Mitekufat Haeven 21:109*-130*.

Goren, Y., Fabian, P. 2002 Kissufim Road A Chalcolithic Mortuary Site, Israel Antiquities Authority Reports 16, Jerusalem. Goring-Morris, N.A. 1987 At the Edge: Terminal Pleistocene Hunter-Gatherers in the Negev and Sinai, British Archaeological Reports International Series 361, Oxford.

Gilead, I., Fabian, P. 1995 The knapped limestone series from Grar. In Grar: a Chalcolithic Site in the Northern Negev, edited by I. Gilead. Ben-Gurion University Press, Be’er-Sheva. 2001

Goring-Morris, A., N., Belfer-Cohen, A. 1998 The articulation of cultural processes and late Quaternary environmental changes in Cisjordan. Paléorient 23:71-93.

Nevatim: a site from the Chalcolithic period in the Northern Negev. In Settlement, Civilization and Culture, edited byA. Meir and I. Baruch, pp. 67-86. Bar-Ilan University Press, Ramat-Gan. (in Hebrew).

Grigson, C. 1987 Shiqmim: Pastoralism and other aspects of animal management in the Chalcolithic of the Northern Negev. In Shiqmim I, Studies Concerning Chalcolithic Societies in the Northern Negev Desert, edited by T. E. Levy, pp. 219-241. B.A.R. International Series 356, Oxford.

Gilead, I., Goren, Y. 1986 Stations of the Chalcolithic period in Nahal Sekher, Northern Negev. Paléorient 12:83-90. 1989

1995

Petrographic analysis of fourth millennium B.C. pottery and stone vessels from the Northern Negev, Israel. Bulletin of the American Schools of Oriental Research 275:5-14.

Gustavson-Gaube, C. 1987 The esh-Shuna North: 1984 and 1985. In Studies in the History and Archaeology of Jordan, edited by A. Hadidi, pp. 237-241. Department of Antiquities, Amman.

The pottery assemblages from Grar. In Grar, A Chalcolithic Site in the Northern Negev, edited by I. Gilead, pp. 137-222. Ben-Gurion University Press, Be’er-Sheva.

Hanbury-Tenison, J. W. 1986 The Late chalcolithic to Early Bronze I Transition in Palestine and Transjordan. British Archaeological Reports International Series 311, Oxford.

Gilead, I., Hermon, S. in prep. The Lithic Assemblages from Abu Matar and Bir es-Safadi (Be’er Sheva), Two Chalcolithic sites in the Northern Negev. CRFJ, Jerusalem.

Hayden, B. 1989 From chopper to celt: the evolution of resharpening techniques. In Time, Energy and Stone Tools, edited by R. Torrence, pp. 7-16. Cambridge University Press, Cambridge.

Gilead, I., Hershman, D., Marder, O. 1995 The flint assemblages from Grar. In Grar, A Chalcolithic Site in the Northern Negev, edited by I. Gilead, pp. 223-280. Ben-Gurion University Press, Be’er-Sheva.

Henessy, J. B. 1969 Preliminary report on the first season of excavations at Teleilat Ghassul. Levant I:1-24.

Goldberg, P., Rosen, A. M. 1987 Early Holocene paleoenvironments of Israel. In Shiqmim I, Studies Concerning Chalcolithic Communities in the Northern Negev, Israel., edited by T. Levy, E., pp. 23-33. B.A.R. International Series, Oxford.

1982

Golden, J. M. 1998 The Dawn of the Metal Age. (unpublished) Ph.D. thesis, University of Pennsylvania.

Teleilat Ghassul: its place in the Archaeology of Jordan. In Studies in the History and Archaeology of Jordan, edited by A. Hadidi, pp. 55-58. Department of Antiquities, Amman.

Henry, D. O. 1985 Late Pleistocene environment and Palaeolithic adoption in Southern Jordan. In Studies in the History and Archaeology of Jordan, edited by A. Hadidi, pp. 67-77. Department of Antiquities, Amman.

Goodfriend, G. A. 1988 Mid-Holocene rainfall in the Negev desert from 13C of land snail shell organic matter. Nature 333:757-760.

1992

123

Seasonal movements of fourth millennium pastoral nomads in the Wadi Hisma, Southern Jordan. In History and

REFERENCES

Archaeology of Jordan, edited by M. Zaghloul, 'Amr, K., pp. 137-143. Department of Antiquities, Amman. 1995

1994

Prehistoric Cultural Ecology and Evolution. Plenum Press, New York.

