<<

A techno-typological analysis of Tor al-Tareeq (WHS 1065): An Epipaleolithic site in west-central

Item Type text; Thesis-Reproduction (electronic)

Authors Stevens, Michelle Nanette, 1965-

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction or presentation (such as public display or performance) of protected items is prohibited except with permission of the author.

Download date 04/10/2021 03:05:54

Link to Item http://hdl.handle.net/10150/291374 INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI films the text directly fi'om the original or copy submitted. Thus, some thesis and dissertation copies are in typewriter face, while others may be from any type of computer printer.

The quality of this reproduction is dependent upon the quality of the copy submitted. Broken or indistinct print, colored or poor quality illustrations and photographs, print bleedthrough, substandard margins, and improper alignment can adversely afifect reproduction.

In the unlikely event that the author did not send UMI a complete manuscript and there are missing pages, these will be noted. Also, if unauthorized copyright material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning the original, begiiming at the upper left-hand comer and continuing from left to right in equal sections with smaU overlaps. Each original is also photographed in one exposure and is included in reduced form at the back of the book.

Photographs included in the original manuscript have been reproduced xerographically in this copy. Higher quality 6" x 9" black and white photographic prints are available for any photographs or illustrations appearing in this copy for an additional charge. Contact UMI directly to order. UMI A Bell & Howell Information Company 300 North Zeeb Road, Ann Arbor MI 48106-1346 USA 313/761^700 800/521-0600

A TECHNO-TYPOLOGICAL ANALYSIS OF TOR AL-TAREEQ (WHS 1065)

AN EPIPALEOLITHIC SITE IN WEST-CENTRAL JORDAN

by

Michelle Nanette Stevens

Copyright ® Michelle Nanette Stevens 1996

A Thesis Submitted to the Faculty of the

DEPARTMENT OF ANTHROPOLOGY

In Partial Fulfillment of the Requirements For the Degree of

MASTER OF ARTS

In the Graduate College

THE UNIVERSITY OF ARIZONA

19 9 5 UMl Number: 1381776

Copyright 1996 by- Stevens / Mxchelle Nanette

All rights reserved.

UMI Microform 1381776 Copyright 1996, by UMI Company. All rights reserved.

This microform edition is protected against unauthorized copying under Title 17, United States Code.

UMI 300 North Zeeb Road Ann Arbor, MI 48103 2

STATEMENT BY AUTHOR

This thesis has been submitted in partial fulfillment of requirements for an advanced degree at The University of Arizona and is deposited in the University Library to be made available to borrowers under the rules of the Library.

Brief quotations from this thesis are allowable without special permission, provided that accurate acknowledgment of source is made. Requests for permission for extended quotation from or reproduction of this manuscript in whole or in part may be granted by the copyright holder.

SIGNED:

APPROVAL BY THESIS DIRECTOR

This thesis has been approved on the date shown below:

m \. IM John W. Olsen Date Pjrofessor of Anthropology 3 ACKNOWLEDGEMENTS

I would like to thank several people who provided assistance and guidence during my preparation of this thesis. Geoff Clark graciously allowed me to use the Step C, WHS 1065 lithic collection and to participate in the 1993 field season of the Wadi Hasa Paleolithic Project (WHPP). Deborah Olszewski introduced me to the WHPP and was also very helpful during the early stages of my analyses. I would also like to thank Mike Neeley for providing me with an advanced copy of the site report and letting me into the ASU lithics laboratory. Steve Kuhn was especially helpful providing many insightful and critical comments on several earlier drafts of this thesis, especially on lithic and statistical analyses. John Olsen and Carol Kramer were helpful not only for their careful readings of this thesis but also for academic guidance during my graduate career. Barbara Mills graciously allowed me to use space in the lab and use some of her laboratory equipment, e.g., calipers and computers. She also provided statistical assistance. 4

TABLE OF CONTENTS

LIST OF FIGURES 5

LIST OF TABLES 6

ABSTRACT 7

CHAPTER 1: INTRODUCTION 10

History of Research 15

Cultural Sequence and Description 23

Conclusions 39

CHAPTER 2: TOR AL-TAREEQ AND PLEISTOCENE LAKE HASA 41

Paleoenvironment and Paleolandscape 41

Previous Research 4 9

Tor al-Tareeq (WHS 1065) - Excavation and Stratigraphy 52

Interpretations 60

CHAPTER 3: RESEARCH METHODOLOGY 63

Acquistion of Raw Material 64

Core Reduction 65

Manufacture, Use and Discard 67

Sampling Rationale 71

Analysis of Cores and Debitage 74

Analysis of Retouched Tools 77

CHAPTER 4: LITHIC ANALYSES 79

Debitage 79

Debitage Morphometries 94

Debitage Summary 98

Retouched Tools - Typology and Technology 101

Major Tool Classes - Typology 103

Major Tool Classes - Technology 105

Microliths 128

Microburin Indices 134 5

TABLE OF CONTENTS - ContAnxied

CHAPTER 5: DISSCUSSION AND CONCLUSIONS 139

Site Formation Processes 13 9

Intra-site Functional Variability 141

Intra-site Variability in Operational Sequences 143

Regional Comparisions of Operational Sequences 148

Conclusions 154

APPENDIX A: WHS 1065 DEBITAGE ANALYSIS CODING LIST 157

APPENDIX B: WHS 1065 TOOL AND CORE ANALYSIS CODING LIST 158

REFERENCES 163 6

LIST OF FIGURES

FIGURE 1.1, Major Eipaleolithic sites in the Levant 11-12

FIGURE 1.2, Schematic illustrations of common Epipaleolithic 27

FIGURE 2.1, Distribution of excavated sites in the Wadi Hasa drainage basin 42

FIGURE 2.2, Site map Tor al-Tareeq (WHS 1065) 53

FIGURE 2.3, The east profile of Steps B and C 56

FIGURE 4.1, Box plot of core weights in grams by groups 91

FIGURE 4.2, Box plot of the widths of unmodified and bladelet blanks by level 95

FIGURE 4.3, Box plot of the thicknesses of unmodified blade and bladelet blanks by level 99

FIGURE 4.4, Box plot of the lengths of unmodified blade and bladelet blanks by level 100

FIGURE 4.5, Box plot of flake tool widths by level 110

FIGURE 4.6, Box plot of flake tool thicknesses by level Ill

FIGURE 4.7, Box plot of flake tool lengths by level 112

FIGURE 4.8, Box plot of blade and bladelet cool widths by level .... 114

FIGURE 4.9, Box plot of blade and bladelet tool thicknesses by level 115

FIGURE 4.10, Histograms of blade and bladelet tool widths for levels C08N-C11N 118

FIGURE 4.11, Histograms of blade and bladelet tool widths for levels C12N-C15 119

FIGURE 4.12, Histograms of unmodified blade and bladelet blank widths for levels C08N-C11N 121

FIGURE 4.13, Histograms of unmodified blade and bladelet blank widths for levels C12N-C15 122 7

LIST OF TABLES

TABLE 1.1, Cultural sequences in the Levant ca. 20,000-10,000 BP .... 24

TABLE 2.1, Correlations of natural and arbitrary levels from Steps B and C, and Units B and C 55

TABLE 2.2, Radiometric dates from Tor al-Tareeq (WHS 1065) 58

TABLE 4.1, Percentages of completeness categories for flakes, blades and bladelets by level 80

TABLE 4.2, Percentages of medial and distal fragments classified as blades and bladelets, and debris by level 80

TABLE 4.3, Debitage and tool percentages by level 81

TABLE 4.4, Ratios and indices of various artifact classes by level .. 81

TABLE 4.5, Percentages of size categories for complete and proximal flakes, blades, and bladelets by level 84

TABLE 4.S, Percentages of debris size categories by level 84

TABLE 4.7, Percentages of cortex for flakes, blades, bladelets and debris by level 8 9

TABLE 4.8, Percentages of cortex for flakes, blades, bladelets by level 8 9

TABLE 4.9, Percentages of cortex on cores by level 89

TABLE 4.10, Summary statistics for core weights by groups 89

TABLE 4.11, Percentages of core types by groups 93

TABLE 4.12, Kolmogrorov-Smirnov two-sided probability test results for blade and bladelet blank widths 97

TABLE 4.13, Percentages of major cool classes by level 104

TABLE 4.14, Row percentages of blank type for major tool classes by level 107

TABLE 4.15, Pearson chi-square test for independence of blank type by level and blank type by group for four major tool classes and all retouched tools 107

TABLE 4.16, Summary statistics for flake tool widths and thicknesses by level 116

TABLE 4.17, Summary statistics for blade and bladelet tool widths by level 116

TABLE 4.18, Summary statistics for bladelet tool widths by level ... 116

TABLE 4.19, Kolmogrorov-Smirnov two-sample probability test for bladelet tool widths by level 116 8

LIST OF TABLES - Continued

TABLE 4.20, Summary statistics for blade and bladelet blank widths by level 124

TABLE 4.21, Distribution of dorsal flake scars on blade and bladelet tools by level 124

TABLE 4.22, Distribution of dorsal flake scars on flake tools by level 124

TABLE 4.23, Percentages of tool platform types by level 127

TABLE 4.24, Percentages of tool retouch types by level 127

TABLE 4.25, Percentages of microliths by level 130

TABLE 4.26, Frequencies of specific microlithic types by level 132

TABLE 4.27, Formulas for microburin indices 136

TABLE 4.28, Percentages of microburins by level 136

TABLE 4.29, Microburin indices by level 136 9

ABSTRACT

A techno-cypological analysis of the chipped stone assemblage from Tor al-Tareeq (WHS 1065), an Epipaleolithic site in Wadi Hasa, west-central

Jordan, suggests that significant typological and technological changes occurred during the occupation of this site. The lowest levels have reliable radiocarbon dates (ca. 17,000-16,000 BP) and are associated with very narrow, backed microliths, single platform bladelet and multi- platform flake and blade cores, and use of the microburin technique.

The overlying, undated levels are associated with wide, short, geometric microliths, bi- and multi-directional flake and blade cores, and absence of the microburin technique. These technological and typological changes, associated with decreased mobility and moister climatic conditions in the upper levels, were not synchronous. The trend towards the manufacture of wide bladelet tools occurred before significantly wider bladelet blanks were being manufactured. The techno-typological characteristics of these assemblages resemble roughly contemporary sites in the Azraq Basin, northeastern Jordan. 10

CHAPTER 1 : INTRODUCTION

Current researchers use microlithic techno-typological variability as the main criterion for distinguishing Epipaleolithic cultures in the

Levant (ca. 20,000/18,000-10,000 BP) (e.g., Bar-Yosef 1991; Byrd 1994;

Goring-Morris 1987, 1995; Henry 1989b). Although the identification of prehistoric cultures should rely on multiple artifact classes and lines of evidence, microlithic tools are abundant and considered to be the most temporally and spatially sensitive artifact class (Bar-Yosef 1981,

1991; Henry 1982, 1989b). However, researchers do not necessarily agree on whether the level of observed variability represents culture groups, subgroups within a single culture, or environmental adaptations of one or more culture groups.

In order to help clarify some of these issues, additional techno- typological studies and studies of operational sequences used to manufacture chipped stone assemblages (i.e., raw material acquisition, core reduction, manufacture, use and discard of lithic artifacts) at multi-component Epipaleolithic sites in a variety of environmental and geographic contexts are necessary. The assumption behind these studies is that techno-typological and operational sequence approaches to lithic analysis have the potential to reflect, in conjunction with other data, different prehistoric culture groups, subsistence strategies, and land use patterns. If lithic assemblages from a variety of environmental and geographic contexts are analyzed, these studies will help fill spatial and temporal gaps in the archaeological record and allow more plausible interpretations of culture variability and change to be postulated.

The present study is a techno-typological analysis of a sample of chipped stone from Tor al-Tareeq (WHS 1065), an early to middle

Epipaleolithic site situated in upper Wadi Hasa, west-central Jordan

(Figure 1.1). Earlier studies and in-field analyses of the lithic FIGURE 1.1: Major Epipaleolithic sites in the Levant. (1) El-Kowm (2) Yabrud III (3) Ksar Akil (4) Azraq Basin: Wadi Uwaynid 18, Uwaynid 14, 6 (5) Wadi Hasa: Tor al Tareeq (WHS 1065) (6) area: Beidha, Wadi Madamagh (7) Wadi Hisma/Wadi Judayid sites (8) Ain Mallaha (9) Hayonim (10) , El-Wad, Nahal Oren (11) Shiokbah (12) (13) Neuville's sites (1934, 1951) (14) Negev sites 12 l&yiy ^v 0

Palmyra -7

Beirut

Damascus Mediterranean

«A Druze

Tel Aviv

El-Jafr 13 assemblage from this site indicate that t^'pclcgical =md technological changes occurred in the early to middle Epipaleolithic deposits, i.e., natural levels 5 and 7, Step C (see Figures 2.2 and 2.3, and Teible 2.1)

(Clark et al. 1987, 1988; Donaldson 1986; Donaldson and Clark 198S). I wanted to study the technological and typological variability associated with this transition (i.e., from a tool assemblage dominated by nongeometric microliths cuid use of the microburin technique to a tool assemblage with high proportions of wide geometric microliths and an almost complete absence of the microburin technique) in order to better understand why these techno-typological changes occurred. Particularly,

I was interested in determining to what extent these technological and typological changes reflect only diachronic change or a combination of diachronic change, and changes in subsistence and land use strategies in response to paleoenvironmental change.

Also since techno-typological characteristics and operational sequences are frequently used to identify cultural variability in the

Epipaleolithic, determining the synchrony of technological and typological change should enable a better understanding of the nature of culture change and variability, as reflected in chipped stone assemblages. If different cultural groups are present in the northern, southern, eastern and western portions of the Levant as some suggest.

Tor al-Tareeq is geographically positioned such that the lithic assemblage from the site may be able inform on the applicability of these cultural groupings.

Several aspects of this site have been previously reported, e.g., preliminary lithic analyses (Clark et al. 1987, 1988; Donaldson 1986;

Donaldson and Clark 1986), surface site structure (Coinman et al.

1989:213-236), paleogeography and paleoenvironment (Schuldenrein and

Clark 1994), and a general site description (Neeley et al. 1995) . This 14

Study is a reanalysis of the eight lowest excavation levels in excavation unit "Step C" (see Figures 2.2 and 2.3, and Table 2.1).

Earlier studies (i.e., Donaldson 198S) relied on in-field analyses of chipped stone debitage and retouched pieces. In-field analyses were conducted by several crew members with varying degrees of lithic expertise. Even though additional analyses were conducted on the chipped stone tools after the excavation season had concluded, in-field analyses did not identify all the tools and special debitage classes like microburins, within the assemblage. Therefore, a reanalysis of the debitage and tool components from this excavation unit is called for.

This study supplements and complements previous research at this site by (1) identifying debitage and tool categories that may have been overlooked or misidentified during in-field analyses; and (2) re­ analyzing the techno-typological variability in the debitage and tool categories in the lowest levels of an excavation unit where previous analyses indicated typological and technological changes in the chipped stone assemblage occurred. These techno-typological changes may be associated with a transition from an earlier Kebaran component to a later Geometric Kebaran component (Clark et al. 1987, 1988; Coinman et al. 1989; Neeley et al. 1995). As little is currently known about this transition in west-central Jordan, this analysis will provide information on lithic techno-typological variability and the operational sequences used to manufacture chipped stone tools in west-central Jordan during these periods. Tor al-Tareeq is particularly important for this because several reliable chronometric dates have been obtained from the lowest excavation levels (Clark et al. 1987, 1988).

This opening chapter presents a general overview of Epipaleolithic research in the Levant. The general Levantine cultural sequence and the characteristic attributes used to define major cultural divisions are 15 discussed. Chapter 2 is a presentation of the paleogeography and paleoenvironment of the Levant and the upper Wadi Hasa in order to position Tor al-Tareeq (WHS 1065) and the Wadi Hasa in their regional and temporal contexts. Also, previous research at Tor al-Tareeq (WHS

1065) including stratigraphic interpretations and relationships, and lithic analyses will be discussed. In chapter 3, a methodological discussion including the sampling rationale, and typological and technological approaches to this analysis is presented. Chapter 4 presents this study's analyses and interpretations of the debitage and tool components of the chipped stone. Finally, this paper concludes with a comparative discussion and interpretation of this site, and the implications of this research at local and regional scales.

History of Research

The Epipaleolithic period in the Levant dates between ca.

20,000/18,000 and ca. 10,000 BP. The Levant is defined here as the area encompassing the modem political states of Israel and Jordan. This region is divided into western and eastern areas by the Wadi Araba and

Jordan River, both of which lie in the Rift Valley. Although some include the modem political regions of and Syria in definitions of the Levant (e.g., Goring-Morris 1995), there is considerably less research on the Epipaleolithic in these areas. Therefore, the discussion of the Levantine Epipaleolithic presented here will not focus on these northern regions.

Prior to 1950, research on the Epipaleolithic in the Levant was dominated by a culture-historical approach in which researchers were primarily interested in establishing chronological sequences through artifact seriations. This early phase of field work was concentrated at cave sites in the Mediterranean vegetation zone where deep cultural deposits were thought to be present. Many of these sites such as Kebara 16

Cave (Garrod 1954; Turville-Petre 1932), al-Watwat, an Nugtah, Wadi

Fallah, and Abu-Usba Caves (Stekelis and Haas 1942, 1952), and Ksar Akil in Lebanon (Ewing 1949) were located in coastal areas (Figure 1.1).

However, Shukbah Cave in the Carmel Mountains (Garrod 1942) and several sites in the Judean Hills (Neuville 1934, 1951) were located further inland, but still far to the east of the present study area.

The transitional nature of the Epipaleolithic was recognized early on by who first identified the at Wadi el-Natuf (Garrod 1932). Influenced by western European concepts and general cultural sequences, Garrod used the term "" for the transition between the Upper Palaeolithic and periods (1932,

1937; c.f. Lubbock 1865). She recognized the period by the appearance of backed and retouched microliths. Neuville (1951) who was working at

Kebaran sites termed the period Epipaleolithic. The term Mesolithic, borrowed from western European archaeologists, was used to denote a closer association with the Neolithic, while the term Epipaleolithic reflected a North African bias and implied closer affinities with the local Levantine sequence (Phase IV) (Neuville 1951).

Since Levantine microlithic assemblages did not have any clear association either chronologically or culturally with the western

European sequence but did exhibit similarities with microlithic assemblages identified in North Africa (Perrot 1966; Tixier 1963), the designation Epipaleolithic eventually replaced the term Mesolithic.

Garrod and Neuville, the most prominent researchers at that time, were relatively progressive in their methodological and analytical approaches when compared to their contemporaries. They employed the standard archaeological approaches of artifact seriation and stratigraphy to construct regional culture histories. However lonlike other excavators, they supplemented these approaches with stratigraphic. 17 geologic, and faunal studies in order to reconstrijct climatological and

environmental sequences (Henry 1989b:7). In one study, the relative

proportions of Persian fallow deer {Dawa mesopotamica) and gazelle were

used to reconstruct past climatic conditions (Garrod and Bate 1937).