Karmon, Y. 1971 Israel a regional geography. Wiley-Interscience, London.

Hermon, S. in prep. The lithic assemblage of Tel-Sheva.

Kenyon, C. 1960 Archaeology in the Holy Land. Frederick A. Praeger, New York.

Hershman, D. 1987 Intra-site Variability of Fourth Millennium Flint Assemblages from Southern Israel. (unpublished) M.A. thesis, The Hebrew University, Jerusalem. (in Hebrew)

Kerner, S. 1997a Specialization in the Chalcolithic in the Southern Levant. In The Prehistory of Jordan, II. Perspectives from 1997, edited by H.G.K. Gebel, Kafafi, Z., Rollefson, G.O., pp. 419-427. ex oriente, Berlin.

Horowitz, A. 1974 Preliminary palynological indications as to the climate of Israel during the last 6,000 years. Paléorient 2:407-414. 1979

The Quaternary of Israel. Academic Press, New York.

1989

Prehistoric cultures of Israel: correlations with oxygen isotope scale. In Investigations in South Levantine Prehistory, edited by O. Bar-Yosef, Vandermeersch, B., pp. 5-17. British Archaeological Reports International Series, Oxford.

1997b

Status, perspectives and future goals in Jordanian Chalcolithic research. In The Prehistory of Jordan, II. Perspecties from 1997, edited by H.G.K. Gebel, Kafafi, Z., Rollefson, G.O., pp. 465-474. ex oriente, Berlin.

Khalaily, M. 1999 The Flint Assemblage of Layer V at Hagoshrim: a Neolithic Assemblage of the Sixth Millennium B.C. in the Hula Basin. (unpublished) M.A. thesis, The Hebrew University of Jerusalem, Jerusalem. (in Hebrew).

Hours, F. 1974 Remarques sur l'utilisation de listes-types pour l'étude du Paléolithique supérieur et de l'Epipaléolithique du Levant. Paléorient 2:3-18.

Khalaily, H., Hermon, S. in press Gat Govrin (Nahal Komem): a late Chalcolithic site in the Northern Negev. Atiqot.

Ibrahim, M., Sauer, J. A., Yassine, K. 1975 The East Jordan valley survey. Bulletin of the American Society for Oriental Research 222:41-66.

Kislev, M. 1987 Chalcolithic plant husbandry and ancient vegetation at Shiqmim. In Shiqmim I, Studies Concerning Chalcolithic Societies in the Northern Negev Desert, edited by T. E. Levy, pp. 251-279. British Archaeological Reports International Series 356, Oxford.

Inizan, M-L., Roche, H., Tixier, J. 1992 Technology of Knapped Stone. C.R.E.P., Meudon. Isaar, H., Bruins, A. 1983 Special climatic conditions in Sinai and the Negev during the most Upper Pleistocene. Paleogeography, Paleoclimatology, Paleoecology 43:67-72.

Koeppel, R. 1938 Tuleilat Ghassul. Quarterly of the Department of Antiquities of Palestine VI:225-226.

Joffe, A. H. 1993 Settlement and Society in the Early Bronze Age I and II, Southern Levant. Sheffield Academic Press, Sheffield.

Koeppel, R., Senes, H., Murphy, J.W., Mahan, G.S. 1940 Teleilat Ghassul II (1932-1936). Pontifical Biblical Institute, Rome.

Joffe, A. H., Dessel, J. P. 1995 Redefining Chronology and Terminology for the Chalcolithic of the Southern Levant. Current Anthropology 36:507-518.

Kozloff, B. 1974 A brief note on the lithic industries of Sinai. Museum Ha'aretz Yearbook 15/16:35-49. Laplace, G. 1966 Recherches sur l'origine et l'évolution des complexes leptolithiques. E de Boccard, Paris.

Kafafi, Z. 1987 The Pottery Neolithic in Jordan in connection with other Near Eastern Regions. In Hadidi, A. (ed.) Studies in the History and Archaeology of Jordan, Vol. III. Department of Antiquities, Amman. Pp. 33-39.

Lee, R. E. 1973 Chalcolithic Ghassul: New Aspects and Master Typology. (unpublished) PhD. dissertation, The Hebrew University, Jerusalem.

Kaplan, J. 1958a Excavations at Teluliyot Batashi in the vale of Sorek. EretzIsrael 5:9-24 (in Hebrew). 1958b

Prehistoric technology: a cognitive science ? In The Ancient Mind, edited by C. Renfrew and E. B. W. Zubrow, pp. 152165. University Press, Cambridge.

Lemonnier, P. 1986 Anthropology as technical systems. Anthropological Archaeology 5:147-187.

Excavations at Wadi Rabah. Israel Exploration Journal 8:149-160.

Journal

of

Karlin, C. 1992 Connaissance et savoir-faire: comment analyser un processus technique en Préhistoire. Paper presented at the conference Tecnologia y cadenas operativas liticas, Barcelona.

1990

Topsy turvy techniques. Remarks on the social representation of techniques. Archaeological Review of Cambridge 9:27-37.

Karlin C, Bodu P, Pigeot N, Ploux S 1992 Some socio-economic aspects of the process of lithic reduction among groups of hunter-gatherers of the ParisBasin area. In The use of tools by humans and non-human primates, edited by Chavaillon J Berthelet A, pp. 318-340. Clarendon Press, Oxford.

1945

Milieu et techniques. Albain Michel, Paris.

1964

Le geste et le parole I - techniques et langage. Albain Michel, Paris.

1965

Le geste et le parole II - la memoire et les rytmes. Albin Michel, Paris.

Leroi-Gourhan, A. 1943 L'Homme et la matiere. Albin Michel, Paris.

Karlin, C., Julien, M.

Leroi-Gourhan A., Bailloud G., Laming-Emperaire A.

124

REFERENCES

1966

La Prehistoire. PUF Nouvelle Clio, Paris.

1940

Levi-Strauss, C. 1976 Structural anthropology. Basic Books, New York.

The stone industry of the campaign 1936. In Teleilat Ghassul II, edited by R. Koeppel, pp. 89-115. Institute Biblique Pontifical, Rome.

Mallon, A. 1930 Les fouilles de l'Institut Biblique Pontifical dans la Vallée du Jourdain. Biblica XI:3-22.

Levy, T. E. 1983 The emergence of specialized pastoralism in the Southern Levant. World Archaeology 15:15-36.

Mallon, A., Koeppel, R., Neuville, R. 1934 Teleilat Ghassul I. Institute Biblique Pontifical, Rome.

1986

The Chalcolithic period. Biblical Archaeologist 49:82-108.

1992

Radiocarbon chronology of the Be’ersheva culture and predynastic Egypt. In The Nile Delta in transition: 4th - 3rd millennium B.C., edited by E. C. M. van den Brink, pp. 345356. Israel Exploration Society, Jerusalem.

Marder, O., Hermon, S. in prep. A tabular scraper cache from the burial site of Eilat.

Cult, metallurgy and rank societies - Chalcolithic period. In The Archaeology of Society in the Holy Land, edited by T E Levy, pp. 226-245. Leicester University Press, London.

Marks, A.E. (editor) 1976 Prehistory and Palaeoenvironment in the Central Negev, Israel. SMU Press, Dallas.

Levy, T. E. (editor) 1987 Shiqmim I, Studies Concerning Chalcolithic Societies in the Northern Negev Desert, British Archaeological Reports International Series 356, Oxford.

Mauss, M. 1927 Division et proportions des divisions de la sociologie. Année sociologique 2.

1995

in prep. The lithic assemblage of the Chalcolithic site of Tel Turmus.

Levy, T. E., Alon, D. 1982 The Chalcolithic mortuary site near Metzad Aluf, Northern Negev desert: A preliminary study. Bulletin of the American Schools of Oriental Research 248:37-59. 1987a

1987b

1936

Les Techniques du corps. Journal de psychologie 32.

1941

Les techniques et la technologie. Journal de psychologie 41.

McConaughy, M. 1979 Formal and Functional Analysis of Chipped Stone Tools from Bab edh Dra. University Microfilms, Ann Arbor.

Settlement patterns along the Nahal Be’ersheva-Lower Besor: Models for subsistence in the Northern Negev. In Shiqmim I, Studies Concerning Chalcolithic Societies in the Northern Negev Desert, edited by T. E. Levy, pp. 45-138. British Archaeological Reports International Series 356, Oxford.

McNicollt, A.W., Edwards, P.C., Hanbury-Tenison, J., Hennsessy, J.B., Potts, T.F., Smith, R.H., Walmsley, A., Watson, P. 1992 Pella in Jordan 2. Meditarch, Sydney. Melaart, J. 1975 The Neolithic of the Near East. Thames and Hudson, London.