Although these studies were innovative, they focused on cultural and

environmental sequencing. Little attempt was made to apply their

climatic and environmental data to test or to explain any cultural or

environmental theory. Although Childe's (1939) hypothesis for the origins of agriculture was published shortly after most of these early reports (Henry 1989b:7), none of the researchers reevaluated their data to support or refute Childe's hypothesis. Although data on this early

research had already been published, Childe did not incorporate the

climatological and environmental sequencing data collected by Garrod and

Neuville into his hypothesis (Henry 1989b;7). Thus, there were two

parallel tracks of Epipaleolithic research, a theoretical approach

focusing on the origins of agriculture and a field oriented culture-

historical approach.

The culture-historical approach emphasizing data collection continued during the 1950s to early 1960s but new geographical and environmental situations were explored. In the western Levant, several important open air sites dating to the late Epipaleolithic to early

Neolithic were investigated. They include Ain Mallaha (Perrot 1962,

1966), Jericho (Kenyon 1959) , and Nahal Oren (Stekelis and Yizraeli

1963). Only Perrot (1962), however, attempted to explain the social and economic significance of his data (Henry 1989b:7). In the eastern

Levant, the first field work was conducted with Diana Kirkbride's survey and test excavations at Wadi Madamagh (Kirkbride 1958) and Beidha

(Kirkbride 1966) in southern Jordan. Again a late Epipaleolithic bias is present. The geographic separation of Kirkbride's sites from those 18 investigated, in the western Levant limited comparisons betv/een these tv/c geographic areas to general stratigraphic and artifact seriation correlations (Henry 1989b).

As chronometric dating techniques were not yet well developed, culture-historical approaches were limited to the relative dating techniques of artifact seriation and stratigraphy. Out of necessity, retouched and backed microliths became the "index fossils" for the entire Epipaleolithic. However, this approach can be somewhat circular.

Since microliths were believed to be only associated with the

Epipaleolithic, all sites with microliths were automatically assigned an

Epipaleolithic date, regardless of other artifact classes present.

Furthermore, the microlithic component of some assemblages may not have been recognized, due to poor recovery techniques. This would be especially true if microliths comprised only a small proportion of the chipped stone assemblage.

After the development of the radiocarbon dating technique, it became apparent that the mere presence of bladelets and retouched and backed microliths was no longer adequate for distinguishing the

Epipaleolithic from earlier or later periods (Byrd 1994). Bladelet blanks and microliths are found at Upper Paleolithic sites dating as early as 30,000 BP (Bar-Yosef and Belfer-Cohen 1977; Bar-Yosef and

Phillips 1977; Byrd 1994:206; Gilead 1983, 1988; 1991:121-125; Phillips

1994). Also, several sites that chronometrically date to the

Epipaleolithic may have predominantly non-microlithic assemblages and industries and have very few retouched or backed microliths (Garrard et al. 1994; Garrard and Byrd 1992; Gilead 1991; Goring-Morris 1987) or contain retouched bladelets that resemble those found in Upper

Paleolithic, i.e., pre-20,000 BP, deposits (Byrd 1994:206). However, the presence, absence and typological characteristics of microliths were 19

only cfitsirici 3.vail5t]?ls to 2rsss5i2rch.s2rs st tinis.

still, even after chronometric dating techniques had been developed, researchers in the mid 1960s to early 1970s continued to rely heavily on index fossils. In fact, the sequence based on forms was further cemented by systematic lithic analyses which categorized tools based on a formal type list approach, both in the

Levant (e.g., Bar-Yosef 1970; Hours 1976; Perrot 1966) and North Africa

(e.g., Tixier 1963). Marked changes in research foci and methodologies became evident in the late 1970s as systematic regional surveys and test excavations expanded to geographic areas outside the Mediterranean vegetation zone, especially into the arid Negev and Sinai (Marks 1976,

1977, 1983 ; Phillips and Mintz 1977) . One goal of these projects was to study diachronic change in regional land use and settlement patterns taking into account regional geomorphology and its effect on the visibility of archaeological sites during a given time period.

Extensive geomorphological and environmental studies enabled the reconstruction of paleolandscapes from which less biased settlement patterns could be derived. One major outcome of this research was Marks and Friedel's (1977) model of radiating versus circulating settlement patterns. This model is based on settlement distribution data. Marks and Friedel (1977) suggest that a semi-sedentary radiating settlement pattern characterized the Negev during moist periods, i.e., the Middle

Paleolithic and Natufian; while a circulating pattern involving a higher degree of mobility was more typical during drier periods, i.e., the

Upper Paleolithic and early Epipaleolithic.

Multi-disciplinary approaches to Paleolithic research were also conducted in the eastern Levant, e.g., the Wadi Hisma/Wadi Judayid area in southern Jordan (Henry 1982, 1983, 1987; Henry and Garrard 1988;

Henry et al. 1983), the Upper Wadi Hasa in west-central Jordan (Clark et 20 al. 1987, 1988; MacDonald et al. 1980, 1982, 1983; MacDonald 19SS), the

Azraq Basin in northeastern Jordan (Garrard et al. 1986, 1987, 1988), and the Palmyra and El-Kowm Basins (Cauvin 1981). The impetus of several of these studies was to test Marks' model of radiating versus circulating settlement patterns and to determine if such land use strategies were used in different paleoenvironmental contexts. If these strategies were used in different paleoenvironmental situations, researchers were interested in determining to what degree settlement

patterns were influenced by changing climatic conditions. Extensive surveys and excavations were also conducted in the western Levant with several projects associated either directly or indirectly with the emergency surveys of the Negev and Sinai (e.g., Goring-Morris 1987).

However as the number of microlithic and non-microlithic

assemblages with chronometric dates increased, it became apparent that

documenting regional and temporal typological variability in lithic

assemblages was no longer sufficient for addressing the new research

questions. To more adequately study regional land use strategies and settlement patterns, more technological approaches to lithic analyses

were required (Henry 1989b). Such technological studies focused on systematically recording the morphometries and types of blanks (e.g.,

Henry 1973; Marks 1983; Olszewski 1989). In addition, some analyzed specific blank attributes such as dorsal flake scar patterns, overall blank shape, and blank curvatures (e.g., Henry 1973, 1977).

This shift in lithic analysis from a typological to a techno- typological approach reflected American trends in lithic research.

These trends were in part responses to the debate between Hordes,

Binford, and others regarding the nature of lithic variability in Middle

Paleolithic assemblages of western Europe (Binford and Binford 1966;

Binford 1973, 1983; Bordes 1961, 1973, 1978; Dibble 1984, 1985, 1987; 21

Mellars 1969; Holland 1977, 1981) . Due to the reductive nature of lithic technology, if chipped stone typological variability was influenced by technological choices made during manufacture and reuse, any stylistic properties manufactured on the original artifact may have been altered during use and maintenance (i.e., resharpening) activities

(Neeley and Barton 1994). In addition, other factors such as the quality and quantity of raw material, and expediency may have more influence on artifact foms than stylistic conventions. Additional research emphasizing lithic reduction sequences and technology is necessary if the meaning of formal variability in artifact forms is to be better understood. Techno-typological variability in lithic manufacture can then be related to regional land use and subsistence strategies, settlement patterns, and economic and social interactions of

Epipaleolithic groups. These factors provide the building blocks in understanding changes in social and economic strategies between a mobile hunter-gatherer lifestyle in the Upper Paleolithic to a sedentary agricultural lifestyle in the Neolithic.

In an effort to explain both the typological and technological variability in Epipaleolithic artifact assemblages, numerous taxonomic classification systems have been developed reflecting a culture- historical or a time-stratigraphic approach (e.g., Aurenche et al. 1981;

Besancon et al. 1975-7; Byrd 1994:206; Garrard et al. 1994; Goring-

Morris 1995:141; Moore 1985). The most widely accepted culture- historical terms are Kebaran, Geometric Kebaran, Harifian, Mushabian, and Natufian (Bar-Yosef l991a; Henry 1989b; Valla 1988a) which some refer to as technocomplexes or cultural complexes (Henry 1989b). Still, some researchers are reluctant to apply even these general terms to both the western and eastern Levant (e.g., Byrd 1994). However, some sort of broad Levantine taxonomic scheme is useful in organizing the variability 22

in material culture, especially in the late Bpipaleolithic when there

seems to be increasing regional variability in lithic assemblages.

Much of the variation in nomenclature reflects the interpretive

levels at which these taxonomic schemes are applied. Classification of

material culture requires selection of arbitrary categories that reflect

varying degrees of technological and typological similarities with other

artifact assemblages (Henry 1989b). Henry's {1989b) hierarchical

classification scheme for scales of material culture during the

Bpipaleolithic recognizes four levels: complex, industry, phase or

facies, and assemblage. This hierarchical scheme is based on Clarke

(1968, 1979) but differs from Clarke's in that two independent

classification schemes are used, one for material culture and another

for prehistoric socio-economic data (Henry I989b:80). This modification

enables material culture to be compared without involving assumptions

regarding the behavior that produced the artifacts (Henry 198 9b).

Socio-economic levels (i.e., culture, culture groups and technocomplex)

are based on environmental, economic, demographic, spatial and temporal

evidence, and previously identified material culture units (Henry

1989b:81). For the Bpipaleolithic, comparisons of material culture are

based on quantitative analyses of chipped stone tools, tool blank

morphology, and microburin indexes.

In the hierarchy of material culture, artifacts are grouped into an assemblage, the smallest building block, which consists of artifacts

in a cultural deposit believed to be deposited over a single interval,

examples include an occupational horizon at a stratified site or debris

from a limited activity site. Assemblages are combined to form phases or facies. Phases have a very high level of technological and typological similarity. Typological similarity includes those

attributes related to form or style and function. Stylistic attributes 23 should be unrelated to fxinction, and include types of retouch (Henry

1973, 1977, 1989b:83), the position and orientation of retouch on certain tools (Close 1978), and specific uses of the microburin technique (Henry 1989b:83; Marks and Simmons 1977). Industries are groups of assemblages with very similar "specialized" technological attributes. The microburin technique is one example of a "specialized" technological attribute. The frequency aind types of major artifact classes in an assemblage can also be used to reflect core reduction strategies and other technological choices that can be used to determine the degree of similarity between assemblages. An "industry" should include specific quantitative boundaries, e.g., bladelet width as well as frequency ranges (Henry 1989b:83). Industries are grouped together to form a "complex" based on the rsinge of percentages of blank production and tool manufacture (Henry 1989b). A complex, the most general grouping of material culture, exhibits a "high level of technological affinity and a comparatively low level of typological affinity" (Henry 1989b:82).

Cultural Sequence and Description

While there is consideraible continuity in the techno-typologies of lithic industries throughout the early Epipaleolithic (ca.

20,000/18,000-15,000 BP), some variability in the microlithic assemblages has been noted. At the largest scale of complex, four, early to middle Epipaleolithic complexes are recognized: Kebaran,

Qalkhan, Geometric Kebaran, and Mushabicui. Several industries have been suggested within each of these complexes that reflect regional and temporal variability (Table 1.1). The Kebaran and Qalkhan Complexes are roughly contemporaneous and appear to be geographically separate. The

Kebaran is found predominately in the Mediterranean vegetation zone in the western Levsint along coastal areas and in upland environments. This 24

Table 1.1: Cultural sequences in the Levant ca. 20,000-10,000 BP. Industry level designations are denoted by complex/industry, e.g., Kebaran/Early Hanursui. Levcintine regional variants (industries) are distinguished by vegetation zones and indicated by (M) - Mediterranecin, (S) - Southern Irano- Turanian and Saharo-Arabicin, (E) - Eastern Irano-Turanian and Saharo-Arabioui, and (EN) - Eastern eUid Northern Irano- Turanian and Saharo-Arsibian.

KYA Climate Western Eastern Levantine Complexes Complexes Complexes (Byrd 1994) (Byrd (Henry (Goring- 1994) 1995) Morris 1995) 10- Humid Harifan Natuf. Late Final Natuf (M) warm eind Natuf. Harifian (S) related EciNatuf. (EN) 11- Dry Natufian indus- Late Natufian cold tries Early Natuf. Early |Ra- 12- Natuf. 1 monian Mush- 1 (S) abian/ 1 13- Humid Geom.-jMushabi- Madam. Geo. 1 warm etric 1 an Non- Geo. metric |Mush- Keb- i Micro- Keb./ Kebaran| abian 14- aran { lithic Hamran 1 (S) 11 Keb­ aran/ Late Kebaran 15- Kebaran E.Ham- ran Nizzanan 15- (Parts of M, S, 1 • • - - 1 and E) 1 17- Dry cold E. Kebaran (M) 1 1 Qalkhan (EN) 18- 1 > Nebekian (EN) 1 Non- Qalk- 19- i 1 Natufian hcui Micro- Masraqan i lithic 20- 1 Late/Teminal Late/Terminal Late Upper Drying Upper Upper Paleolithic >20 Paleolithic Paleolithic 25 distribution may in part reflect the relatively dry and cold conditions at that time. The Qalkhan Complex (Henry 1995) or the Non-Natufian

Microlithic (Byrd 1994) is foiand in eastern steppic regions, generally east of the Rift Valley. Although, it may be present in the northern

Levant as well (Henry 1995). The Geometric KebarcUi and Mushabian are roughly contemporaneous complexes. Both are found in the western and eastern Levant eind appear to succeed either the Kebarsui or Qalkhan

Complexes depending on the geographic region in which they are foxond.

Brief descriptions of the lithic assemblages, temporal range, geographical distribution, and some regional industry level variants of the Late/Terminal Upper Paleolithic Kebaran, Qalkhan, Geometric,

Mushabian Complexes are presented in the following section.

The Late/Terminal Upper Paleolithic (pre-20,000-ca. 16,000 BP) assemblages encompass considerable lithic variability but are represented in relatively few (six) archaeological horizons with reliable chronometric dates (Byrd 1994:208). Most researchers accept a terminal date of 20,000 BP for the Late/Terminal Upper Paleolithic period (Table 1.1). However in the western Levant, similarities in lithic tool traditions between Late/Terminal Upper Paleolithic assemblages and later sites that chronometrically overlap with Kebaran sites, suggests to some that the Late/Terminal Upper Paleolithic tool tradition may have continued for another 4,000 years in the western

Levant (Byrd 1994:208). Generally, these assemblages are characterized by the presence of some microliths (between 10% and 60% of the total tool assemblage). The microliths differ typologically from those found in contemporaneous Kebaran sites. Both twisted bladelets typically with interior lateral retouch, and finely retouched microliths are common, but generally not in the same assemblage (Byrd 1994:208). Core reduc­ tion technologies also seem to differ between Kebaran and Late/Terminal 26

Upper Paleolithic assemblages (Byrd 1994; Goring-Morris 1995).

The early Epipaleolithic Kebaran Complex (ca. 19,000-14,500 BP) was discovered by Turville-Petre in 1931 during his excavations in

Kebara Cave, Mt. Carmel (Turville-Petre 1932), and was later described by Garrod. In general, identification of the Kebaran is based on the high frequency of microliths and the predominaince of certain types of non-geometric microliths (Bar-Yosef 1975, 1981; see Figure 1.2 for illustrations of some common Epipaleolithic microliths). Also, the microburin technique was infrequently used to segment microliths prior to backing.

Various subdivisions of the Kebaran have been suggested. Bar-

Yosef (1970) postulated a scheme that subdivided the Kebaran into four contemporaneous culture groups. He has subsequently revised his earlier groupings, incorporating results from Lebanon (Hours 1976) and more recent investigations elsewhere (Bar-Yosef and Vogel 1987; Goring-Morris

1995:153). Group A is represented by narrow micropoints with basal modifications or truncations smd broad micropoints. As it is found in the southern coastal plain, it probably represents a regional variant.

Some suggest that coastal Kebaran sites, many of which are not well dated, may actually date to the Upper Paleolithic (Byrd 1994:209; Gilead

1991). The wide distribution of the other three groups B, C, and D suggests that these groups are not regional variants. Group B is represented by curved, backed, and pointed retouched bladelets; some of which have basal truncations. Group C contains a combination of narrow micropoints and obliquely truncated and backed bladelets. Group D has obliquely truncated backed bladelets (Kebara points) and narrow, curved backed bladelets. Bar-Yosef now recognizes that these groupings represent both temporal and spatial lithic variability in Kebaran assemblages. 27

Figure 1.2: Schematic illustrations of common Epipaleolithic microliths (after Goring-Morris 1987; Henry 1989; Muheisen 1988; Tixier 1963): (a) curved/arched ijacked bladelet, (b) obliquely truncated and backed, (c) scalene triangle, (d) trapeze/rectangle, (e) wide trapeze/rectangles, (f) microgravette, (g)regular microburin, (h) piquant tiedre microburin, (i) Krukowski microburin, (j) lunate, (k) Qalkhan point, (1) la Mouillah point. 28

The Kebaran has also been divided into early and late phases based on the frequency of geometric microliths. In general, micropoints, curved or arched backed bladelets, and microgravettes are more common in the early Kebaran and tend to stratigraphically precede Kebara points

(large, obliquely tnoncated, backed bladelets). Therefore, in Bar-

Yosef's original scheme (1970), Group C is considered earlier than Group

B. However, some sites that are dominated by curved micropoints have relatively late dates which suggests to some that early Kebaran dates should be extended upwards (Byrd 1994:209; Hovers and Marder 1991).

Although the late Kebaran is dominated by obliquely tr\incated and backed forms, curved and arched backed microliths still persist in low numbers in mainy assemblages (Bar-Yosef and Vogel 1987:225) . A technological trend is also recognized in a change from fine and inverse retouch to abrupt and bipolar retouch on the microliths. It has also been suggested that the early Kebaran has greater lithic variability than the late Kebaran {Bar-Yosef 1981:392-393, 1990). This greater lithic variability may account for most of the spatial subdivisions of the Kebaran (Byrd 1994:208).

In general, Kebaran sites are small (usually ca. 25-100 m^ and rarely larger than 250 m^) and are located in the Mediterranean vegetation zone (Goring-Morris 1995:153). During this period, climatic conditions are considered to be drier and colder than those at present based on low arboreal pollen frequencies in pollen diagrams throughout the Levant (Baruch 1994; Bottema 1987) . In addition, the early

Epipaleolithic and Kebaran Complex are coincident with the last glacial maximum of the Pleistocene, ca. 22,000-18,000 BP. As vegetation zones would have shifted north, southern and eastern Levantine areas would have been relatively drier and colder, as the Sahara-Arabian vegetation zone expanded northward. The steppic, Irano-Turanian zone would have 29 also moved north and may have been somewhat restricted. The

Mediterranean vegetation zone would also have been restricted, being found only along coastal areas, the Jordan Valley, and the west-central and northern Levantine areas.

Essentially no horizontal data exist at Kebaran sites; therefore, intra-site patterning of features and artifacts is not well studied

(Goring-Morris 1995:153). Unlike later Epipaleolithic complexes, few skeletal remains have been recovered from Kebaran sites. Based on the presence of bone that appeared to be cremated at Kebara C, it is suggested that the lack of skeletal remains may be related to cremation mortuary practices (Goring-Morris 1995:153). Most sites contain some marine mollusks, but, generally in low frequencies. Bone artifacts are also infrequent, probably due to poor preservation conditions, and include awls, points, spatulas, and very occasionally, an art object made on bone (Goring-Morris 1995:153).