Excavations in Shiqmim cemetrey 3: final report on the 1982 season. In Shiqmim I, Studies Concerning Chalcolithic Societies in the Northern Negev Desert, edited by T. E. Levy, pp. 333-356. British Archaeological Reports International Series 356, Oxford.

Mellars, P. 1996

The Neanderthal Legacy. Princeton University Press, New Jersey.

Levy, T.E., Grigson, C., Holl, A., Goldberg, P., Yorke, R., Smith, P. 1991 Subterranean settlement in the Negev desert, c. 4500-3700 BC. National Geographic Research & Exploration 7:394413.

Milevski, I., Marder, O., Khalaily, H., Sonntag, F.

Levy, T. E., Rosen, S. A. 1987 The chipped stone industry. In Shiqmim I, Studies Concerning Chalcolithic Societies in the Northern Negev Desert, edited by T. E. Levy, pp. 281-294, British Archaeological Reports International Series 356, Oxford.

Miller, R. 1987 Flaked stone from workman's village. In Amarna Reports IV, edited by B.J. Kemp. The Egypt Exploration Society, London.

2006

Miller, J. M. (editor) 1991 Archaeological Survey of the Kerak Plateau. Scholars Press, Atlanta, Georgia.

Lovell, J. L. 2001 The Late Neolithic and Chalcolithic Periods in the Southern Levant. British Archaeological Reports International Series 974, Oxford.

Moore, A. M. T. 1973 The late Neolithic in Palestine. Levant V: 36-68.

Loyd, S., Safar, F. 1945 Tell Hassuna. Journal of Near Eastern Studies 4:255-289. Luz, B. 1982

The flint assemblages, in Yannai, E., 'En Esur ('Ein Asawir) I Excavations at a Protohistoric Site in the Coastal Plain of Israel, pp. 179-210. IAA Reports 31, Jerusalem.

Paleoclimatic interpretation of the last 20,000 yr record of deep sea cored around the Middle East. In Palaeoenvironments and Human Communities in the Western Mediterranean Region in Latter Prehistory, edited by W Van Zeist and J L Bintliff, pp. 41-61. British Archaeological Reports International Series, Oxford.

1982

A four stage sequence for the Levantine Neolithic, ca. 85003750. Bulletin of the American Schools of Oriental Research 264: 1-34.

1985

The development of Neolithic societies in the Near East. In Advances in World Archaeology, edited by F. Wendorf, Close, A. E., pp. 1-69. Academic Press, Orlando.

Nachshoni, P., Goren, Y., Marder, O., Goring-Morris, N. 2002 A Chalcolithic site at Ramot Nof, Be’er-Sheva. Atiqot XLIII:*1-*24 (Hebrew).

MacDonald, E. 1932 Prehistoric Fara, Beth Pelet II. The British School of Archaeology in Egypt, London.

Nasralleh, J. 1948 Une station Ghassoulien du Horan. Revue Biblique LV:81103.

Mackay, E. 1921 The cutting and preparation of tomb-chapels in the Theban necropolis. Journal of Egyptian Archaeology 7:154-168.

Navarro i Barberan, C. 1997 Trois ateliers de taille de silex a Abu Hamid (Jordanie). In The Prehistory of Jordan, II. Perspectives from 1997, edited

Mahan, G. S.

125

REFERENCES

by H.G.K. Gebel, Kafafi, Z., Rollefson, G.O., pp. 383-393. ex oriente, Berlin.

Perrot, J., Ladiray, D. 1980 Tombes et Ossuaires de la Région Cotiere Palestinienne au IVe Millénaire avant l'Ère Chrétienne. Association Paleorient, Paris.

Neev, D., Hall, J. K. 1977 Climatic fluctuations during the Holocene as reflected by the Dead Sea levels. Paper presented at the International Conference on Terminal Lakes, Ogden, Utah.

Perrot, J., Zori, N., Reich, Y. 1967 Neve Ur, un nouvel aspect du Ghassoulien. Israel Exploration Journal 17:201-232.

Neuville, R. 1934 L'outillage en silex. In Teleilat Ghassul, edited by A. Mallon, Koeppel, R., Neuville, R., pp. 55-65. Institut Biblique Pontifical, Rome.

Pigeot, N. 1990 Technical and social actors: flint knapping specialists and apprentices at Magdalenian Etiolles. Archaeological Review from Cambridge 9:126-141.

Neuville, R., Mallon, A. 1931 Les débuts de l'age des métaux dans les grottes du désert de Judée. Syria XII:24-47.

1991

Réflexions sur l'histoire technique del'homme: de l'évolution cognitive a l'évolution culturelle. Paleo 3.

North, R. S. J. 1961 Ghassul 1960 Excavation Report. Biblical Pontifical Institute, Rome.

Ronen, A. 1970 Flint Implements from Southern Sinai: Preliminary Report. Palestine Exploration Quarterly 102:30-41.

Noy, T. 1998

Rosen, A. 1989 Environmental change at the end of Early Bronze Age of Palestine. In L'Urbanisation de la Palestine a l'Age du Bronze Ancien, Bilan et Perspective de Recherches Actuelles, edited by P. de Miroschedji, pp. 247-255. British Archaeological Reports International Series 527, Oxford.

Flint artifacts. In The Chalcolithic Culture of the Golan, edited by C. Epstein. Israel Antiquities Authority, Jerusalem.

Oren, E., Gilead, I. 1981

Chalcolithic Sites in Northeastern Sinai. Tel Aviv 8:25-45.

Orton, C. 1980 Mathematics in Archaeology. Collins, London. Pelegrin, J. 1985 Reflection sur le comportement technique. In L'Apport des sols d'habitat a l'étude de l'outillage technique., edited by M Otte, pp. 72-91. British Archaeological Reports, Oxford. 1990

Prehistoric lithic technology: some aspects of research. Archaeological review from Cambridge 9:116-125.

1991

Les savoir - faire: une très longue histoire. Terrain 16:106113.

1987

The potentials of lithic analysis in the Chalcolithic of the Northern Negev. In Shiqmim I, Studies Concerning Chalcolithic Societies in the Northern Negev Desert, edited by T. E. Levy, pp. 295-312. British Archaeological Reports International Series 356, Oxford.

1997

Lithics After the Stone Age. Altamira Press, Walnut Creek.

Rosen, S. A., Eldar, I. 1993 Horvat Beter revisited: the 1982 salvage excavations. Atiqot 22:13-27. Roshwalb, A. F. 1981 Prehistory in the wadi Ghazzeh: A typological and technological study based on Macdonald Excavations. (unpublished) Ph.D. dissertation, London University, London.

Pelegrin, J., Karlin, C., Bodu, P. 1988 Chaînes opératoires: un outil pour le préhistorien. In Technologie Lithique, edited by J. Tixier, pp. 55-62. CNRS, Paris.

Rowan, Y. M.

Perrot, J. 1952 Les industries lithiques Palestiniennes de la fin du Mésolithique a l'age du Bronze. Israel Exploration Society 2:73-81.

2006

Flint Tool Production at Gilat. In Archaeology, Anthropology and Cult The Sanctuary at Gilat, Israel, edited by T.E. Levy, Equinox, London.

1955

The excavations at Tell Abu Matar near Be’ersheba. Israel Exploration Journal 5:17-40, 73-84, 167-189.

Rothenberg, B. 1978 Excavations at Timna site 39. Archaeo-Metallurgy 1:1-20.

1962a

Palestine-Syria-Cilicia. In Courses Toward Urban Life, edited by R. J. Braidwood, Willey, G. R., pp. 147-165. Aldine Publishing Company, Chicago.

1962b

Chronique Archéologique: Nahal Besor. Revue Biblique 69:388-391.

Rothenberg, B., Glass, J. 1992 The beginnings and the development of early metallurgy and the settlement and chronology of western Arabah, from the Chalcolithic period to the Early Bronze Age IV. Levant XXIV:141-157.

1966

La préhistorire palestinienne. Supplément au Dictionnaire de la Bible. Letouzy and Ane, Paris. Pp. 286-446.

1984

Structures d'habitat, mode de la vie et environnent: Les villages souterrains des pasteurs de Be’ersheva dans le Sud d'Israël, au IVe millénaire avant l'ère chrétienne. Paléorient 10:75-92.

1987

Introduction. In La Poterie de Abu Matar et de l"ouadi Zumeili (Be’ersheva) au IVe millénaire avant l'ère chrétienne, edited by C. Commenge-Pellerin, pp. 15-18. Association Paleorient, Paris.

1990

Schiffer, M. B. 1987 Formation Processes of the Archaeological University of New Mexico Press, Albuquerque.

Record.