Although most researchers use the term Kebaran to refer to early

Epipaleolithic assemblages in the western Levemt, some (e.g., Ferring

1977, 1988; Goring-Morris 1987, 1995; Marks 1976) believe there is sufficient lithic variability in the earliest Epipaleolithic to warrant its separation from the Kebaran into the Masraqan industry. The

Masraqan industry was first described at Masaraq an-Na'aj in the Judean

Desert (Perrot 1955). It is found throughout the Levant including the

Mediterranean, Irano-Turanian, and Saharo Arabian vegetation zones. As indicated in Table 1.1, this industry is coeval with some Late/Terminal

Upper Paleolithic, early Kebaran, Qalkhan or Non-Natufian Microlithic

Complexes. This industry has been subsumed imder the early Kebaran

Complex in the west, and the Qalkhan or Non-Natufian Microlithic

Complexes in the east.

Generally, Masraqan sites include small sites (ca. 25-250 m^) that 30 have a single hearth and an associated artifact scatter, and larger sites that have several hearths with each hearth having its own artifact scatter (Goring-Morris 1995:151). Occasionally, hearths are outlined by stones. Some vegetal remains have been recovered from at least one

Masraqan site, i.e., Ohalo II {Goring-Morris 1995:151). Gronndstone and bone points and awls are occasionally present at these sites. Marine shells, usually from the Mediterrsinean, are also present, sometimes in relatively high frec[uencies .

Masraqan lithic assemblages are characterized by elongated, narrow single platform cores, and evidence of intensive preparatory abrasion, resulting in carination (Goring-Morris 1995:151). The core technology used to manufacture larger blade blanks appears to differ from that used to manufacture bladelets. Bladelets are narrow, elongated and thin, with incurvate, not twisted, profiles (Goring-Morris 1995:151). Retouch includes Ouchtata, semi-abrupt, and abrupt types usually along the length of the blade; the distal and proximal ends of the blade are commonly not modified (Goring-Morris 1995:151). The microburin technique is absent.

In the eastern Levant, more variability in classification schemes exist (Table l.l). While some researchers are very clear about the hierarchial organization of complexes, industries, and phases (e.g.,

Henry 1989b, 1995), others are not and continue to split the early

Epipaleolithic into numerous industries without clearly identifying the relationship between them (e.g., Goring-Morris 1995). Given the plethora of extent classification schemes, some seem reluctant to jump into the fray opting to use the category Non-Natufian Microlithic until more studies and radiocarbon dates are generated from the eastern Levant

(e.g., Byrd 1994). Clearly more research is needed before the sequence and nature of interaction between groups during the early Epipaleolithic 31 can be hypothesized. It seems that the identification of numerous

industries has less utility, tending to obscure techno-typological

changes behind nomenclature, than demonstrations of technological

affinities between industries. This is especially true if these

industries are not particularly well dated.

The Qalkhan Complex (ca. 20,000-15,500 BP) seems to be a very

early Epipaleolithic complex that was first recognized in the Ras en-

Naqb/Hisma area of southern Jordan (Henry 1982, 1983, 1989b, 1995).

This complex appears to be an arid Icind adaptation parallel to the

Kebaran Complex in the Mediterranean woodlands. Few chronometric dates are associated with this complex, especially in southern Jordan.

However, this complex has been identified at Petra in southern Jordan

(Schyle aJid Uerpmann 1988), the Azraq Basin in northeastern Jordan (Byrd

1988; Garrard and Byrd 1992; Garrard et al. 1985, 1986, 1987; Garrard

and Gebel 1988), layers 4-7 at Yabrud III rockshelter in western Syria

(Rust 1950), and perhaps even further north in the El-Kowm Basin of

northeastern Syria (Cauvin 1981; Cauvin and Coqueugniot 1990; Cauvin et

al. 1979) by Henry (1995:38). Based on stratigraphy and radiocarbon

assays from sites in the Azraq Basin, this complex seems to date between

20,000-15,500 BP. Although its geographic distribution seems to be

restricted to the eastern steppic regions, it may also be present at

Mahal Lavan lOlOS in the western Negev (Goring-Morris 1995:152).

Qalkhan sites have a bimodal distribution with small sites (ca.

50-200 m^) located in open-air and sheltered situations in southern

piedmont areas, and larger sites (ca. 1,200-1,400 m^) located in more

northern areas (Henry 1995:235). Although the general climate is characterized as being dry and cold, climatic conditions in the Levant

were moister than at present. In southern Jordan, Qalkhan sites contain only chipped stone artifacts; no features or other artifact classes have 32

been foiind in association with these sites (Henry 1995:38)• The lithic

assemblages contain relatively long narrow bladelets as well as many

relatively wide bladelet tools which are often wider than the normal

range of microlithic tools. Therefore, these wide bladelet tools may

also be described as blade tools. The microlithic assemblage has

Qalkhan points which are shouldered triangles manufactured by the

microburin technique. The microburin scar on Qalkhan points is

unretouched (Goring-Morris 1995:152). Large la Mouillah points, double

truncated bladelets, and relatively high proportions of notches axid

denticulates are frequently found in tool assemblages (Goring-Morris

1995:152). Importantly, this complex represents the first appearance

and consistent use of the microburin technique.

Byrd (1994:210) uses the term Non-Natufian Microlithic, rather

than Qalkhan Complex, to describe lithic variability in the eastern

Levant during the early Epipaleolithic. All eastern regions exhibit a

development from the production of bladelets manufactured into narrow,

backed and retouched microliths to more geometric, tnoncated and/or

backed microliths. This period is best represented in the Azraq Basin

where sites have abundant evidence for the use of the microburin

technique (Byrd 1988; Byrd and Garrard 1990; Garrard et al. 1986, 1987,

1988; Garrard et al. 1994; Henry 1988, 1989b; Muheisen 1988). Earlier

sites in the Azraq area have narrow, finely retouched and backed, and

arched backed bladelets; while younger sites have larger backed tools

with more la Mouillah points and double truncated pieces (Byrd 1988), and wide geometric, truncated and/or backed microliths (Muheisen 1983,

1988) .

Variability and lack of reliable chronometric dates across the eastern Levant make Byrd (1994) reluctant to group these assemblages into a single complex. For example, in the Jordan Valley, narrow, 33 straight backed and obliquely truncated, non-geometric micrcliths dominate some early assemblages (Edwards 1987, 1990; Edwards et al.

1988). However, other assemblages are dominated by unbacked, obliquely truncated, non-geometric microliths (Bar-Yosef 1970). In these later assemblages, straight backed non-geometric microliths occur in very low frequencies. Tabaqat el-Bumma in west-central Jordan has four internally inconsistent dates and an assemblage that contains narrow, obliquely truncated and backed bladelets, and some micropoints (Banning et al. 1992) . These assemblages did not have evidence for the use of the microburin technique (Bcinning et al. 1992; Edwards 1987, 1990) and appear to be associated with the earlier Epipaleolithic based on typological considerations. Since these assemblages are more similar to the Kebaran or Late Upper Paleolithic Complexes of the western Levant, they may be contemporaneous with the western Levantine Kebaran and Late

Upper Paleolithic Complexes (Byrd 1994:211) and some would include them in Masraqan industry (Goring-Morris 1995).

Ironically, Byrd (1994) also views several southern Jordanian sites, the same sites used to define the Qalkhan Complex and industry, as problematic because they have few reliable chronometric dates and because of typological differences between these southern sites and northern sites in the eastern Levant. However, several important similarities between the lithic assemblages, particularly the early use of the microburin technique, suggest that these assemblages should be grouped together into a larger technological complex.

Several industries that appear to have strong affinities and geographical overlap with the Qalkhan Complex as defined by Henry

(1995), may be included in this complex such as the Nebekian, Qalkhan and Nizzian. The Nebekian industry (ca. 20,000-18,000 BP) was originally described by Rust (1950) . Geographically, it is found east 34 of the Rift Valley. Despite the fact that asseinblages assigned to this industry have not been well described, Goring-Morris (1995:152) distinguishes this industry from others based on its high frequencies of elongated, narrow and symmetrically curved, pointed arch-backed pieces which are almost oblique tnancations. Retouch is abrupt and invasive

(Goring-Morris 1995:152). The larger tool assemblages are characterized by truncated pieces and non-standardized retouched pieces (Goring-Morris

1995:152). Although several assemblages in northeast Jordan have affinities with the Nebekian (Garrard et al. 1994), Goring-Morris

(1995:152) suggests that several southern Jordanian sites may also be related, i.e., Wadi Madamagh levels A1-A2, Tor Hamar C, and J431.

However, Henry (1995) assigns these later sites, and many of the northern Jordan sites that define the Nebekian, into the Qalkhan industry.

The Qalkhan industry as described by Goring-Morris (1995) includes most of the characteristics used to define the Qalkhan Complex. The criteria for distinguishing the Nebekian from the Qalkhan industry appear to be that the Qalkhan industry has wide bladelet tools, Qalkhan points with a retouched microburin scar, large la Mouillah points, double tr\incated bladelets, and abundant notches and denticulates. It is also chronologically later than the Nebekian (Goring-Morris

1995:152). Henry (1995), however, describes the Qalkhan industry essentially as he does the Qalkhain Complex, using identical criteria and an expanded geographical distribution that includes arid regions along the eastern edge of the Mediterranean woodland zone from southern Jordan to northeastern Syria.

The Nizzian (ca. 17,000-15,000 BP) includes small scalene and isosceles triangles made by the microburin technique and microgravettes

('spiky points') (Goring-Morris 1995:152). This industry is frequently 35 svibsumed under the Geometric Kebaran Complex; yet. radiometric determinations indicate this industry overlaps with the end of the

Kebaran. Therefore, it may represent either a regional contemporaneous variant or a later phase of the Qalkhan (Goring-Morris 1995:155). The lithic assemblage contains exhausted cores, usually pyramidal, single platform cores with some opposed platform cores present. The tool assemblage also includes well-made scrapers and burins cind numerous dihedral burins (Goring-Morris 1995:155). Scalene bladelets and intensive use of the microburin technique are common. It has been suggested that this assemblage may be a precursor or predecessor to the

Mushabian (Goring-Morris 1995:155).

The middle Epipaleolithic (ca. 14,500-12,500 BP) includes the

Geometric Kebaran and Mushabian Complexes. Climatic conditions during this period changed siibstantially and are associated with the retreat of the glaciers after the last glacial maximum at the end of the

Pleistocene. In general, climatic conditions throughout the Levant became relatively warmer and moister than the previous period based on an increase in the amount of arboreal pollen in pollen diagrams across the Levant (Baruch 1994; Bottema 1987). The warmer and moister climatic conditions during this period favored the southward migration of the major vegetation zones such that the Mediterranean woodlands, and steppic environments would have expcind as the Saharo-Arabian zone contracted or was pushed further south. The geographical distribution of middle Epipaleolithic sites reflects these favorable climatic conditions with an increased number of sites found in desert regions.

The Geometric Kebaran Complex (14,500-13,000 BP) is characterized by straighter backed, non-geometric microliths than in earlier periods, and the appearance and widespread use of geometric, trapeze-rectangle microliths (Bar-Yosef and Belfer-Cohen 1989:462-453; Byrd 1994; Henry 36

1989b:93-94; Valla 1988b). Bladelet blanks tend to be slightly wider and longer than blanks fovind in the Mushabian Complex. In general, the microburin technique was only occasionally used to truncate microliths cind there is considerable variability in microburin indices. When the technique is used, it is more often associated with the manufacture of triangles than with trapezes and rectangles (Henry 1989b:93). The blanks for rectangular and trapeze microliths were apparently snapped without the microburin technique. Well dated Geometric Kebaran sites are found primarily in the Sinai and Negev (Byrd 1994:209).

As previously mentioned, settlement patterns during the Geometric

Kebaran differ from Kebaran patterns in that there is a noticeable increase in the number of Geometric Kebaran sites in the desert regions.

Geometric Kebaran sites are generally still found near major water sources including perennial springs, seasonal playa type settings or confluences of major drainages (Goring-Morris 1995:161). This strong association of Geometric Kebaran sites with major water sources, and ameliorating climatic conditions at this time suggest to some that the

Geometric Kebaran originally derived from an early Mediterranean woodland adaptation of the Kebaran (Goring-Morris 1995:161). Although one expects both prehistoric and modem groups to inhabit areas close to reliable water sources, the idea is that Geometric Kebaran sites do not represent a different cultural adaption to a more arid environment.

Rather, the ameliorated climate during this period would have expanded the Mediterranean woodland zone and Geometric Kebaran sites would have been located in environmental situations similar to those occupied during the Kebaran. Compared to early Epipaleolithic sites, there is decreased evidence for the use of marine shell, perhaps due to rising sea levels at this time (Goring-Morris 1995:161).

The Geometric Kebaran has been divided into early and late phases 37

based on widths of backed microliths. Narrcv/ r^icrcliths are generally

considered older (Bar-Yosef 1981:397-398); however, recent dates from

Neve David may not support this assumption (Byrd 1994:209). In southern

Jordan, a regional variant of the Geometric Kebaran Complex, the Hamran

industry, has been recognized (Henry 1989b, 1995). The Middle Hamran

industry is typical of the Geometric Kebaran Complex with relatively

narrow bladelets from which trapeze/rectangles were manufactured without

the use of the microburin technique. Sites are typically 320-650 and contain hearths, basalt pestles, and ornamental shells (Henry 1995:39).

The Late and Final Hamran differ from the Middle Hamran by having shorter bladelets, the microburin technique, and the appearance of

lunates (Henry 1995:39). The Late and Final Hamran are distinguished by

the relative frequencies of trapeze/rectangles and lunates. The Late

Hamran has more trapeze/rectangles than lunates, while the Final Hamran

has more lunates than trapeze/rectangles. Also, in the Final Hamran,

there is evidence for increased use of the microburin technique, and

some lunates have Helwan retouch. Based on the appearance of lunates

and Helwan retouch, and more substantial architectural features (i.e.,

stone-lined hearths), these Final Hamran sites show a clear

developmental trend towards the Natufian (Henry 1995:39).

The Mushabian (ca. 14,000-12,800 BP) was defined in the Gebel

Maghara, Northern Sinai and some other sites in the Negev (phillips and

Mintz 1977). It is characterized by abundant backed bladelets, mostly

arched backed bladelets, scalene bladelets, and la Mouillah points

manufactured with the microburin technique (Phillips and Mintz 1977) .

The prominent use of the microburin technique is seen in the Mushabian's

high microburin indices. Originally, the high microburin indices led

some to postulate that the origins of this complex were in North Africa

(Bar-Yosef and Vogel 1987; Phillips and Mintz 1977). Geometric 38 microliths are present but occur in IGV: frequencies and have variable forms (Henry 1989b:91). In general, the bladelet blanks are relatively- short and wide.

The settlement patterns and site locations of the Mushabian

Complex resemble those of the Geometric Kebaran in many respects.

However, it differs from Geometric Kebaran settlement patterns in that many sites do not appear to be as strongly tied to water sources as

Geometric Kebaran sites (Goring-Morris 1995:164). Mushabian sites are more frequently located in exposed areas, or areas with good views of the local landscape. Such locations may have been advantageous in monitoring the movement of game in steppic environments.

Many regional variants, i.e., industries, of the Mushabian Complex have been identified; the most prominent being the Ramonian (Goring-

Morris 1987, 1995), and the Madamaghan (Henry 1989b, 1995). The

Ramonian is distinguished from other Mushabian industries based on differences in raw material usage, single platform cores, pyramidal cores, and blades and bladelets that are frequently quite narrow and elongated. The Ramon point, a concave backed and obliquely truncated bladelet manufactured using the microburin technique, is present. The morphological dimensions of many Ramon points exhibit little variation which suggests that the morphology of these points may have been somewhat standardized. Settlement patterns differ from those found in preceding periods by an increase in the number of high elevation sites located outside the Mediterranean woodland zone in more arid environments. Most Ramonian sites were located in the Irano-Turanian vegetation zone and the lowland Saharo-Arabian dune localities in the

Negev. Since highland sites have generally smaller and more diverse chipped stone assemblages, it has been suggested that highland and lowland sites may represent a seasonal pattern of dispersion and 39 aggregation by groups whose Local ranges transected the different environmental zones. Ramonian sites frequently have larger tool assemblages and appear to have separate reduction sequences for larger tools such as scrapers and smaller tools such as microliths.

Henry (1989b, 1995) has identified the Madamaghan industry within the Mushabian complex. Technologically, this industry has relatively long, wide bladelets, and high microburin frequencies. Typologically, there are moderate percentages of points including la Mouillah, microgravette, and arched backed varieties, and low percentages of geometric microliths consisting of trapeze/rectangles, and normal and

Helwan lunates. Although the characteristics of this industry seem to exhibit good affinities with other Mushsibian sites, it is not clear that the type site for this industry, Wadi Madamagh, actually dates to the middle Epipaleolithic. Although Henry (1989b, 1995) asserts that the lithic assemblage from the rockshelter in Wadi Madamagh typologically resembles other assemblages in the Mushabian Complex, the original interpretation of the site (Kirkbride 1958) and later references to this site (e.g., Byrd 1994:205) indicate that Henry's assertion is not universally accepted.

Conclusions

There is considerable typological and technological variability in early and middle Epipaleolithic lithic assemblages. As a result, numerous classification schemes have been postulated that reflect techno-typological differences in lithic assemblages during this period.

Henry's (1989b) hierarchical scheme of complex, industry, phase, assemblage is being adopted more frequently. Still, a number of authors are reluctant to apply this scheme in both the eastern and western

Levant, or continue to identify numerous groups at only the industry level. All researchers agree that additional quantitative technological 40 and typological descriptions of lithic assemblages are necessary from sites with reliable chronometric dates before this situation can be satisfactorily resolved. The analyses presented here should be a step in this direction. 41

CHAPTER 2: TOR AL-TAREEO AND PLEISTOCENE LAKE HASA

Tor al-Tareeq (WHS 1065) is located on the north bank of a small tributary of Wadi Hasa in west-central Jordan (Figure 2.1). WHS 1065 is spread over a 812 m^ area and has cultural deposits with an average depth of about one meter. Originally, Tor al-Tareeq was interpreted as a series of superimposed Kebaran (and possibly Natufian) basecamps

(Clark et al. 1987, 1988; Coinman eC al. 1989). This interpretation

was based on the high density of artifacts in well-defined natural

levels; typological similarities between the WHS 1065 and Kebaran chipped stone assemblages; and several radiocarbon determinations from the lower excavated levels at WHS 1065 that cluster between 16,500-

15,000 BP (Clark et al. 1987, 1988; Coinman et al. 198 9; Neeley et al.

1995). Several features including hearths, pits, overridden wall sections, midden deposits under a collapsed rockshelter, and a fossil spring deposit were used to support this initial interpretation (Clark

et al. 1987, 1988; Coinman et al. 1989). Recently, a broader

interpretation has been suggested, i.e., that the site may have been

occupied repeatedly for relatively short periods and/or intensively at

several points in time (Neeley et al. 1995). The temporal and cultural

interpretation of the upper levels has also changed from a possible

Natufian basecamp to several Geometric Kebaran occupations (Neeley et

al. 1995).

Paleoenvironment and paleolandscape

The modem Levantine climate is characterized by long, hot, dry

summers that alternate with shorter, wetter winters. Annual

precipitation varies between 50-100 mm in the arid regions of the Negev,

Sinai, and southern Jordan to 200 mm in the northern Negev, Jordanian

Plateau, and lower Jordan Valley and to 200-300 mm in the upland and coastal regions of the northern Levant. Today, WHS 1065 is located at 42

Figure 2.1: Distribution of excavated sites in the Wadi Hasa drainage basin (after Neeley et al. 19951 43

815 m asl in the Irano-Turanian steppe environment, a setni-arid zone

that receives about 100-200 mm of precipitation annually. Modem

rainfall usually occurs between December and March. However as

precipitation is commonly torrential and localized, the annual amount

and intensity of precipitation varies considerably within and across the

region.