Schlanger, N. 1994 Mindful technology: unleashing the chaîne opératoire for an archaeology of mind. In The ancient mind, edited by Zubrow E. B. W. Renfrew C, pp. 143-152. University Press, Cambridge. Setton-Karr, W.H. 1905 How the tomb galleries at Thebes were cut and the limestone quarried at the prehistoric flint-mines of the E. desert. Annales du Service des Antiquités de l'Égypte 6:176-187.

Introduction. In La Poterie de Safadi (Be’ersheva) au IVe millénaire avant l'ère chrétienne, edited by C CommengePellerin. Association Paleorient, Paris.

Shalev, S., Northover, P. J. 1987 Chalcolithic metal and metalworking from Shiqmim. In Shiqmim I, Studies Concerning Chalcolithic Societies in the

126

REFERENCES

Northern Negev Desert, edited by T. E. Levy, pp. 357-371. British Archaeological Reports International Series 356, Oxford.

1977

Ussishkin, D. 1970 The Chalcolithic period in Eretz-Israel (Hebrew). Qadmoniot III:110-126.

Schick, T. 1978 Flint Elements, Strata V-I. in Early Arad, by R. Amiran, pp. 58-63, Israel Exploration Society, Jerusalem.

1980

Shipton, G. M. 1939 Notes on the Megiddo Pottery of Strata VI-XX. University of Chicago Press, Chicago.

1970

Stager, L. E. 1992 The periodization of Palestine from Neolithic through Early Bronze Age times. In Chronologies in Old World Archaeology, edited by R. Ehrich, pp. 22-42. University of Chicago Press, Chicago.

Villa, P. 1991

1972

The Yarmukian Culture (in Hebrew), Magnes Press, Jerusalem.

Wreschner, E. E. 1977 Sea level changes and settlement location in the coastal plain of Israel during the Holocene. Eretz Israel 13:277*-282*. Yeivin, E. 1959 Flint tools from Horvat Beter. Atiqot 2:40-44.

Tixier, J. 1963 Typologie de l'Epipaléolithique du Maghreb. Arts et Métiers Graphique, Paris.

Yogev, O. 1983 A fifth millennium sanctuary in the Uvda valley [in Hebrew]. Qadmoniot 16:499-518.

Procèdes d'analyse et questions de terminologie dans l'étude des ensembles industriels du Paléolithique récent et de l'epipaléolithique en Afrique du Nord-Ouest, BurgWartenstein, Austria.

Yorke, R. 1990 A Chalcolithic Chipped Stone Assemblage from the Northern Negev Desert: Phase II (1987-1989) Investigations at Shiqmim, University of Texas at Austin, Austin, Texas.

Tixier, J., Inizan, M-L., Roche, H. 1980 Préhistoire de la pierre taillée, 1: Terminologie et technologie. CREP, Valbonne.

Yorke, R., Levy, T. E. 1991 Use wear analysis of a chalcolithic scraper assemblage from Shiqmim. Journal of the Israel Prehistoric Society (Mitekufat Haeven) 24:112-132.

Trigger, B. C. 1968 Beyond History: the Methods of Prehistory. Holt, Reinhart and Winston, New York.

1994

Tsori, N. 1959

Neolithic and Chalcolithic sites in the valley of Beth-Shean. Palestine Exploration Quarterly 91:44-51.

1962

An archaeological survey of the Beth Shean Valley. Paper presented at The Beth Shean Valley, the 17th Archaeological Convention (Hebrew), Jerusalem.

1967

On two pithoi from the Beth-Shean region and the Jordan Valley. Palestine Exploration Quarterly 99:101-103.

From debitage chips to social models of production: the refitting method in Old World Archaeology. The Review of Archaeology 12:24-30.

Winter, M. C. 1976 The household cluster in the valley of Oaxaca. In The Early Mesoamerican Village, edited by K. Flannery, V., pp. 25-31. Academic Press, New York.

Sugar, A., N. 2000 Archaeometallurgical investigation of the Chalcolithic site of Abu Matar, Israel: a reassessment of technology and its implications for the Ghassulian culture. (unpublished) Ph.D. dissertation, University College, London.

1967

Palestine during the Neolithic and Chalcolithic periods. In Cambridge Ancient History, edited by I. E. S. Edwards, Gadd, C. J., Hammond, N. G. L., pp. 498-553. University Press, Cambridge.

Willey, G., R., Phillips, P. 1958 Method and Theory in American Archaeology. Chicago University Press, Chicago.

Stekelis, M. 1935 Les monuments mégalithique de Palestine. Masson, Paris. Traces of Chalcolithic culture. Eretz Israel 8:88-94 (Hebrew); 71* (English summary).

The Ghassulian shrine at En-Gedi. Tel Aviv 7:1-44.

de Vaux, R. 1961 Les fouilles de Tell el-Far'a, rapport préliminaire sur les 7e, 8e, 9e campagnes. Revue Biblique 68:557-592.

Smith, A.M. II, Stevens, M., Niemi, T. 1997 The southeast Arabah archaeological survey: a preliminary report of the 1994 season. Bulletin of the American Schools of Oriental Research 305:45-72.

1967

Beth She'an in the Chalcolithic period (Hebrew). Eretz-Israel 13:76-82.

Proto-Canaanean blades of the Chalcolithic period. Levant XXVI: 167-174.

Zbenovich, V. in prep. The flint assemblage of the Chalcolithic settlement Kh. el'Alya, East. Zohary, D. 1959 The Pleistocene Period. Hutchinson, London. Zori, N. 1958

127

Neolithic and Chalcolithic Sites in the Valley of Beith-Shan. Palestine Exploration Quarterly 90:45-51.

LIST OF SITES AND THEIR COORDINATES

LIST OF SITES AND THEIR COORDINATES Abu Matar Bir es- Safadi Meitar Mitcham C Nevatim Ramot 3 Ramot Nof Shiqmim Tel Arad Tel Sheva

128.50 128.60 143.83 130.90 139.50 131.65 131.70 115.00 162.00 135.30

71.40 70.70 82.20 72.50 70.00 76.00 76.50 67.00 75.00 73.00

A301 Ghazzeh A Ghazzeh B Ghazzeh D Ghazzeh E Ghazzeh M Ghazzeh O Grar R45 R48 R79

53.54 101.00 101.05 102.00 103.00 103.05 102.05 114.59 67.92 67.54 68.08

67.55 75.00 75.02 75.04 75.08 75.10 75.06 88.68 74.23 73.81 73.64

Teleilat Ghassul

208.00 135.00

Kh. Aliya East Beith Netofa Kaukab

146.00 123.00 170.00 210.00 174.45 248.45

PLATES

Plate 1. 1: core-chopper (Abu Matar), 2: blade core (Safadi)

129

PLATES

Plate 2. Beer-Sheva scrapers. 1: Abu Matar; 2: Safadi

130

PLATES

Plate 3. Safadi. 1-2: tabular scrapers 131

PLATES

Plate 4. 1: micro drill; 2: drill; 3: awl; 4-7: micro end-scrapers

132

PLATES

Plate 5. Abu Matar. 1-12: sickle-blades

133

PLATES

Plate 6. Safadi. 1-2, 5-7: piquant trièdre; 3-4, 8-12: microburins; 13-16: refitted piquant trièdre and microburins

134

PLATES

Plate 7. 1: axe; 2: adze

135

PLATES

Plate 8: 1: chisel; 2: perforated dics (drawing 2/3 of the original size)

136

PLATES

Plate 9. 1-3: picks (drawing 2/3 of original size)

137

PLATES

Plate 10. Tel-Sheva. 1-2: flakes knapped with Kombewa technique; 3: flake-core; 4: core-chopper 138

PLATES

Plate 11. Tel-Sheva. 1-5: bladelet cores found in one locus 139

PLATES

Plate 12. Tel-Sheva. 1-2 Discoidal cores; 3-4 Levallois-like flakes

140

PLATES

Plate 13. Tel-Sheva. 1-2: refitted flakes 141

PLATES

Plate 14. Tel-Sheva. 1-4: tabular scrapers 142

PLATES

Plate 15. Tel-Sheva. 1: drill; 2-6: awls 143

PLATES

Plate 16. Tel-Sheva. 1-10, 13: micrograttoirs; 11-12, 14-15: retouched bladelets

144

PLATES

Plate 17. Tel-Sheva. Pick

145

PLATES

Plate 18. Tel-Sheva. 1-2: adzes

146

PLATES

Plate 19. Tel-Sheva. 1-2: axes

147

PLATES

Plate 20. Mitcham C. 1: flake core; 2-3: bladelet cores; 4: blade core; 5: core-chopper

148

PLATES

Plate 21. Mitcham C. 1:awl; 2-4: scrapers

149

PLATES

Plate 22. Mitcham C. 1: microburin; 2, 7-9: retouched blades; 3, 5-6, 10: backed blades; 4: sickle-blade; 11: geometric sickle-blade