A description of the paleoenvironmental and paleoclimatic

conditions of the Levant during the Late Pleistocene (ca. 25,000-10,000

BP) is presented in more detail here than in the preceding chapter. An

understanding of paleoenvironmental variability and consistency across

the Levant is helpful in positioning the occupations at WHS 1065 in

their regional, environmental and temporal contexts. It should become

clear that although the timing may vary between regions, the

paleoenvironmental sequence for the eastern Hasa region is largely

consistent with other Levantine paleoenvironmental reconstructions.

Climatological conditions in the Late Pleistocene produced long-

term, multi-annual cycles that brought about fluctuations between

periods of drought and periods of higher humidity (Katsnelson 1964;

Rosenan 1970). While dry periods probably had somewhat higher humidity

levels than present, humid periods may have had as much as 50-100%

greater humidity (Issar and Bruin 1983:75), perhaps even more. As a

consequence, precipitation and recharge rates across the Levant were generally greater and aquifers tended to fill (Issar and Bruins

1983:63), so that, spring activity was probably greater throughout the

Late Pleistocene.

In the Late Pleistocene, WHS 1065 would have been located at the northwest edge of a large Pleistocene lake (Schuldenrein and Clark

1994). Several depositional changes at WHS 1065 are present such as the

interfingering of cultural materials with marshy sediments, the 44 deposition of cultural materials over marl deposits, and the movement of cultural materials by colluviura. The interfingering of palustrine (pond deposited) silts and clays, and colluvium indicate that the local environment changed as lake levels in Pleistocene Lake Hasa fluctuated between mesic and xeric conditions (Schuldenrein and Clark 1994) .

Sometime between 25,000 and 22,000 BP, climatic conditions became increasingly arid throughout the Levant. Supporting geomorphologic evidence includes: (1) the initial lowering of Lake Lisan between

21,000 and 18,000 BP (Gat and Magaritz 1980; Kaufman et al. 1992;

Yechieli et al. 1993); (2) the first appearance of eolian sand in drainages in the Sinai and Negev around 25,000 BP (Goldberg 1986:242);

(3) the absence of dated fluvial sediments between 22,000-17,000 BP indicating widespread valley erosion in the Negev (Marks 1977) ; and (4) an erosional period between 25,000-20,000 BP in the Judayid Basin-Wadi

Hisitia area of southern Jordan (Henry 1986).

Evidence for a drying trend, i.e., lower lake levels and reduced base levels, appears somewhat later (ca. 20,000 and 17,000 BP) in the eastern or Upper Hasa (Schuldenrein and Clark 1994:49). Between 26,000-

20,000 BP, Pleistocene Lake Hasa had high lake levels which are associated with increased spring activity, and more mesic conditions.

The Late Ahmarian occupation in Wadi Hasa (WHS 618) is associated with a variety of spring and marsh environments during this period. The environmental situation at WHS 1065 is directly associated with fluctuations in lake levels (Schuldenrein and Clark 1994:40). Between ca. 20,000-17,000 BP, the Hasa Marls, which were deposited in subaqueous conditions, eroded and stream margin environments around Lake Hasa formed (Copeland and Vita-Finzi 1978; Schuldenrein and Clark 1994:46;

Vita-Finzi 1966) . Evidence from WHS 1065 indicates that this interval is associated with lower lake levels, drier conditions and erosion of 45

laminar lake beds (Schuldenrein and Clark 1994). The lack cf riparian

taxa in the pollen spectrum also supports this interpretation of cooler

and drier conditions; this interpretation is consistent with regional

interpretations (Neeley et al. 1995). Still, conditions at the site

were probably at least seasonally marshy as these deposits are

associated with marl and palustrine silts and clays (i.e., marshy-

lacustrine silts and clays) (Clark et al. 1987, 1988; Schuldenrein and

Clark 1994:47). Similarly, the onset of drier conditions in northeastern Jordan occurs sometime after 21,000 BP, 2,000-3,000 years

later than in more southern regions based on evidence from Wadi Jilat 9

in Azrag Basin (Garrard et al. 1987).

Palynological evidence suggests that moist conditions actually

continued across the Levant until ca. 22,000 BP which is closer to the

Azraq situation (Horowitz 1979:341). However, subsequent palynological

studies from a well dated core in the Hula Basin (Bottema and Van Zeist

1981) suggest that the climatic reconstruction of Marks (1977) and

Goldberg (1981, 1986) in which there is a change to more arid conditions

ca. 25,000 BP, is more appropriate. Therefore, the "lag" in the onset

of drier conditions in the Azraq Basin and Wadi Rasa is most likely due

to its deeper placement in the northern climatological belt. As the

climatic belts migrated north during drier periods, the Azraq Basin and

Wadi Hasa would have been affected later than more southern regions.

The rate of environmental change during this period and its

potential affects on human and animal populations throughout the Levant

can be reconstructed to some extent from these data. As a result of drier climatic conditions and erosion. Late Upper Paleolithic and early

Epipaleolithic sites (ca. 22,000-18,000 BP) are scarce (Goldberg and

Bar-Yosef 1982; Henry 1986) especially in the southern Levant. If any archaeological material was deposited during this period, it would most 46 likely have been eroded. The few Late Upper Paleolithic sites that do exist in the Negev and southern Jordan are found associated with eolian sediments and dunes (Goldberg 1986:242; Henry 1986).

Between ca. 17,000-12,000 BP, a moist period is suggested by (1) renewed alluviation and aggradation in the northern Sinai aind western

Negev (Goldberg 1986:240); (2) new fan deltas at the junction of major tributaries; (3) local palustrine basins in the Hasa (Schuldenrein and

Clark 1994) and lower Jordan Valley (Schuldenrein and Goldberg 1981);

(4) high lake level stands at Lake Lisan ca. 14,000 BP (Begin et al.

1985; Druckman et al. 1987); (5) spring heads reaching their highest level at Wadi al Hammeh ca. 15,000-11,000 BP (Macumber and Head

1991:72); (6) higher water tables and an increase in relative moisture

from lake-like deposits in Wadi Mushabi and extensive gleying in Qadesh

Bamea associated with middle Epipaleolithic sites ca. 14,000 BP

(Goldberg 1981:60-21; 1986:240); (7) well developed paleosols in the

Negev highlands and the Gebel Maghara area of northwestern Sinai by

14,000 BP indicating moist, stable conditions must have prevailed in the

region for at least 3,000-4,000 years (Goldberg 1981); and (8) the

pollen diagram from the Hula pollen core (Baruch and Bottema 1991).

At Wadi Hasa, there is evidence for higher lake levels, an

increase in spring activity, and deposition of tributary alluvium on

eroded surfaces between ca. 17,000-15,000 BP (Schuldenrein and Clark

1994). This period corresponds to the earliest cultural deposits at WHS

1065 (some of which are analyzed in this study). In general, high lake

levels are associated with moister climates and transitions to paludal

(pond-like or marshy) environments. Pollen evidence supports this

interpretation showing a steady increase in the ratio of arboreal to

steppe plants (Neeley et al. 1995). Wadis Salibiya and Fazael in the

lower Jordan Valley show slightly earlier climatic amelioration (Darmon 47

1987; Leroi-Gourhan 3ind Darrr.cn 19S7) . However, the northern position of the lower Jordan Valley would naturally result in this area experiencing the effects of climatic amelioration earlier than more southern areas as climatological belts migrated southward.

How much moister was this "moist" period? Based on the well developed paleosols in the Negev highlands and the Gebel Maghara area of northwestern Sinai, climatic conditions may have been considerably wetter with 500 mm of annual precipitation (2.5 times greater than present) by 14,000 BP (Goldberg 1981). In southeastern Jordan, Judayid

504, a rockshelter overlooking a dry lake bed, lies in red sands with high frequencies of oak, walnut, and conifer pollen (Henry 1982). This is indicative of an environment with 200-300 mm of precipitation annually (again, 2-3 times greater than present) (Henry 1982). Based on the Negev data, rainfall gradients must have been depressed about 150 km to the south and by extension across the entire Levant and Sinai

(Goodfriend and Magaritz 1988) .

Not surprisingly, land use patterns also suggest more mesic climatic conditions as an increased number of late Kebaran and Geometric

Kebaran sites are found outside the Mediterranean woodland vegetation zone as reconstructed for the early Kebaran period. This suggests that the Mediterranean woodland zone extended further south and east than previously. Moister conditions continued in the northeastern Levant until at least ca. 14,500 BP in the Azraq Basin (Garrard et aJ. 1986).

Unfortunately, the geomorphic stratigraphy of the Judayid Basin-Wadi

Hisma area is not well developed and paleoenvironmental reconstructions rely heavily on palynological studies at a few locations. However, evidence from these locales suggests that moister conditions continued to approximately 12,000 BP (Henry 1986).

A drying trend with more arid conditions than the previous dry 48 period began about 11,000-12,000 BP and continued to about 10,000 BP.

It corresponds to the Geometric Kebaran in the Negev and Sinai and

Natufian across the Levant. Several locations provide sedimentological and palynological evidence for drier conditions at this time including

(1) initiation of the halite deposition in the Rift Valley between

11,000-8,500 BP (Yechieli et al. 1993); (2) a dramatic decrease in arboreal-nonarboreal ratios from 75% arboreal pollen in 11,500 BP to 30%

by 10,500 BP in the Hula pollen core (Baruch and Bottema 1991); (3) absence of lake deposition at Wadi al Hammeh ca. 11,000 BP in response to significant increases in evapo-transpiration rates (Macumber and Head

1991); (4) a highly evaporated sabkha deposit in the northern Negev ca.

11,000 BP (Magaritz 1986); (5) archaeological sites in eolian deposits

in the Azraq Basin defined at Kharaneh 4, Phase C and D (Garrard et al.

1987) and at Judayid 2 in Wadi Judayid (Henry 1986) ; and (6) massive cycles of colluviation and slope erosion in the lower Jordan Valley

(Schuldenrein and Goldberg 1981) and Wadi Hasa at WHS 1065 (Schuldenrein and Clark 1994). Land use patterns also suggest increased aridity as

Geometric Kebaran and Natufian sites in the southern Levant and Sinai

have limited distributions during this period.

In sum, during the end of the Pleistocene, major fluctuations in climatic conditions created corresponding fluctuations in depositional

and erosional sequences. Moist conditions prevailed before 25,000 BP

and are associated with Late Upper Paleolithic occupations in alluvial

sediments. This was followed by a significant drying trend between

25,000 and 17,000 BP and is associated with a decrease in the number of

Late Upper Paleolithic and Kebaran occupations found in most southern

Levantine areas. While the dry period affected the entire Levantine

region, its effects were more significant in the south where massive

erosion and deposition of eolian sediments occurred. A brief return to 49

moister conditions between 17,000 and 12,000 BP marks a period of human

expansion into previously arid regions and the development of paleosols

at a number of locations in the Negev aind Sinai. A drier period between

12,000 and 10,000 BP is indicated by the lowering of Lake Lisan, the

accumulation of evaporite deposits as shallow lakes dried, the incision

of most wadi systems across the Levant and Sinai, and the deposition of

eolian sands in more localized regions.

The timing of the climatic fluctuations and subsequent changes in geomorphic stratigraphies differs slightly between regions. In general, southern regions were more significantly affected by increasing aridity

and had longer arid periods than northern areas. During these drier

periods, human occupation of the region was restricted to springs,

marshes and lake margins especially in southern regions.

Previous Research

WHS 1065 was discovered in 1982 by Burton MacDonald during the

Wadi Hasa Survey (WHS) (MacDonald et al. 1980, 1982, 1983; MacDonald

1988). Between 1979 and 1983, the WHS conducted a fairly systematic

survey on the south bank of Wadi Hasa drainage system (Figure 2.1). The

Wadi Hasa drainage begins in central Jordan, near Qa el-Jinz and flows

west to its mouth at the Dead Sea near As-Safi. During the WHS, over

1074 sites were identified; 222 of these were lithic sites from the

Lower, Middle, and Upper Paleolithic, Epipaleolithic, and Prepottery

Neolithic periods (Clark et al. 1987, 1988; MacDonald et al. 1983;

MacDonald 1988). The number of Paleolithic sites identified during the

WHS suggested that additional work, focusing specifically on the

Paleolithic, had the potential to yield valuable information on land use

and settlement patterns of Paleolithic groups in this region.

Although the WHS recorded prehistoric sites, its research design

emphasized identifying historic period sites. Therefore, many 50 geomorphological, topographic and environmental situations adjacent to

Wadi Hasa were not necessarily explored (Clark ec al. 1987). As current and past geomorpological conditions can significantly affect the distribution, preservation and visibility of Paleolithic sites, these conditions need to be considered when interpreting the distribution of cultural and temporal periods in the WHS data. Although some geoarchaeological research relating landforms to archaeological sites in the Wadi Hasa had been conducted (Copeland and Vita-Finz 1978; Vita-

Finzi 1964, 1966, 1982), the dating of the landforms relied heavily on artifacts thought to be temporally diagnostic (Schuldenrein and Clark

1994:33). Only one radiocarbon determination from the lowest terrace in the present Wadi Hasa drainage system was obtained. The aggradation event associated with this lowest terrace (called the Hasa Formation) dated to ca. 8,000 BP (Copeland and Vita-Finzi 1978).

Additional research was clearly needed to (1) obtain radiocarbon dates to secure the cultural and environmental sequence in central

Jordan, (2) reconstruct land use and settlement patterns around Wadi

Hasa, and (3) evaluate the existing settlement pattern and land use models. Two lines of research that would supplement the WHS and previous geoarchaeological work were necessary including: (l) systematically surveying a variety of geomorphologic and environmental contexts around Wadi Hasa that had not yet been surveyed; and (2) defining and refining the environmental and cultural sequence for the

Paleolithic period in central Jordan.

In 1984, the Wadi Hasa Paleolithic Project (WHPP) was developed in order to gather data to address the additional research topics mentioned above. The goals of the WHPP were to "(1) acquire adequate samples of lithic assemblages, (2) recover faunal, floral and other kinds of paleoenviromental information, (3) undertake a geomorphological study of 51 the east Hasa drainage, (4) establish the beginnings of a radiocarbon chronology for west-central Jordan, and (5) map the extent of

Pleistocene Lake Hasa with which most of the archaeological sites are associated" (Clark et al. 1987:19). Once the environmental and cultural sequences for west-central Jordan were developed and settlement distribution data collected, these data could be used to evaluate two prominent settlement system models: Marks' model for paleoenviromental change and human adaptation (radiating verses circulating settlement systems) developed for the central Negev highlands (Marks and Friedel

1977; Mortensen 1972) and Henry's model of transhumance developed for southern Jordan (Henry 1982, 1983).

Marks' radiating settlement pattern has a sedentary to semi- sedentary residential camp surrounded by small, limited activity sites during relatively moist climatic periods (Marks cind Friedel 1977) .

These moist periods would correspond to expansions in the Mediterranean vegetation zone. A circulating settlement pattern consists of small, relatively similar multi-purpose sites. This pattern is associated with more xeric conditions and the formation of the Irano-Turanian vegetation zone or dry siab-desert and steppe environments. Henry's transhumance model suggests that Pleistocene hunter-gatherers moved between different elevational zones on a seasonal basis (Henry 1982, 1983). This seasonal movement affected group size and composition and would ultimately be reflected in the size and character of the site. Therefore, larger residential sites are located at lower elevations in the winter while smaller sites tend to be found at higher elevations during the summer.

Thus, a "radiating" settlement system was employed in the winter, a

"circulating" system characterized the summer settlement system. This model differs from Marks and Friedel's (1977) model in that settlement pattern changes occur on a seasonal basis; one settlement pattern does 52 not persist through thousands of years in any given mesic or xeric period.

To gather the appropriate environmental and artifactual evidence, the WHPP conducted a systematic survey along the north bank of Wadi Hasa in 1984, 1992 and 1993 (Clark 1992; Clark et al. 1987, 1988, 1992;

Coinman et al. 1989). In addition, six Paleolithic sites identified during the WHS were tested in 1984. Five of these sites had buried, stratified deposits (Clark et al. 1987, 1988; Coinman et al. 1989).

Additional excavations at two Epipaleolithic sites (WHS 1065 and WHS

784) were conducted in 1992 and 1993, respectively (Neeley et al. 1995;

Olszewski et al. 1994). The Late Upper Paleolithic to late

Epipaleolithic sequence was primarily developed from the excavations at three Upper or eastern Hasa sites: WHS 618 (a late Ahmarian site), WHS

1065 (a middle Epipaleolithic site), and WHS 784 (a late Ahmarian and middle-late Epipaleolithic site) (Figure 2.1). Both WHS 1065 and WHS

784 are collapsed rockshelters on the southern beink of Wadi Hasa.

Tor al Tareeq (WHS 1065) - Excavation and Stratigraphy

A description of the excavations conducted at WHS 1065 in 1984 and

1992 by the WHPP is useful in understanding how natural levels 5 and 7,

Step C (the chipped stone material from these levels are analyzed in this study) relate stratigraphically and culturally to the other levels excavated at the site. The 1984 research conducted at WHS 1065 included an intensive surface collection in which 95% of surface artifacts were collected (Coinman et al. 1989) and the excavation of a 44 meter long, stepped trench through the middle of the site (Coinman et al. 1989;

Clark et al. 1987, 1988).

The trench, excavated in eight 5x1 meter trenches (Steps A-H) and one 4x1 meter trench (Step I) (Figure 2.2), was placed

perpendicular to the slope through what appeared to be the heaviest 53

Fiqxire 2.2: Site mao of Tor al-Tareeq (WHS 1065) (after Neeley et al. 1995).

TVPRNIF RFFFINANL

Rf.DRfH'K MORTARS

mil J CIVRSUM RI! V

N\V OIJAI)

RT)RIINI.R.

(INTL ( EXMSRN BEDROCK RACKDIRR

50 <0 N

WHS SITE 1065 rONfOI'R tNTI-RVAI. I M MIDPt-Nd ATPI IN M

SCALE MARI.S 54

concentration of surface artifacts (Clark et al. 1987, 1988). Unless

natural stratigraphic levels were discovered, the steps were excavated

in 10 cm arbitrary levels. The arbitrary levels were later combined

into natural stratigraphic levels based on sedimentological differences

in the excavation profiles. As might be expected, not all of the

natural levels were identified in each excavated step. A correlation of

the natural levels of Step B, Unit B, Unit C and Step C, and the

arbitrary levels from Step C is presented in Table 2.1. Profile

drawings of Steps B and C are presented in Figure 2.3. The arbitrary

and natural levels used in this analysis are presented in bold type. In

this chapter, levels will be used to refer to natural stratigraphic

levels unless otherwise indicated. Over 160,000 pieces of chipped stone

were collected during the 1984 field season (Neeley ec al. 1995).

The best stratigraphic sequences and all of the radiocarbon dates

on charcoal were obtained from Steps A, B, and C (Clark et al. 1987,

1988, 1992; Neeley et al. 1995:23). Step A contained a wall fragment

probably associated with levels 1 and/or 2 (Neeley et al. 1995:23).

Natural levels 1 and 2 are dominated by larger tools and contained very-

few microliths (less than 10%) compared to levels 3 and 4 (62-65%). The

majority of microliths in levels 3 and 4 were non-geometric; the most

abundant non-geometric microlith were straight backed and curved or

arched backed bladelets which comprised 84% of the microlithic types

from Step A (Neeley et al. 1995:23). No geometric microliths were

recorded.

In Step B, non-geometric microliths comprise 50-60% of the

retouched tools in all levels (Clark et al. 1987, 1988; Neeley et al.

1995:23). However, level 5 contained the highest percentages of non- geometric microliths. Geometric microliths were foiand in low frequencies (less than 4%) in all levels except level 1 (11% 55

Table 2.1: Correlations of natural and arbitrary' levels from Steps B and C (excavated in 1984) and Units B and C (excavated in 1992) (after Clark et al. 1987:59 and Neeley et al. 1995:Table 2). The lithic assemblage from arbitrary levels 8-15 in Step C (in bold type) are the focus of this analysis. Note that the natural levels do not necessarily correspond between excavated steps and units.

Natural Natural Sediments Natural Arbitrary Natural Level Level Level Level Level Unit B Step B Step C Step C Unit C (1992) (1984) (1984) (1984) (1992) I 1 mixed 1 1 I surface deposits

-- -- grey ashy 2 2 -- intrusive

-- -- grey 3 2 -- intrusive

III 2 grey -- ?4 II midden

III 3 grey -- -- II calcreted midden IV 4 grey/brwn 4 3-7 II to brown silt/sand

-- -- grey/brwn 5 8-10 II to brown silt/sand

-- -- cobble 6 -- layer V 5 compact 7 11-15 III, IV light tan silt Figure 2.3: The east profile of Seeps B and C (after Clark et al. 1987, 1988)

WHS SITE 1065

N60 STEP B EAST PROFILE

P FS "si . BEDROCK ®

UNEXCAVATED

N55 STEP C EAST PROFILE

POTHOLE BACKOIRT

N50

UNEXCAVATED RODENT

1.0 METER

SCALE 57 microliths). Since the straight and arched backed bladelets are the most dominant type of microliths throughout the levels and "there is some consistency and continuity between these levels", they relate broadly to other industries in the Kebaran Complex (Neeley et al.

1995 :24) .

Step C differs from Step B in that it contained relatively higher frequencies of geometric microliths in the upper levels (levels 1-5).

Level 7 was the only level that contained percentages of geometric (4%) and non-geometries (54%) microliths similar to those of level 5, Step B

(2%, 58%) (Neeley et al. 1995:24). Based on the typological similarities and the greater abundance of geometric microliths within levels 1-5, two temporal occupations are suggested, a stratigraphically higher occupation (levels 1-5) related to a Geometric Kebaran component, with levels 4-5 containing the highest frequencies of the atypical, wide

(Hasa) lunates and bitruncated bladelets, and a stratigraphically lower occupation (level 7) related to industries in the Kebaran Complex (Clark et al. 1987, 1988; Neeley et al. 1995:24).

There are several similarities between the basal levels of Steps B and C. Both basal levels were comprised of a compact, light tan silt that appeared to be an in situ deposit (Clark et al. 1987). step C also contained two hearths (Features 4 and 5) in the lowest level (level 7).

Level 7 contained the highest relative frequencies of bitruncated bladelets and has arched backed bladelets that occur in frequencies similar to those of straight backed bladelets. This trend was also noted in Step B (Neeley et al. 1995:25) . Six charcoal determinations from the lowest levels in Steps A, B, and C yielded dates that range from 15,580 to 16,900 BP (Table 2.2) (Clark et al. 1987, 1988;

Schuldenrein and Clark 1994:34; Neeley et al. 1995).

The lower two-thirds of the trench (Steps D-I) had a 20 cm layer Table 2.2: Radiometric dates from Tor al-Tareeq (WHS 1065) (Clark et al. 1987, 1988; Neeley et al. 1995).

Aae (BP) Level Material Lab. No. 9,010 + :100 Bill Soil/sediment Beta-57898 11,280 + 290 Bill S o i1/sediment Beta-57899 15,580 + 250 B4 Charcoal/hearth UA-43 92 15,860 + 430 B4 Charcoal/hearth UA-43 94 16,570 + 380 C7 Charcoal/hearth UA-4390 16,670 + 270 BIV Charcoal/hearth Beta-57900 16,790 ± 370 C7 Charcoal UA-4393 16,900 + 500 A4 Charcoal UA-43 91 59 of artifacts in a colluvial deposit that was derived frcm upslcpe (Clark et al. 1987, 1988; Coinman et al. 1989). In contrast to Steps A-C, the upper levels of Steps D, E, and F have denser concentrations of artifacts than the lower levels. Since these upper levels include some geometric microliths, they probably correspond to the Geometric Kebaran occupation (Neeley et al. 1995:26). Steps D-I appear to lack a Kebaran component. Therefore, the Kebaran component seems to be confined to the upper slope (Steps A, B, C) (Neeley et al. 1995:26). Steps D, E, and F were probably beneath Lake Hasa during the earliest occupation of the site (Clark et al. 1987, 1988; Neeley et al. 1995:24).

Steps G, H, and I are comprised of marls with small gravel and sand lenses from fluvial and colluvial depositions (Neeley et al.

1995:10). These marls and calcareous deposits are overlain by sands and colluvial debris derived from upslope (Neeley et al. 1995:10). The lack of artifacts in a marl and calcareous deposit suggests that this area was below Lake Hasa during relatively moist periods with high lake levels.

In 1992, two extension units (each 2 x 2 m) were excavated at WHS

1065 in order to further define the Kebaran and possible Geometric

Kebaran occupations, and isolate/define a possible Natufian component

(Clark et al. 1992; Neeley et al. 1995:4). One extension. Unit B, was placed east of Step B and the other, Unit C, west of Step C (Figure

2.2). Over 40,000 lithic artifacts were collected in addition to fauna, shell and groundstone (Neeley et al. 1995). In Unit B, five natural levels were excavated which seem to show a continuation of the stratigraphy and Kebaran component noted in Step B in the 1984 excavations (Neeley et al. 1995) (see Table 2.1). In general, artifact densities are high throughout Unit B. In Unit C, four natural levels were excavated (Neeley et al. 1995:9). Unlike Step C, the four natural 60 levels excavated in Unit C appear to be colluvial deposits (Meeley st al. 1995). Unit C differs from Unit B in that Unit C seems to have less preservation of distinct natural levels from the Kebaran occupation(s) along the lake margin. The stratigraphy in Unit C is probably less distinctive because of greater colluviation in this portion of the site

(Neeley et al. 1995).

The major differences between Unit C and Step C are that the features and living surfaces identified in Step C, levels 7 (east profile) and 4 (west profile), are not present in Unit C, levels II-IV

(Neeley ec al. 1995) . Despite the lack of features and continuous living surfaces between Step C and Unit C, the chipped stone assemblage in both excavated areas exhibit similarities in the frequency and types of artifacts represented (Neeley et al. 1995). It is believed that only these two areas have adequate evidence for differentiating an earlier,

Kebaran occupation (Unit C, levels III and IV, and Step C, level 7) and a later, Geometric Kebaran occupation (Unit C levels I and II, and Step

C, level 5) (Neeley et al. 1995) .

Interpretations

Both the faunal assemblage and pollen profile indicate the presence of an open woodland/steppe mosaic environment throughout most of the occupation of the site (Clark et al. 1987, 1988; Schuldenrein and

Clark 1944:39; Neeley et al. 1995). The faunal assemblage which contains tortoise, gazelle, auroch, equids, ovicaprines, and avifauna indicates diversity in the natural environment (Neeley et al. 1995).

The pollen profile includes cattail iTypha), riparian willow iSalix), alder {Alnus), oak (Quercus) , Chenopodiaceae (usually greater than 50% in most samples), and grasses (Clark et al. 1987, 1988; Schuldenrein and

Clark 1944:39; Neeley et al. 1995).

Although there are changes in these pollen percentages, steppic 61 pollen is dominate throughout the entire occupation of the site v.'hich probably reflects vegetation both at the site and in adjacent areas as airborne pollen from the surrounding landscape is deposited. However, the later occupations at the site, i.e., those occupations associated with greater frequencies of geometric microliths, have increased proportions of riparian taxa. These riparian taxa occur relatively rapidly suggesting that the later occupations at this site enjoyed warmer, more mesic climatic conditions (Neeley et al. 1995). The absence of riparian taxa in the earlier occupations suggests drier, cooler climatic conditions. Based on these data, the margins of

Pleistocene Lake Hasa would have provided a broad range of lake and marsh resources that differed considerably from the surrounding steppe environment (Neeley et al. 1995). This diversity was probably a major factor for prehistoric groups to return repeatedly to the Hasa basin during the Epipaleolithic (Neeley et al. 1995).

Radiocarbon determinations clustering between 15,900-15,000 BP place the occupation(s) of the lowest levels as contemporary with the

Late/Terminal Upper Paleolithic (pre-20,000-ca. 15,000 BP) and the

Kebaran (ca. 19,000-14,500 BP) periods in the Levant (Bar-Yosef and

Vogel 1987; Byrd 1994). However, the overall character of the lithic assemblage, specifically the high proportion of narrow backed bladelets, suggests that most of the cultural deposits are associated with the

Kebaran Complex (Bar-Yosef 1975, 1981; Clark et al. 1987, 1988; Neeley et al. 1995). Typological changes in the microlithic component of the flaked stone assemblage, specifically an increase in the number of geometric microliths in the upper levels of some of the excavated units, suggest a later Epipaleolithic occupation possibly associated with the

Geometric Kebaran Complex (Clark et al. 1987, 1988; Neeley et al. 1995).

The occupations at the site probably moved up and down the slope 62 following the rise and fall of the lake levels (Schuldenrein and Clark

1994). This interpretation is based on the fact that the oldest Kebaran levels are upslope and younger Geometric Kebaran levels are downslope

(Neeley et al. 1995:12). Although lake levels clearly fluctuated over time, the occupational surfaces and in situ deposits may have been scattered along the upper slope without regard to the lake levels

(Neeley et al. 1995:12). This is based on the Kebaran levels in Step B having yoimger radiocarbon determinations than levels in Steps A and C which have the greatest concentrations of artifacts possibly associated with the Geometric Kebaran and Kebaran Complexes (Neeley ec al.

1995:12) . In either case, Tor al-Tareeq was occupied repeatedly because of its location near the margins of Pleistocene Lake Hasa. 63

CHAPTER 3: RESEARCH METHODOLOGY

Chipped stone material comprises the largest artifactual component of Paleolithic sites; therefore, it provides an important line of evidence for examining variations in activities and functions between and within prehistoric settlements. As a result, many inferences and interpretations regarding human adaptions during the Paleolithic are based on typological and technological analyses of chipped stone material. Several systematic typological and technological approaches have been employed in the study of Epipaleolithic chipped stone assemblages in the Levant. A recently developed approach, the chains operatoire or operational sequence, is applied in this study. This approach, largely adapted from French studies (e.g., Lemonnier 1983,

1990), is based on identifying and tracking an artifact through all stages of production from the acquisition of raw materials to core reduction and the manufacture, use, and final discard of the artifact

(Bar-Yosef 1991b). A similar approach, employed by some American archaeologists, is termed the life history approach (e.g., Schiffer

1975:46-48; Rathje and Schiffer 1982:84-89).

An important assumption behind this approach is that operational sequences used to manufacture chipped stone artifacts reflect technical traditions and were learned behaviors "passed from one generation to the next" (Bar-Yosef 1991b:320). If various production strategies and sequences can be identified, Epipaleolithic groups can be distinguished more reliably and relationships between contemporaneous groups better understood (Bar-Yosef 1991b:322). Many archaeologists working in the

Levant have already observed that there is a relationship between reduction strategies and subsistence issues, mobility practices and settlement patterns (e.g., Byrd 1988; Garrard et al. 1987; Henry 1989a,

1989b; Marks 1976, 1977, 1983; Marks and Volkman 1983). Additional 64 studies will enable more meaningful interpretations cf regional variability to be generated.

Acquisition of Raw Material

The use life of an artifact begins with raw material procurement which is influenced by the abundance, availability, quality, and size of the raw material. Researchers frequently note the ubiquity of lithic raw material sources, mainly chert or flint, throughout the Levant

(e.g., Goring-Morris 1995:143). Therefore, raw material constraints are generally not considered to be significant in influencing core reduction strategies. However, the quality and size of the raw material varies throughout the Levant and some preferences in lithic material between geographic regions have been noted (Goring-Morris 1995:143). During mesic climatic periods, a denser vegetation cover would have obscured some of the available raw material. Although there would be less ground cover during xeric periods and lithic raw material would have been more visible, there is little evidence for human occupation in these arid regions during most of the Late Pleistocene. Therefore, suitable raw material may not have been as ubiquitous as some suggest.

There is clearly some relationship between the size of the core and the size of the flake produced. Tools manufactured on larger blanks such as scrapers and burins, may be constrained by the dimensions of the core. However, the size and quality of raw material is not always directly related to the size of the debitage (Bar-Yosef 1991b:322).

Bladelets in many assemblages were produced from varying sizes eind qualities of raw material. Also, the reduction strategy for manufacturing bladelets was not necessarily determined by the size of the raw material (Bar-Yosef 1991b:322). In Epipaleolithic assemblages, bladelet dimensions are somewhat standard and there are only small variations between contemporaneous assemblages (Bar-Yosef 1970, 1991; 65

Henry 1989b). By definition, bladelets are less than 1.2 cm wide and have a length to width ratio of 2:l (Tixier 1963). Some of the potential variability in morphological dimensions is limited by this definition. Regardless of this definition, the fact that bladelets exhibit some uniformity despite variations in core quality and morphology suggests that Epipaleolithic knappers had a preconceived notion on the appropriate dimensions of bladelet blanks (Bar-Yosef

1991:322).

Core Reduction

Several systematic methods and techniques have been employed in the study of Epipaleolithic core reduction strategies, and tool manufacture, use, and discard. Core reduction strategies represent primary lithic technologies, while blank modification and tool manufacture represent secondary lithic technologies. Initially, many of these techniques were pursued individually and focused on only one step in the manufacturing process. The chaines operatoires approach incorporates many of these earlier methods and techniques.

Primary lithic technology or core reduction includes initial core preparation (including the removal of cortical material), preparation of a striking platform, flake or blank removal, and core rejuvenation. The refitting of lithic artifacts to the core from which they were removed is an extremely informative, albeit extremely time consuming, method with which to study primary lithic technology and core reduction strategies. Refitting studies have been used to document the transition between the Middle and Upper Paleolithic showing that cores, flakes, blades, and elongate Mousterian points were produced during the reduction of single nodules (Marks 1983; Marks and Volkman 1983) and in interpretations of site formation processes and discard patterns in the

Upper Paleolithic of Southern Sinai (Phillips and Gladfelter 1991; 66

Phillips 1991). Refitting studies have been employed less frequently cn

Epipaleolithic assemblages; although, there are some notable exceptions

(e.g., Hamifagash IV, western Negev cited in Goring-Morris 1995:156).

Sites with greater variability in lithic raw material and small sites that appear to represent a single chipping episode will be more easily refit than large, multi-component sites which have little variability in lithic raw material. As Tor al-Tareeq falls into this later category and much of the lithic assemblage was fragmentary, this type of study was not considered for this analysis.

Since core reduction or primary lithic technology includes the initial core preparation, preparation of a striking platform, flake removal and core rejuvenation, this study records the number, type cind specific attributes of debitage and cores (see below for definitions of debitage classes and Appendices A and B for coding sheets used in this analysis). There are several assumptions associated with this type of study. Generally, assemblages that contain high proportions of cortical flakes are interpreted as representing initial reduction activities.

The ratio of cores to debitage and the percentage of core rejuvenation elements in an assemblage are used as proxy measures for the relative amount of core reduction occurring at a site. Low core to debitage ratios in an assemblage suggest either intensive core reduction activities (more debitage produced from a single core) or the importation of debitage to the site. High proportions of core rejuvenation elements in an assemblage generally indicate more intensive lithic reduction as the striking platform of the core was rejuvenated so that additional flakes and blades could be removed. Assemblages with more intensive core reduction strategies are frequently interpreted as being more conservative, i.e., lithic raw materials were used more conservatively. Intensive core reduction has also been associated with 67 an increase in mobility (Henry 1989a; Kuhn 1994) .

Secondary lithic technologies include those reduction activities used to modify a blank and are usually associated with tool manufacture.

The most frequent observations about secondary technologies are the type of retouch used for tool manufacture. Although similar edge angles may be produced with different types of retouch, retouch types are often used to define specific tool classes. For example, scrapers are partly defined by their steep angled retouch while microliths are associated with abrupt retouch. Another common secondary technology in some

Epipaleolithic assemblages is the microburin technicjue, a method for truncating or segmenting a bladelet whereby a bladelet is notched and snapped. Microburins are the waste or end product of this technique and are generally considered to be an intermediate stage in the manufacture of some microlithic tools (Henry 1974; Tixier 1973).

Manufacture, Use and Discard

Analyses of retouched tools have also been approached from a variety of avenues relating to the typology (form and function) and technology (primary and secondary manufacture) of the artifact.

Retouched tools are traditionally associated more with typological, rather than technological studies. Early analyses of chipped stone tools involved categorizing them based on a formal type list (e.g., Bar-

Yosef 1970 and Hours 1974 in the Levant). In such typological studies, typological variability was identified but interpreting the variability was more problematic. Some believed that typological variability within an artifact class could be used to identify individual culture groups

(e.g., Bar-Yosef 1970; Hours 1974). As researchers recognized the importance of raw material constraints, tool maintenance activities, curation, and discard behavior on tool morphology, the assumptions behind typological approaches were reevaluated. Although typologies are 68 useful for providing a standard set of terms that can be used to facilitate commimication between researchers, interpretations of the variability represented in tool typologies, especially those relating to functional and stylistic variations, have been modified.

Functional variability in lithic assemblages may be reflected to some extent in the relative frequencies of tool classes such as scrapers and microliths. Current functional interpretations for selected stone tool types are based on ethnographic analogy, microwear analysis of prehistoric tools, experimental research on microwear, and evidence for prehistoric hafting and use. Some of these functional interpretations for the major tool types are presented here so that the interpretive biases of past and present researchers will be made obvious. Backed microliths are generally thought to be associated with either weapons or cutting plant products like grasses (J. Clark 1954,- Clarke 1976; Curwen

1941). Geometric microliths were hafted as projectile points on wooden arrows in Predynastic and Dynastic Egypt (J. Clark et al. 1974; J. Clark

1975-1977). In addition, use wear studies of chipped stone material from el Wad, Ain Mallaha, and Abu Hureyra suggest that geometric microliths were hafted on bone or wood shafts and had polish from cutting meat (Anderson-Gerfaud 1983:78-85; Buller 1983:110-112).

Smaller geometries from Abu Hureyra were interpreted as having been used on the tips of shafts and as barbs (Anderson-Gerfaud 1983) . At el Wad and Ain Mallaha, nongeometric microliths were used on meat probably as a composite tool hafted in wood (Buller 1983). At el Wad (Garrod and Bate

1937) and (Edwards 1990), nongeometric microliths were recovered from bone sickles and were probably used for plant processing.

Therefore, nongeometric microliths are associated with either hiinting or plant processing, and may be associated with other activities as well.