150

PLATES

Plate 23. Mitcham C. 1: varia

151

PLATES

Plate 24. Meitar. 1: flake core; 2: bladelet core; 3: sickle-blade; 4: micrograttoir; 5: scraper; 6: axe 152

PLATES

Plate 25. Nahal Komem. 1: flake/bladelet core; 2: bladelet core; 3: scraper 153

PLATES

Plate 26. Nahal Komem. 1-2: tabular scrapers

154

PLATES

Plate 27. Nahal Komem. 1, 3: canaanean retouched blades; 2: sickle-blade on canaanean blade; 4-6: sickle-blades; 7: micrograttoir; 8: retouched bladelet 155

PLATES

Plate 28. Kaukab. 1: discoidal core; 2: bladelet core; 3: flake/blade/bladelet core; 4: flake core

156

PLATES

Plate 29. Kaukab. 1-2: scrapers; 3: tabular scraper; 4: drill; 5: awl; 6: chisel

157

PLATES

Plate 30. Kaukab. 1-3: sickle-blades; 4-5, 7: micrograttoirs; 6: retouched bladelet

158

PLATES

Plate 31. Kaukab. 1: chisel; 2: adze

159

PLATES

Plate 32. Tel Turmus. 1: discoidal core; 2: flake core; 3: blade core; 4-5: bladelet cores 160

PLATES

Plate 33. Tel Turmis. 1-2: tabular scrapers; 3-4: scrapers; 5: drill; 6: awl 161

PLATES

Plate 34. Tel Turmus. 1: retouched and backed blade; 2: backed blade; 3-5: backed and truncated blades

162

PLATES

Plate 35. Tel Turmus. 1, 5-8: sickle-blades; 2-4: sickle-blades on canaanean blades 163

PLATES

Plate 36. Tel Turmus. 1: chisel; 2: adze

164

PLATES

Plate 37. Tel Turmus. 1: adze; 2: perforated disc

165

PLATES

Plate 38. Tel Turmus. 1-2: pieces esquilles; 3: awl on sickle-blade; 4: scraper on perforated disc; 5: burin on truncation; 6: burin on broken bifacial 166

PLATES

Plate 39. Tel Turmus. 1: chisel; 2: adze 167

PLATES

Plate 40. Bir Sawane. 1: scraper; 2: tabular scraper; 3: retouched flake; 4-5: massive awls 168

PLATES

Plate 41. Wadi Wattir. 1: flake-core; 2-3: varia 169

PLATES

Plate 42: Wadi Wattir. 1: axe; 2: adze; 3: chisel 170

PLATES

Plate 43. Wadi Zalaka. 1, 3: tabular scrapers; 2: awl on tabular scraper; 4: massive denticulate; 5: scraper; 6: awl 171

PLATES

Plate 44. Wadi Sa’al. 1: scraper; 2, 5: tabular scrapers; 3: retouched blade; 4: macro-lunate; 6: awl; 7: denticulate

172

TABLES

173

TABLES

174

TABLES

175

TABLES

176

TABLES

177

TABLES

178

TABLES

179

TABLES

180

TABLES

181

TABLES

182

TABLES

183

TABLES

184

TABLES

185

TABLES

186

TABLES

187

TABLES

188

TABLES

189

TABLES

190

TABLES

191

TABLES

192

TABLES

193

TABLES

194

TABLES

195

TABLES

196

TABLES

197

FIGURES

Figure 1. Comparison of width distributions of blades and sickle-blades of Abu Matar – Safadi

Figure 2. Comparison of sizes of bladelets, micro-endscrapers and microliths from Tel-Sheva

198

FIGURES

Figure 3. Distribution of flint raw materials at Tel Turmus

Figure 4. Distribution of flint raw materials at Tel Turmus by main tool types

199

FIGURES

Figure 5. Size distribution of Tel Turmus cores and flakes

200

FIGURES

Figure 6. Tel Turmus: distribution of perforated discs’ diameter and central hole diameter

201

FIGURES

Figure 7. Visualization of main ideas of theoretical framework adopted in this work

202

FIGURES

Figure 8. Graphic representation of the interpretation scheme of work’s results

203

MAPS

Map 1a. Map of main climate and rain zones of southern Levant

204

MAPS

Map 1b. Map of main phytogeographic zones of southern Levant

205

MAPS

Map 2. Map showing location of sites analysed

206