Based on ethnographic analogies, scrapers were probably used for 69 hide processing (Hayden 1979; Nissen and Dittemore 1974) or v.'ccdv.'crking

(Gould et al. 1971; Hayden 1977:182). Burins were probably used as shavers and engravers of wood, antler and bone (Hayden 1977:185; Keeley

1980; Newcomer 1974; Seminov 1964). It has also been suggested that the bit may have been hafted onto wood or bone shafts, and that the blow was one manner of blunting the end of the burin for hafting (Buller

1983:109-110; Mortensen 1970). Burins at Beidha, are sometimes retouched on the ends with very small burin blows (Byrd 1987:102). If the burin was hafted, a burin blow on the lateral edge of the blank may also have served to narrow the hafted end of the piece. Retouched pieces, notches, and denticulates were probably used for a variety of tasks. Denticulates were probably used for cutting sinew and other tough materials that would cut more easily with a serrated edge.

Unretouched pieces that exhibit evidence of utilization were probably also used for a variety of tasks as an expedient tool.

Typological variation within individual tool classes such as scrapers, burins and microliths is attributed to stylistic variability, if such variability is thought unrelated to the function of the implement. Typological variations may be produced by the overall morphology, extent, invasiveness, and positioning of retouch on the modified blank. Some stylistic variability in Epipaleolithic assemblages can be attributed to different temporal and cultural periods, e.g., lunates with Helwan retouch in the Natufian (Henry 1977,

1989b). Some suggest that stylistic variability between contemporaneous lithic assemblages can be used to identify different social groups moving about the landscape (e.g., Bar-Yosef et al. 1992; Close 1978,

1989; Goring-Morris 1987; Henry 1977, 1989b, 1995). Others question the ability to identify prehistoric ethnicity in chipped stone assemblages because of the difficvlty isolating lithic attributes encoded with 70 social symbols (Clark 1989, 1991).

While it is relatively easy to identify attributes that might encode stylistic information, interpreting the meaning of this variability is less obvious. Much of the stylistic variability used to distinguish ethnicity in the Epipaleolithic is associated with the morphology of microlithic tools. Most would agree that microliths are frequently hafted either singly as points or knives, or together as composite tools. If the shape and/or retouch of microliths encode social information, this social information may be intended for distinguishing group identity (emblemic style) or individual identity

(assertive style) (Wiessner 1983:257-258). If this were the case, much of the social information would be "lost" when the microlith is hafted.

In addition, much of the typological variation within the microlithic tool class would be completely unrecognizable from distances greater than two or three meters away (c.f. Wobst 1977). Therefore, it is unlikely that microlithic variability is associated with group recognition or identification. The typological variability may also represent emic notions of how to "correctly" retouch a microlith.

Although these notions could be established ethnographically, they are difficult to demonstrate archaeologically. Therefore in this study, more emphasis is placed on variations in the proportion of tool classes than on variation within a single tool class.

The view here is that technological variability in microliths and other chipped stone tools has potential to inform on social groups if one accepts that technological approaches to core reduction and tool manufacture are learned behaviors. Technology of retouched tools includes the preferential selection of blanks for tool manufacture.

Several studies have been conducted on the selection of blanks for tool manufacture, (e.g., Henry 1973; Marks 1983; Olszewski 1989). Secondary 71 technology includes the microburin techniqu- snd manner of retouch.

Although the type of retouch may not vary significantly within a major tool class, secondary technologies may indicate technological choices used in the manufacturing process to compensate for other factors such as expediency and raw material (specifically quality, size and shape of the raw material) constraints.

The discard patterns in the assemblages analyzed here will be identified by core to debitage and debitage to tool ratios, and percentages of major tool classes. These patterns will then be compared to other contemporaneous levels which will help determine intra-site functional variability. Contemporaneous levels will be determined either by reliable radiocarbon determinations and stratigraphic positioning, or by a combination of stratigraphic positioning and typological similarities in the microliths {i.e., the relative proportions of geometric and non-geometric microliths). Although the later approach is less reliable, the absence of radiocarbon determinations in some levels prohibits the use of the first approach.

Sampling Rationale

Previous analyses of the Tor al-Tareeq lithic assemblage indicated the presence of the microburin technique in the site's earliest levels,

(Step B level 5, Unit B level V, Step C level 7, and Unit C levels III and IV) (Clark et al. 1987, 1988; Donaldson 1986; Neeley et ai. 1995).

In stratigraphically younger levels (Step C level 5 and Unit C level

II), there was an increase in the proportion of geometric microliths accompanied by a significant decrease in the frequency of the microburin technique (Clark et al. 1987, 1988; Donaldson 1986; Neeley et al. 1995).

I wanted to study the technological and typological variability associated with this transition (i.e., from a tool assemblage dominated by nongeometric microliths and use of the microburin technique to a tool 72 assemblage with, high proportions of wide geometric microliths and an almost complete absence of the microburin technicjue) in order to better understand why these techno-typological changes occurred. Particularly,

I was interested in determining to what extent these technological and typological changes reflect only diachronic change or a combination of diachronic change, and changes in subsistence and land use strategies in response to paleoenvironmental change.

This transition was best represented in natural levels 5 and 7 of

Step C (Neeley et al. 1995). I reanalyzed all of the lithic material

(debitage, cores and retouched tools) from the northern half of natural levels 5 and 7 in Step C. Natural level 5, which is characterized by grayish brown to brown silt/sand sediments, corresponds to arbitrary levels COS-CIO (Table 2.1). Natural level 6, an intervening cobble layer, probably indicates an erosional period of unknown duration and lies between natural levels 5 and 7 but has no associated artifacts

(Clark et al. 1987, 1988). Natural level 7, a compact light brown sediment, corresponds to arbitrary levels C11-C15.

As the references to arbitrary and natural levels, and units and steps can be confusing, I want to clarify that most of the following discussion refers to Step C arbitrary levels 8-15 (i.e., C08N-C15, see

Table 2.1). Arbitrary levels will always be referred to by the step, level, and provenience (north, south, or undifferentiated). Therefore, level C08N refers to Step C, arbitrary level 8, northern half of the excavated level; level 014 refers to Step C, arbitrary level 14, and the combined northern and southern halves of the excavated level. The corresponding natural levels in Step C will be referred to as natural levels 5 and 7.

This step and these specific levels were chosen for various reasons. First, the technological and typological transition was best 73 represented in Step C natural levels 5 and 7. Althaugh the general character of the assemblage from Step C is known from in-field analyses and a sample of the microliths has been previously analyzed (Clark et al. 1987:52-67; Donaldson 1986), a thorough description of this assemblage has not yet been conducted. Also, since the microburin technique is not usually associated with Kebaran lithic industries

(e.g., Bar-Yosef 1970) a strict late Kebaran terminology (Clark et al.

1987, 1988; Neeley et al. 1995) does not seem appropriate for the assemblage from Step C, natural level 7. Even using the general term

Kebaran Complex, though certainly chronologically acceptable, does not significantly aid in sorting out the relationship of this site to others in the eastern Levant. Levels COBN-CIO exhibit an increase in the proportion of geometric microliths which suggests an association with a different, later microlithic complex, perhaps the Geometric Kebaran

(14,500-13,000 BP) or Mushabian (14,000-12,5000 BP) as defined by

Phillips and Mintz (1977) . Based on the wide, atypical lunates in the upper levels of Step C (specifically levels C08-C10) , it has been suggested by some that the upper levels may be associated with a

Natufian component at the site (12,800/12,500-10,500 BP)(Clark et al.

1987, 1988; Coinman et al. 1989). Epipaleolithic cultures are largely defined (both temporally and spatially) on the basis of this techno- typological variability in the flaked stone assemblage. Therefore, identifying and interpreting this variability at the site level is necessary before occupations of the site can be placed in a regional context.

Furthermore, the in-field analyses were conducted by people with different levels of expertise. As a result, many retouched pieces and microburins were not consistently identified. The lowest natural levels in Step C (levels 5 and 7) had high artifact densities and are 74 considered to be in situ. Finally, level C13 has two reliable radiocarbon determinations on charcoal, one from an associated hearth

(feature 5) dated to 16,570 + 380 BP {UA-4390) and another from charcoal found in level C13 dated to IS,790 + 340 BP (UA-4393) (Clark et ai.

1987, 1988) . This situation facilitates comparisons between this study and studies on contemporaneous assemblages from other Levantine areas.

During the 1984 excavation, the northern and southern areas of these levels were excavated and collected separately. Since the north end produced more artifacts, I analyzed only the lithics recovered from the north end of Step C (levels C08N, C09N, C11N-C13N). There were, however, three exceptions. Although the north and south ends of level

CIO were excavated and collected separately, this provenience information was not recorded on the three bags of lithics recovered from that level. I selected the largest bag (by volume) from level CIO for analysis. Also as the artifact density decreased significantly in levels C14 and CIS, the north and south ends were not collected separately. Therefore, I analyzed all of the lithics collected from these levels. While the uncertainties with sampling prohibit direct comparisons of artifact frequencies (counts) between levels, the relative frequencies can still be compared. A total of 7952 pieces

(7093 pieces of debitage and 859 retouched pieces) were analyzed for this study.

Analysis of Cores and Debitage

The technological analysis of the debitage is concerned with primary (core reduction) and secondary (retouch and microburin technique) lithic technologies. First, the chipped stone was sorted into major debitage, core, and tool categories. Debitage represents any potential blank for tool manufacture. Debitage was sorted by major blank categories into flakes, blades, bladelets, microburins, burin 75 spalls, core rejuvenation elements, debris and shatter (for debitage coding list see Appendix A).

Flakes represent any blank with a single interior surface

(Sullivan and Rozen 1985:759) and does not contain any of the characteristics used to define the other debitage classes. In this analysis, all complete flakes and proximal flake fragments are counted as flakes. Blades are flakes that are at least twice as long as they are wide. Although blades frequently have parallel sides, they do not necessarily have to. Bladelets are blades with widths less than 1.2 cm and lengths generally less than 3 cm (Tixier 1963) . The division between blades and bladelets is arbitrary; therefore, "blades" will be used to refer to both blades and bladelets unless otherwise specified.

For both blades and bladelets, the number of dorsal flake scars was recorded. A first order blade is a blade with one ridge (a high point between two flake scars) on its dorsal surface. A second order blade has two or more parallel ridges on its dorsal surface. Burin spalls are generally small flakes with triangular or rectangular cross-sections and are removed during burin manufacture. Core rejuvenation elements are any flake or blade that was removed to prepare the core for additional flake removals such as crested blades. Debris consists of medial or distal flake fragments. This category is also referred to simply as

"flake fragments" in other studies (e.g., Sullivan and Rozen 1985). The shatter category, as used in this analysis, consists of any angular fragment produced during lithic manufacture that does not have a single, discemable, interior surface. This category has also been referred to as "debris" by some researchers (e.g., Sullivan and Rozen 1985:759). In other analyses, the shatter category sometimes combines shatter and debris categories as they are defined here (e.g., Clark et al. 1987,

1988) . 76

Microburins, as previously mentioned, are formed through the intentional notching and snapping of bladelets. Based on the appeareince of the microburin scar, three types of microburins are recognized, regular, piquant tiedre, and Krukowski (Tixier 1963). Although part of the debitage, analyses of microburins are frequently presented in discussions of retouched tools. Here, microburins are counted as part of the debitage. However, specific technological and typological attributes of microburins such as type of striking platform, length, width, and thickness were recorded during the attribute analysis of retouched tools (see Appendix B for a list of coded attributes). For each debitage class, the number, presence of cortex, and completeness category (complete or proximal, medial, and distal fragment) were recorded. Additional analyses, mainly recording length, width, and thickness of blade and bladelets, were also conducted.

Cores include all pieces from which three or more blanks have been removed. Core fragments are cores that appear fractured. During the analysis of the debitage, the frequency of cores and core fragments was recorded. Additional information (core type, amount of cortex, lithic raw material, and maximum length, width, thickness and weight) was also recorded during the attribute analyses (see Appendix B for a list of coded attributes).

Much of the chipped stone material is incomplete or fragmentary.

Therefore in an effort to learn more about the geologic integrity of these deposits, some of these debitage categories (flakes and debris) were further subdivided into four length categories (<1, 1-2, 2-3, >3 cm). The assumption is that breakage rates and size categories may be used as proxy measures for the amount of post-depositional disturbance of these assemblages. Archaeological deposits sxibjected to greater post-depositional disturbance may exhibit an increase in the proportion 11 of fragmentary debitage. Also, these data may be useful in interpreting patterns in the debitage cuid tool components of the assemblage.

Analysis of Retouched Tools

In the analysis of retouched tools, I recorded technological

(primary and secondary lithic technologies) as well as typological

(formal) attributes of retouched and utilized tools in each assemblage.

Therefore, an attribute analysis was conducted on all of the retouched and utilized pieces from levels C08N-C15. The coding sheets used in this analysis are a combination of those employed by several researchers. The tool typology with some modifications is based largely on Tixier (1963) and Goring-Morris (1987)• while the technological attributes are based on Phillips and Yerkes (1979).

In the technological analysis of retouched and utilized pieces, the primary technological attributes recorded for each retouched tool and utilized piece included original blank type, amoimt of cortex, distribution of dorsal flake scars, platform type, lithic raw material, length, width, thickness and weight. Several secondary technological attributes (placement, orientation and type of retouch) were also recorded.

The typological analysis included assigning each artifact to a major tool class. The classes used were endscrapers, burins, multi­ purpose tools, retouched blades, tnancated flakes and blades, nongeometric raicroliths, geometric microliths, retouched flakes, retouched pieces, notches and denticulates. Bach of these major tool classes was further divided into subtypes based on the type of blank, location and type retouch, or the overall morphology of the tool. In addition, the shape of the distal end and lateral edges were recorded for each tool. Although some of this information is encoded in the subdivisions of specific tool types, this information was recorded 78 separately so that variations between artifact classes could be mere easily accessed and quantified.

Not all of the attributes recorded are analyzed in this study.

Many of the attributes recorded relate to secondary lithic technologies and/or subtle typological variations. Although this analysis focuses on primary lithic technologies associated with core reduction activities and tool typology, some of the secondary lithic technologies and typological variation exhibited considerable consistency between levels and did not show significant variability within the small sample size.

All of the data were entered into dBase IV version 4.0 during analysis.

Later, these files were imported into SYSTAT version 5.04 for data manipulation and statistical analyses. 79

CHAPTER 4: LITHIC ANALYSES

Technological and typological attributes of the debitage and tool components of the chipped stone assemblage from stratigraphic levels

COSN-CIS were analyzed according to the method presented in the previous chapter. Chipped stone debitage represents both primary and secondary lithic technologies {Goring-Morris 1987:48). Primary technology includes core preparation and blank removal, while secondary technology includes s\absequent blank modification such as the microburin technique, i.e., purposefully sectioning blanks for later use. Chipped stone tools also represent technological (blank selection) and typological

(stylistic) choices. This chapter presents a technological description and analysis of the chipped stone debitage followed by technological and typological analyses of the chipped stone tools for levels C08N-C15.

Debitage

The classification of debitage was hampered by the fragmentary nature of the chipped stone assemblage. Only 8.55-16.34% of the chipped stone flakes, blades, and bladelets in each stratigraphic level are complete (Table 4.1). Proximal debitage fragments make up 25.42-34.41% of each assemblage. The remaining debitage (53.38-66.10% of each assemblage) is comprised of medial or distal fragments. The completeness of cores, core fragments, core rejuvenation elements, microburins, and burin spalls was not recorded. These categories occur only in low frequencies, 4.28% of the entire chipped stone assemblage

(Table 4.3). Fragmentation may have affected the recognition of some categories such as burin spalls since identification of burin spalls requires a triangular cross section and a hinged flake scar termination which may be absent in fragments. Other categories can be recognized despite their degree of completeness. Therefore, burin spalls tended to be small (1-2 cm) and complete. Since there is little evidence for 80

Teible 4.1: Percentages of completeness categories for *flakes, blades and bladelets by level.

C08N C09N CIO CllN C12N C13N C14 C15 TOTAL

Complete 1 8.55 16.34 10.91 11.43 12.25 9.29 12.41 8.47 | 10.90 I I Proximal |32.60 27.45 33.45 33.58 34.38 34.41 30.08 25.42 | 32.57 fragment| | Distal |15.38 8.82 8.63 7.08 14.00 7.60 10.65 3.39 | 10.26 fragment| | Medial |43.47 47.39 47.00 47.91 39.38 48.70 46.87 62.71 | 46.28 fragment TOTAL 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 NUMBER 865 306 834 551 800 1119 798 177 5450

* Completeness categories for cores, core fragments, core rejuvenation elements, microburins and burin spalls were not recorded.

Table 4.2: Percentages of medial and distal fragments classified as blades and bladelets, and debris by level.

C08N C09N CIO CllN C12N C13N C14 C15 TOTAL

Blades andl20.04 23.26 27.59 25.74 20.61 18.57 20.70 20.5l| 21.81 bladelets| { Debris |79.96 76.74 72.41 74.26 79.39 81.43 79.30 79.49| 78.19

TOTAL 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.Q NUMBER 509 172 464 303 427 630 459 117 3081 81

Table 4.3: Debitage and tool percentages by level.

C08N C09N CIO CllN C12N C13N C14 CIS TOTAL

Flakes 24.70 19.55 24.09 23 .93 to 92 22.18 20.00 19.43 22.95

Blades/ 15.13 15.89 20.02 16.07 13.65 13.57 14.44 14.57 15.29 bladelets Core rejuv. 1. 57 1.83 1.77 1.84 1.80 1.12 1. 75 .00 1.56

Microburins • 09 • 61 • 35 1. 35 1.03 2.01 1. 51 .40 1.07

Burin spall 17 • 41 • 27 - 12 • 00 00 • 16 .00 .13

Debris 35. 39 26.88 29. 76 27.61 29.10 30.27 28.89 37.65 30.29

Shatter 12 . 35 14.66 12. 05 12. 76 16.65 19. 29 21. 35 23 .89 16.39

Cores • 51 1.22 1.15 • 86 1.03 65 • 16 .00 .73

Core frags. • 26 • 81 • 80 1.60 1.29 • 71 • 56 .00 .79

Tools 7. 13 12.22 7. 35 9.69 5.67 6.43 7.46 3.24 7.31

Tool frags. 2.61 5.91 2.39 4. 17 3. 86 3.78 3 .73 .81 3 .50

TOTAL 100. 0 100.0 100.0 100.0 100. 0 100.0 100.0 100. 0 100.0 NUMBER 1150 491 1129 815 1165 1695 1260 247 7952

Table 4.4: Ratios and indices of various artifact classes by level.

C08N C09N CIO CllN C12N C13N C14 CIS Flake:Blade 1.63 1.23 1.20 1.49 1.90 1.64 1.38 1.33 Debitage:Core 47.9 18.8 23.86 17.65 18.30 28.65 53.0 0.00 Shatter:Debitage 0.30 0.38 0.26 0.30 0.39 0.50 0.56 0.69 Fragments:Debitage 0.85 0.70 0.64 0.64 0.67 0.78 0.76 1.09 82 burins in the chipped stone tools (Table 4.13). it is doubtful that the fragmentary nature of the assemblage would have significantly affected the recognition of burin spalls.

Since the preponderance of flakes, blades, and bladelets in each level are fragmentary, statements about the frequency of blade and flake blank manufacture and the selection of blanks for tool manufacture could be affected by the treatment of this fragmentary material. The classification of the broken pieces used in this analysis follows Byrd

(1987:91). If a broken piece still had a length to width ratio that was

2:1 along its striking axis, it was classified as a blade or bladelet.

If the length to width ratio was less than 2:1, it was placed either into proximal or medial and distal fragment categories. Complete blcinks and proximal blank fragments were used to determine the total number of flake and blade blanks in each assemblages (Table 4.3). The medial and distal fragments with a length to width ratio less than 2: l were not counted among flake and blade blanks since multiple medial fragments may be produced from a single blank. The inclusion of medial and distal fragments in the overall flake and blade counts might artificially inflate this category.

The number of flakes recorded in this analysis may be more

"upwardly biased" than other classification schemes. Medial blade/bladelet fragments with 2:1 length to width ratios were counted as blades/bladelets. However, if the length to width ratio was less than

2:1, the proximal end could be considered a flake. Therefore, some of the proximal blank fragments may actually have derived from blades and bladelets, not flakes. However, this situation would be equally problematic if proximal blank fragments with length to width ratios of less than 2:1 were assigned to the blade/bladelet category. Some may assign any parallel-sided, proximal blank fragment to a blade and 83 bladelet category. This too can be problematic as the proximal ends of some flakes may have parallel sides at the proximal end. In addition, the length of proximal fragments influences the appearance of the sides of the flake. Although 72.41-81.43% of the medial and distal fragments were debris, i.e., not classifiable to a specific blank category, 18.57-

27.59% of the medial and distal fragments had length to width ratios of at least 2:1 (Table 4.2) and were classified as blades/bladelets.

Some variability in the frequencies of debitage classes exists between levels, but there is considerable consistency overall (Table

4.3). The proportions of flakes (19.43-25.92%) and blades and bladelets

(13.57-16.07%) vary only slightly between levels. Blades and bladelets comprise slightly greater percentages of the total assemblage in levels

C08N-C11N than in levels C12N-C15. Bladelets comprise approximately one-third (30.1-33.3%) of the total debitage blanks, i.e., flakes, blades, and bladelets, from levels C10-C13N (Table 4.5). In the remaining levels, the frequency of bladelets is slightly higher in levels C09N, C14 cind C15 than in level C08N (25.5%) . Blades comprise almost twice the proportion of the total assemblage in the upper levels

(C08N, ClO-CllN) as in the lower levels (C12N-C15)(Table 4.5). Again,

C09N has an anomalously low percentage of blades and a high percentage of bladelets. The variability in levels C09N and CIS may be attributable to the relatively small samples from these levels.

Debris comprises approximately one-third (27.61-37.65%) of each assemblage (Table 4.3). Microburins are rare in the upper levels {C08N-

ClO) but markedly more abundant in levels C11N-C14. Microburin frequencies and their relationship to geometric and non-geometric microlith frequencies will be discussed in greater detail below. Burin spalls are rare in all levels, but slightly more abundant in levels

C08N-C11N. Cores, core fragments and core trimming elements are 84

Table 4.5: Percentages of size categories for complete and proximal flakes, blades, and bladelets by level.

C08N C09N CIO CllN C12N C13N C14 C15 Total Flakes <1 cm 13 54 17.24 16.67 19.94 25.60 28.05 24.88 25.00 21.60

1-2 cm 31 44 25.86 24.90 29.45 27. 98 26.57 23.96 20.24 26. 96

2-3 cm 12 01 8. 05 7.63 5.83 7.59 4.95 7.14 7.14 7.50

>3 cm 5 02 4 .02 5.42 4.60 4 .34 2.48 2.07 4.76 3.95

Blades 12 45 5 .17 12.45 10.74 3 .47 6.94 4.38 5.95 8.55

Bladelets 25 55 39 .66 32.93 29.45 31.02 31.02 37.56 36.90 31.93

TOTAL 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 NUMBER 458 174 498 326 461 606 434 84 3041

Table 4.6: Percentages of debris size categories by level.

C08N C09N CIO CllN C12N C13N C14 CIS Total

<1 cm |43.49 40.91 39.88 40 .44 50.44 61.60 56.32 49.46 49.56 1 1-2 cm i40.29 43 .18 37.50 44.00 37.76 30. 02 37.91 46.24 37.73 j 2-3 cm [12.29 12.88 16.37 12.00 9. 73 6.63 5.49 4.30 9.96 j >3 cm 1 3.93 3.03 6.25 3.56 2.06 1. 75 .27 .00 2.74

TOTAL 100.0 100.0 100.0 100.0 100.0 100 .0 100 .0 100.0 100.0 NtMBER 407 132 336 225 339 513 364 93 2409 85 relatively rare in all levels. This result is consistent with in-field analyses of the debitage (Clark at al. 1987:59). Levels C09N-C12N have slightly higher percentages of cores, core fragments, and core trimming elements than other levels. Shatter varies between 12.05-23.85% of each assemblage (Table 4.3). The lower levels (C10-C15) have higher percentages of shatter than the upper levels (C08N-C09N).

Differences in the frequencies of blade and bladelet blanks between levels are affected by the relative frequency of other debitage and tool categories in that level (Table 4.3) . Flake to blade ratios may be a better comparative measure because variation in other debitage classes will not affect these ratio values (Table 4.4). However, if bladelet blanks are preferentially selected for tool manufacture, their numbers would be reduced in the debitage. Tool blanks were not included in comparisons of blank frequencies (Table 4.4). In-field debitage analyses of natural levels 5 and 7 from Step C indicated that nearly equal numbers (0.8 and 0.7 flake to blade ratios, respectively) of flakes and blades were manufactured (Clark et al. 1987; Donaldson 1986).

Earlier studies suggested that blade manufacture was more heavily emphasized in the earlier occupations of the site, specifically, natural level 2 of Step C and natural levels 3-4 of Step B (Clark et al. 1987;

Donaldson 1986). Similar results were obtained in this analysis for levels C09N and CIO. However, levels C08N and C12N produced almost twice (1.77-1.85) as many flakes as blades. Levels CllN and C13N-C15 produced 1.25-1.48 times as many flakes as blades (Table 4.4).

Since the assemblage is so fragmentary, flakes, flake fragments, and debris were further classified into 1 cm size categories (Tables 4.5 and 4.6). It was hoped that this size classification scheme might be able to inform on (1) the degree to which post-depositional disturbance and breakage rates differed between levels, (2) the relationship between 86 the number and size of the flakes, flake fragments, and debris, and the number ajid size of whole blade blanks, and (3) the potential effect of these values on flake to blade ratios. First, the frequencies of blades, bladelets, and flakes were determined (Table 4.3). Then, complete and proximal flake categories were further subdivided into four, 1 cm size classes (Table 4.5). The unclassifiable debris was also divided into four, 1 cm size classes (Table 4.6).

The size distribution pattern of complete and proximal flake fragments differs slightly from the pattern found in debris. Complete and proximal flakes in the 1-2 cm length category comprised the majority of the flake blanks (24.9-31.4%) in levels C08N-C11N (Table 4.5).

Levels C13N-C15 have only slightly more <1 cm size flakes (by 0.9-4.8%) than 1-2 cm flakes. In contrast, the <1 cm debris is consistently the

most abimdant size category (50.4-61.2%) with the 1-2 cm debris (29.1-

46.2%) being the second most abimdant size category. However, the relative cibiandance of these categories varied between levels. In levels

C12N-C14, the <1 cm size debris was 12.68-31.58% more aibundant than in other levels (Table 4.6). In levels C08N-C11N and C15, the difference between the <1 cm and 1-2 cm size debris was less pronounced varying only 2.7-3.7% between levels. The 2-3 cm and >3 cm size debris are also more abundant in levels C08N-C11N thsin levels C12N-C15. Although the patterning of 2-3 cm size debris is not present in complete and proximal flakes, there is a trend for slightly more >3 cm size flakes in all levels than >3 cm size debris.

There is an inverse relationship between the length of debris and the relative abundance of debris in all levels. Interestingly, the lowest levels (C12N-C15) have at least 10-20% more <1 cm size debris than the upper levels (CGSN-CllN). This break between levels CllN and

C12N in the proportion of small debris is also present between these 87

levels in the proportion of small sized flakes. Based on this

information, it appears that the lowest levels have more small debris

either as a result of debris being eroded down slope or being broken

into smaller fragments in these levels.

The original blank morphology may influence its susceptibility to

post-depositional breakage. In general, blade and bladelet blanks in

levels C12N-C14 are narrower and very slightly thinner than in levels

C08N-C11N (see discussion on debitage morphometries and Figures 4.2-4.6

below). These data cannot be used to support a causal relationship

between blank morphology and general size of debris. However, they do

suggest that assemblages with narrow and thin blanks are more

susceptible to post-depositional stress than wide and thick blanks.

Additional comparisons of flake widths and thicknesses from each level

would provide additional support for this relationship (see section on

the morphometries of flake tools below).

High frequencies of shatter usually indicate more primary core

reduction (Clark et al. 1987, 1988; Sullivan and Rozen 1985); however,

extreme weather conditions in desert environments may inflate the

frequency of shatter as surface scatters have "baked and cooled for

millennia in the desert environments of the " (Clark et al.

1987:54). Shatter percentages reported in previous analyses (Clark et

al. 1987, 1988) are considerably higher than those reported here. From

in-field analyses, shatter was "more common in the excavated sample than

on the site surface (42.9% verses 34.2%)" (Clark et al. 1987:59).

Shatter to debitage ratios from in-field analyses produced values greater than one for all natural subsurface levels except Step C, level

4 (0,76) which may have been exposed to surface weathering processes for a longer period of time than the other natural levels (Clark et al.

1987, 1988) . 88

In this analysis, the amoiant of shatter varied between 12.05-

23.89% of each assemblage (Table 4.3). The proportion of shatter is 2-4% lower in levels C08N-C11N than levels C12N-C15. Also, shatter percentages increase progressively between levels C12N and C15. Shatter to debitage ratios for levels C08N-C15 indicate lower ratios for levels

C08N-C12N and higher ratios for levels C13N-C15 (Table 4.4).

Differences between in-field amalyses and this cinalysis are probably the result of differences in the definition of shatter between the two studies. Clark's (ec al. 1987) shatter category seems to include both shatter and debris. If the debris and shatter categories presented in this study are combined, values more similar to those presented by Clark

(et al. 1987) are obtained.

If increased amounts of shatter represent increased frequencies of primary core reduction as have been suggested by some (Sullivan and

Rozen 1985; Clark et al. 1987), then more primary core reduction may have occurred in the lower than in the upper levels. If this is true, other indicators of primary core reduction, such as the frequency of debitage with cortex, should be higher. However, a comparison of cortex frequency in the debitage and debris by level indicates that the frequency of cortex varies little between levels (Tables 4.7 and 4.8).

Only level C13N has a higher percentage of cortex (by 8.76-16.03%) than other levels, suggesting that more initial core reduction occurred in this level than in other levels. This conclusion is supported somewhat by an analysis of the amount of cortex remaining on cores.

The majority of cores (50-100%) in all levels had some cortex.

However, there is some variability in the amount of cortex (Table 4.9).

Approximately 50% of the cores in levels C08N and C09N were non- cortical; no cores had more than 50% cortex. In contrast, levels ClO-

C13N had low frequencies (1-2) of cores with 51-99% cortex. Levels 89

Table 4.7: Percentages of cortex for flakes, blades, bladeletS; and debris by level.

C08N C09N CIO CllN C12N C13N C14 CIS TOTAL

No cortex 79.88 78.76 76.86 78.40 79.50 71.58 76.57 77.40 76.88

Cortex 20.12 21.24 23.14 21.60 20.50 28.42 23.43 22.60 23 .12

TOTAL 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 NUMBER 865 306 834 551 800 1119 798 177 5450

Table 4.8: Percentages of cortex for flakes, blades, eind bladelets by level.

C08N C09N CIO CllN C12N C13N C14 C15 TOTAL

No cortex 78.60 79.89 78.11 79.45 75.27 63.86 73.27 72.62j 74.32 I Cortex 21.40 20.11 21.89 20.55 24.73 36.14 26.73 27.38| 25.68

TOTAL 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 NUMBER 458 174 498 326 461 606 434 84 3041

Table 4.9: Percentages of cortex on cores by level.

C08N C09N CIO CllN C12N C13N C14 TOTAL

No cortex 42.86 50.00 38.46 .00 25.00 27.27 .00 29.31

1-50% cortex 57.14 50.00 46.15 85.71 66.67 63.64 100.00 62.07

51-99% cortex .00 .00 15.38 14.29 8.33 9.09 .00 8 .62

TOTAL 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 N 7 6 13 7 12 11 2 58

Table 4.10: Sununary statistics for core weights by groups: (1) levels C08N-C10, (2) levels C11N-C13N, and (3) level C14.

GROUP 1 GROUP 2 GROUP 3 NUMBER 26 30 2 MINIMUM 10.1 12.1 23.6 MAXIMUM 157.7 154.7 45.7 RANGE 147.6 142.6 22.1 MEAN 46.2 41.2 34.7 MEDIAN 36.1 28.3 34.7 90

C11N-C13N had more cores with some cortex (1-50% cortex) and fev/er ncn-

cortical cores than levels C08N-C09N. Level C14 has such a small

sample that the significance of the frequency of cortical cores in this

level cannot be evaluated. These data suggest that more initial core

reduction occurred in levels C10-C13N than in levels C08N and C09N.

Core weights were used to access the relative size of cores. The

assumption is that more intensive core reduction should result in

smaller cores. Since over 90% of the entire chipped stone assemblage is

comprised of grayish brown chert, core weights can be used as proxy data

for relative core size. If different lithic materials were being used,

significcuit variability in core weights might reflect differences in raw

material densities, not relative core sizes. Since the frequency of

cores in all levels was low, levels were combined into three groups

which reflect natural levels and changes in artifact densities. Group i

corresponds to levels C08N-C10, group 2 to levels C11N-C13N, and group 3

to level C14. A box plot of core weights by group was produced in order

to determine if core sizes varied significantly between groups (Figure

4.1). Based on the overlapping notches in each group, these cores

cannot be separated into two statistically different populations. The

median values of core weights, while very similar, are slightly greater

in group l (levels C08N-C10) than in group 2 (levels C11N-C13N) (Table

4.10). Based on core weights alone, the relative intensity of core

reduction between levels cannot be determined.

The relative intensity of core reduction in each level can be further evaluated by comparing debitage to core ratios between levels

(Table 4.4). Levels C08N and C14 have the highest debitage to core ratios, 2-3 times higher than other levels. Levels CIO and C13N have only slightly greater values than the other levels. Still, these ratios suggest that more core reduction occurred in levels C08N and C14 than in 91

Fiaure 4.1: Box plot of core weights in grams by groups: (1) levels C08N-C10, (2) levels"cilN-C13N, and (3) level C14.

200

iOU

= 100

d)

'I

S'rangraDhic group 92 levels CIO and C13N. Levels CIO and C13N had only slightly more secondary core reduction than other levels. Alternatively, core reduction activities may have occurred at another portion of the site.

Therefore, the debitage to core ratios presented here may represent only the relative proportion of secondary core reduction activities at this location.

Although cores occurred in relatively low frequencies in all levels, some general trends by groups (levels C08N-C10 and levels CllN-

C14) are present. These trends reflect differences in lithic reduction techniques between stratigraphic groupings. The most ubiquitous core types include single platform blade, bi-directional blade (a combination of opposed, bipolar, and double platform blade cores), multiple platform flake, and multiple platform flake and blade cores (Table 4.11).

Interestingly, group 2 (levels C11N-C14) has about 11% more single platform blade cores, and about 10% more multi-platform flake and blade cores than group l (levels C08N-C10). In contrast, bi-directional cores comprise roughly 21% more of the core assemblage in group 1 than in group 2. Single and multi-platform flake cores occur in slightly higher percentages in group l than in group 2, while multi-platform amorphous cores occur only in group 2.

In the debitage, flake to blade ratios suggested that more flakes than blades were being manufactured in all levels. However, blade cores are the most abundant core type in both stratigraphic groups. Although flake cores occur in all levels, they occur in slightly higher percentages (by about 8%) in group 1 than in group 2. This study suggests that the lithic assemblage in group 2 was more likely to be manufactured from single platform blade and multi-platform flake and blade cores than group 1. Group 1 was more likely to have to have been manufactured from either flake or blade cores. Combination cores (i.e.. Table 4.11: Percentages of core types by groups: (i) levels CO8N-CI0T (2) levels cilN-ci4T

Group 1 Group 2 TOTAL Platform type Single platform blade 23 08 34 .38 29.31

Opposed platform blade 11 54 6.25 8.62

Bipolar platform blade 7 69 .00 3 .45

Double platform blade 11 54 3 .12 6.90

Single platform flake 7 69 3 .12 5 .17

Double platform flake 3 .85 9 .38 6 .90

Multiple platform flake 15 38 6 .25 10 .34

Opposed platform .00 3 .12 1.72 flake and blade Multiple platform 15 .38 25 .00 20 .69 flake and blade Multiple platform .00 9 .38 5 .17 amorphous Other 3 85 .00 1.72

TOTAL 100 00 100.00 100 .00 NUMBER 26 32 58 94 flake and blade cores) in group l occur in low percentages.

Debitage Morphometries

During the Epipaleolithic, there is a trend towards the manufacture of increasingly wide and short bladelet blanks (Henry

1989b) . Since chronometric dates have only been obtained from level

C13N and bladelet blank morphology is somewhat temporally sensitive, the morphometries of blade and bladelet blanks were recorded. In addition, morphometric data can inform on the size criteria of blanks selected for tool manufacture and potentially inform on the issue of post- depositional breakage. Width and thickness measurements were recorded for all blade and bladelet blanks in the debitage. Length measurements were recorded only for complete blade and bladelet blanks.

A notched box plot of the widths of unmodified blade and bladelet blanks (Figure 4.2) shows that widths in levels C12N-C14 are statistically different with a 0.5% confidence interval from widths in levels C08N-C10, CllN and CIS. In notched box plots, notches represent

95% confidence intervals for the sample median. If box plot notches overlap, then the two samples cannot be separated into two statistically different populations. Surprisingly, statistical differences between bladelet widths did not occur only between the natural levels identified during excavation (i.e., levels CIO and CllN). The unmodified blade and bladelets blanks in levels C08N-C10 and C15 are wider than levels C12N-

C14. Level CllN has unmodified blade and bladelet blanks with widths between those foimd in levels CIO and C12N.

Since I expected less statistically variability in the box plots,

I wanted to determine if the shape of the distribution of bladelet widths between adjacent levels also had statistically significant variability. Therefore, a Kolmogorov-Smimov two sample test was conducted to more accurately determine the statistical difference 95

Fioure 4.2: Box plot of the widths of untnodified blsde and blsdslet blanks by level.

40

30

E E - 20 n

10

j_ 0

Level 96 between the shape of the distribution of bladelet widths in each pair of adjacent levels (Table 4.12). This test is appropriate for these data because blade and bladelet widths are not normally distributed, metric, and derived from two independent samples. The null hypothesis that the two samples are the same Ccin be rejected for all of the tests except those between levels CIO and CllN. Since the probability between CIO and CllN (O.OOl) lies below 0.01, the null hypothesis can be rejected.

Therefore, levels CIO and CllN probably do not come from the same population.

Again, differences in blade and bladelet blank widths can be seen in levels C08N-C10 and C11N-C13N. This division coincides with the natural levels identified in this excavation unit. Although the median width value was greater in level CllN than in levels C12N-C14, the shape of the width distribution for level CllN does not differ statistically from levels C12N-C14. The median width value and shape of distribution of unmodified blade and bladelet widths in level CllN suggests that lithic manufacture in level CllN may be transitional reflecting a change in lithic manufacture from the manufacture of narrow bladelets in levels

C12N-C15 towards the manufacture of wide bladelets in levels C08N-C10.

No significant differences in the thickness of blade and bladelet blanks (Figure 4.3) between levels are discemable. Initially, it was thought that if a relationship between width and thickness could be established, i.e., wide bladelet blanks are also thick, then thickness measurements on tools could be used as proxy data for determining the width of the original blank. However, there is little variation in the thickness of blade and bladelet blanks. The variability that does exist does not have a consistent relationship with the width of blade auid bladelet blanks.

Although blade and bladelet blank lengths vary, no clear trend is 97

Table 4.12: Kolmogorov-Smimov two-sided probability test results for blade and bladelet blank widths.

C08N C09N CIO CllN C12N C13N C14 C09N 0.026 CIO . . . 0.038 CllN 0.001 C12N 0.093 C13N 0.558 C14 0.454 C15 0.065 98 apparent (Figure 4.4). However, this result is certainly influenced by the very small sample size in each level. The median values for length of blade and bladelet blanks are somewhat shorter in the upper levels

(C08N-C11N), than in the lower levels (C12N-C13N). However all of the notches of the box plot overlap indicating that the different levels cannot be statistically separated into different groups. One exception, between levels C13N and C14, is present. The lengths in C13N have a narrower range than other levels. The relative amount of post- depositional disturbance and fracturing may be an influencing factor as the above discussion on flake, flake fragment, and debris size classes has suggested.

Debitage Summary

These analyses indicate that technological differences exist between levels C08N-C10 and levels C11N-C13N. The sample from levels

C14 and C15 are small and may produce questionable, unrepresentative results. Based on the data presented above, the lower levels (CllN-

C13N) have higher flake to blade ratios than those presented in earlier studies, suggesting differences in the treatment of fragmentary debitage can significantly affect flake to blade ratios. High flake to blade ratios in the lower levels may be influenced by greater post- depositional stresses and breakage rates in the lower levels. Lower levels (C11N-C13N) have higher percentages of short flakes, flake fragments, and debris than upper levels (C08N-C10). This result was partially influenced by the narrow and thin blade and bladelet blanks in the lower levels. It was suggested that similar morphometric trends

(i.e., narrow and thin dimensions) would also be present in flakes and flake fragments (see the discussion on the morphometries of flake tools below). However, debris to debitage ratios are lower in levels C09N-

C12N than in levels C08N and C13N-C15 indicating that relatively few Picture 4.3: Box olot of the thicknesses of unmodified blade and bladelet blanks by level.

20

15

E E c S 10 0 c X o r. 1-

5

0

Level 100

Fiaure 4.4: Box plot of the lengths of unmodified blade and bladelet blanks by level.

O 40

- m

Level 101 blanks were broken in the middle levels.

Debitage to core ratios are higher in levels C08N, C14 and CIS than in levels C09N-C13N which suggests that these levels may have more intensive core reduction than middle levels. However based on the frequency of cortex and shatter to debitage ratios, levels C13N-C15 have more primary core reduction. It was suggested that variability in these ratios may reflect relative amounts of surface exposure since extreme temperature fluctuations in desert environments may increase the frequency of shatter in an assemblage.

Core reduction strategies and the morphometries of blade and bladelet blanks varied between levels. In general, the lower levels

(C11N-C15) had more single platform bladelet cores and combination

(flake and blade) cores. Blade and bladelet widths were shown to be statistically different between the upper (C08N-C10) and lower (CllN-

C15) levels. However, thicknesses and lengths of blade and bladelet blanks were not statistical different between levels.

Retouched Tools - Typology and Technology

All tools from levels C08N-C15 were placed into general typological categories based on Tixier's (1963, 1974) Epipaleolithic typology. Some modifications, mostly compressing some of the tool types, were made following Goring-Morris (1987). At the most general level, typologies may reflect fxmctional categories between major tool classes including scrapers, burins, notches and denticulates, and microliths. Variations within categories are most likely to exhibit stylistic influences as seen in the shape of microliths. Although raw material constraints may influence variation within a category (Close

1978:3), many stylistic attributes of Epipaleolithic tool assemblages have been used to infer social groups within a chronological period

(e.g., Henry 1989b). Throughout the Epipaleolithic, simple end 102 scrapers, angle burins, and notches and denticulates on blades and bladelets are the most common forms in the major tool classes (Henry

1989b). In this analysis, more emphasis is placed on general tool categories than on stylistic differences within a single category.

Although variation within general tool classes is likely to exhibit stylistic influences and responses to raw material constraints, the sample of non-microlithic tools from levels C08N-C15 was relatively small, i.e., less than 30 non-microlithic tools per level. Generally, this sample size is insufficient for statistically significcint comparisons.

In previous analyses, excavation levels were typologically analyzed then grouped by natural levels when preliminary statistical analyses were conducted (Clark et al. 1987, 1988). Some discemable changes between natural levels 5 and 7 existed in the microlithic component of the assemblage. The most notable differences included the proportions of geometric and non-geometric microliths, and differences in the microburin index (Clark et al. 1987, 1988; Donaldson 1986).

However before lumping excavation levels into natural levels for my analyses, I wanted to first determine if the sedimentological chainge used to distinguish natural levels occurs at the same stratigraphic position as typological and technological changes in the tool assemblage, i.e., between levels CIO and CllN. This is important since combining levels might mask some typological and technological variability in the assemblage. As techno-typological changes are also used to explain cultural variability, determining the synchrony of techno-typological and paleoenvironmental changes should aid in the interpretation of culture change during this period. To facilitate analysis and discussion, the typological and technological attributes of major tool classes will be presented. This will be followed by a 103 discussion of the typological and secondary' technological attributes of microlithic tools.

Major Tool Classes - Typology

All levels have relatively low percentages of retouched tools

(3.24-12.22%) and tool fragments (0.81-5.91%) (Table 4.3). No clear trends are apparent in the proportion of tools between levels, although the tool percentages are slightly higher, by about 1-4.5%, in levels

C08N-C11N and C14, than in levels C12N-C13N. Similarly, no patterned variability in the proportion of major tool classes by level is present

(Table 4.13). Scrapers, carinated tools, burins, multi-purpose tools, retouched and backed blades, truncated pieces, utilized blanks, and

"other" types, each comprise 4% or less of the entire tool assemblage.

Notches and denticulates, and retouched flakes comprise slightly higher percentages of the total tool assemblage (10.71-12.46%) and are slightly more abiondant in levels C08N and C09N than in lower levels. Retouched flakes comprise 5-10.5% more of the assemblage in levels C08N-C11N, C14 and C15, than levels C12N-C13N. In general, notches and denticulates, and retouched flakes comprise greater proportions of tools in upper levels (C08N-C11N) than in lower levels (C12N-C14).

Tool fragments form the second largest tool category (Table 4.13) .

In Table 4.3, the number of tool fragments varied only 1-2% between levels. However, Table 4.3 presents the percentages of both the debitage and tool components of the assemblage. Since tools and tool fragments comprise only a small proportion of the entire assemblage, some tool variability was masked as the result of the larger sample used in Table 4.3. Except for level C09N, levels C11N-C14 had 4-16% more tool fragments than levels C08N and CIO. The relatively high percentages of tool fragments in the lower levels resemble the trend found in the debitage; however with a smaller sample size, this trend 104

Table 4.13: Percentages of major tool classes by level.

C08N C09N CIO CllN C12N C13N C14 CIS TOTAL

Scrapers 3 .57 3 .37 1.82 6.19 2.70 1.16 2.84 10.00 3 03

Carinated .00 1.12 .00 1. 77 .00 .58 .71 .00 58

Burins .89 .00 .91 .00 2.70 .58 .00 .00 70

Multiple .89 .00 2 .73 1.77 .00 .58 .71 .00 93 tools Ret./bckd. .89 1.12 3 .64 2.65 .90 1.16 .71 10.00 1 63 blades Trunc­ .89 3 .37 .91 4.42 3 .60 5.78 2.13 10. 00 3 26 ations Microliths 33 .93 22.47 38.18 22 .12 36.04 26.75 24.82 40.00 29 34

Notches/ 18 .75 23 .60 2 .73 11.50 9.01 14.45 9.93 .00 12 46 dentic. Retouched 9 .82 12.36 13 .64 14.16 4.50 3 .47 19.15 10.00 10 71 flake Retouched 3 .57 11.24 8 .18 11.50 6.31 2.89 5.67 , 00 6 52 piece Utilized 3 .57 .00 10 .00 4.42 .00 4.05 5.67 .00 4 07 blank Other .00 .00 .91 .88 .00 3 .47 .00 .00 93

Ret./bckd. 23 .21 21.35 16 .36 18 .58 34 .23 34.10 27.66 20.00 25 84 fragment NUMBER 112 89 110 113 111 173 141 10 859 TOTAL % 100 100 100 100 100 100 100 100 100.0 105 appears more pronounced in the tool component.

Microliths comprise the most abundant tool category in all levels

(ca. 22.12-38.18%) (Table 4.13). For now, microlithic tools will be treated as one of the major Cool classes. Stylistic variation within this tool class and its implications will be discussed in greater detail below.

Major Tool Classes - Technology

Some technological variability exists in the selection of blanks for some tools. Blank selection criteria for major tool classes (i.e., scrapers, truncations, notches and denticulates, and utilized pieces) were cinalyzed by determining variations in the proportion of blanks in each tool class by level (Table 4.14). Several of the major tool classes were not analyzed because they either occur in low frequencies

(e.g., carinated tools, burins, and multiple tools) or their blank type is explicit in the tool type (e.g., retouched and backed blades, microliths, and retouched flakes). In order to determine if specific blanks were being preferentially selected for tool manufacture, a

Pearson chi-square test was applied to four tool categories (i.e., scrapers, trvincations, notches and denticulates, and utilized pieces).

Pearson chi-square is an appropriate test for these data because the data are nominal and in a symmetrical, polytomy table. However, one major problem with this independence test is that it is strongly affected by small sample sizes. In order to decrease the number of sparse cells in this analysis, level C15 which has a very small sample was excluded. In addition, the Pearson chi-square test was conducted on the frequency of blank type by level (i.e., C08N-C14) and by group.

Levels were grouped into group 1 (natural level 5, levels C08N-C10) and group 2 (upper portion natural level 7, levels C11N-C13N) in order to decrease the number of sparse cells and to compare differences between lOS natural levels.

For truncations, and notches and denticulates, comparisons of blank type by level and by group produced probability values greater than 0.05 (Table 5.15). If the stcuidard confidence level of 0.05 is used, the null hypothesis that the rows and column totals in the table are independent caimot be rejected. Therefore, there is no statistical difference in bleink selection either by level or by group for truncations, and notches and denticulates.

For scrapers and utilized pieces, comparisons of blank type by level and by group produced probability values less than 0.05.

Therefore, the null hypothesis can be rejected and blank selection for scrapers and utilized pieces by level and by group is statistically dependent. Scrapers in levels C08N-C10 are more often manufactured on blades, while scrapers in levels C11N-C15 are more likely to be manufactured on bladelets and flakes (Tables 4.14 and 4.15). The scrapers on bladelets were small thumbnail scrapers and foiand only in levels CllN and C12N. This is interesting because simple end scrapers on blades and bladelets are usually the most common scraper type in

Epipaleolithic assemblages (Henry I989b:85). Utilized pieces are manufactured more frequently on flakes in levels C08N-C11N than in levels C12N-C13N. Despite the grouping of levels, these significemce tests may still be suspect as some sample sizes remained low. In order to determine if these statistical differences reflect a general trend in retouched tools, a Pearson chi-square test was conducted on the blank types of all retouched tools by level and by group (Tables 4.14 and

4.15) . Blank selection for retouched tools by level is statistically significant. In contrast, blank selection for retouched tools by groups is not significant indicating that blank selection does not vary between upper (C08N-C10) and lower levels (C11N-C13N). 107

Table 4.14: Row percentages of blank t^'pe fcr major tccl classes by level. (Level CIS is excluded in order to decrease the number of sparse cells.)

C08N C09N CIO CllN C12N C13N C14 TOTAL % Scraoers Blade 100.0 66.7 100.0 14 .3 33 .3 0 0 36.4 Bladelet 0 0 0 71.4 33 .3 0 0 27.3 Flake 0 33 3 0 14 .3 33 .4 100 .0 100.0 36.3 (Number = 22) 100.0 Truncations Blade 0 33. 3 100.0 60 .0 25 .0 90.0 66,7 64.3 Bladelet 0 0 0 0 25 .0 0 0 3.6 Flake ! 100.0 66. 7 0 40.0 50 .0 10.0 33 .3 32.1 (Number = 28) 100.0 Notches/dents. Blade 35.0 50.0 100.0 46.2 50.0 39.1 14.3 40.8 Bladelet 25.0 15. 0 0 15 .4 10 .0 26.1 28 .6 20.4 Flake J 40. 0 35. 0 0 38 .5 40 .0 34 .8 57.1 38.8 (Number = 103) 100.0 Utilized oiece Blade 0 0 36 .4 20.0 0 33 .3 50.0 1 32.4 Bladelet 25.0 0 27.3 40.0 0 66.7 0 1 29.4 Flake 75.0 0 36.4 40.0 0 0 50. 0 ! 38 .2 (Number = 34) 100.0

All retouched tools Blade 15 1 27 8 18 9 17 2 12.9 25 5 9 .3 118 .26 Bladelet 57 5 40 7 57 0 46 7 66.1 59 2 44.2 |53 .02 Flake J 27 4 31 5 24 1 36 0 21.0 15 3 46.5 1 28 .72 (Number = 527) 100.0

TABLE 4.15: Pearson chi-square test for independence of blank type by level and blank type by group* for four major tool classes and all retouched tools.

Scrapers VALUE DF PROB blank type by level 24.270 12 0,019 blank type by group 9. 090 2 0.028 Truncations blank type by level 13.806 12 0.313 blank type by group 2 .122 2 0.346 Notches and denticulates blank type by level 11.064 12 0.523 blank type by group 0.087 2 0.993 Utilized blank type by level 11.787 8 0.000 blank type by group 27.559 2 0.000 All retouched tools blank type by level 36 . 76 14 0.001 blank type by group 0. 965 2 0.617

* Groups are group 1 (levels C08N-C10) and group 2 (levels C11N-C13N).

** More than one-fifth of fitted cells were sparse (frequency < 5). Therefore, significance tests are suspect. Boldface type indicates statistically significant values. 108

Since significance tests for the preferentially selection of blade, bladelet, and flake blanks for tool manufacture by levels and by groups are frequently suspect, perhaps the size and shape of blanks selected for tool manufacture would exhibit significant variability.

Therefore, morphometric data were recorded on all tools in order to determine if blanks with a certain range of measurements were being preferentially selected for tool manufacture. Nomtially, researchers emphasize microlithic tools and bladelet blanks in Epipaleolithic assemblages. As previously mentioned, the definition of bladelets places an arbitrary division between blades and bladelets such that bladelets have lengths less than 4 cm and widths less than 1.2 cm

(Tixier 1963). While this bladelet definition is now conventional, I wanted to analyze the widths of both blades and bladelets together so that changes in width distributions would more accurately reflect sample-wide width distribution data. Therefore, tools are grouped by blank types into blade/bladelet and flake tool categories. Furthermore, since many major tool categories occur in low frequencies, analyses of morphometric variations within tool classes would likely be heavily influenced by small sample sizes.

As there is a trend throughout the Epipaleolithic towards increasingly wide and short bladelet blanks and the resulting microlithic tools (Henry 1989b:93; Neeley and Barton 1994:282), variability in blank and microlithic tool morphology might reflect regional trends and help to identify temporal variations within the assemblage. I was interested in determining whether this trend towards wide and short blanks is found in both the blade/bladelet and flake components of this assemblage. However as the lithic assemblage at WHS

1065 is 98% incomplete or fragmentary, the utility of blank morphology data, at least for length measurements and width to length ratios, is 109 limited.

Blade/bladelet blanks and tools are more likely to break perpendicular to their length than parallel to it when siibjected to compaction stresses and colluviation. Unmodified blanks and retouched tools having a "rounder" morphology would be more likely to break around the edges rather than perpendicular to the length of the blank. Since tool widths and thicknesses are more likely to be preserved than tool lengths, analyz .ng the widths and thicknesses of tools in these levels enables a small aspect of overall blank morphology to be investigated.

Length measurements are analyzed only for flake tools because blade and bladelet tools are not present in sufficient quantities to enable statistically meaningful results.

Flake tool widths, thicknesses, ajad lengths were plotted in separate notched box plots (Figures 4.5-4.7), In the notched box plot of flake tool widths (Figure 4.5), all levels have overlapping notches.

Therefore, they cannot be separated into to statistically different populations. However, flake tools in levels C08N-C11N have wider median values than levels C12N-C14 and range between 19.8-25.9 mm. The mean for flake tool widths in levels C08N-C11N ranges between 20.46-24.70 mm.

The mean and median widths in levels C12N-C14 ranges between 17.4S-X9.19 mm and 14-16.5 mm, respectively. Level CIS has a very small sample size

(two pieces) and will not be discussed further.

The plots of flake tool thicknesses and lengths exhibit similar trends to those found in flake tool widths; although, specific values vary. Flake tool thicknesses have fairly confined ranges (Figure 4.6).

Median thickness values are slightly higher (by about 1 mm) in the upper levels (C08N-C11N) than in the lower levels (C12N-C14). Flake tool lengths again seem to show a break between upper levels (C08N-C11N) and lower levels (C12N-C14) (Figure 4.7). Figure 4.5: Box plot of flake tool widths by level.

DO

40 r

E 30 C y W \ I on / 1 U/' I i_l

' / ri 1 'J I \ I \ ^ A 1C r

r,\0, Q\6

Levei Ill

Fiqure 4.6: Box plot of flake tool thicknesses by level.

Level Figure 4.7: Box plot of flake cool lengths by level.

J

i 1 r i 1 1 I

! : i 1 ) : i i « -'inti 1 * n r- . ' u 1 i;H \1 i !\H ! ' in: U n M H ' : n ^ n \ - i m : i i : 1 1 ; i i '

J 1 :

•.J^ -- 1^, 'j • O 'V./

Level 113

Figures 4.8 and 4.9 are notched box plots of blade and bladelet widths and thicknesses by levels. In Figure 4.8, the confidence intervals (notches) in levels C08N, C09N, CIO, and CllN overlap with each other but do not overlap with levels C12N, C13N and C14. Levels

C12N, C13N and C14 also have notches that overlap with each other. This suggests that tool blank selection (at least for widths) in levels C08N-

CllN is statistically different with a 95% confidence interval from lower levels {C12N-C14). Level CIS has a notch which is below the lower quartile. Although its notches overlap with all of the other levels, this is probably due to the small, variable sample (8 pieces) from this level,

Summary statistics for the blade and bladelet tool widths reflect the same trend, showing the most pronounced break in median values between levels CllN and C12N (Tcible 4.17) . However, mean values exhibit a more gradual change with the greatest difference between levels CIO and CllN. In order to determine the extent to which outliers and stragglers in level CIO influenced these results, another set of summary statistics was generated for bladelet tool widths (Table 4.18); blades were excluded. Again, a clear break can be seen between levels CllN and

C12N.

A series of Kolmogorov-Smimov two sample probability tests were conducted on bladelet tool widths to determine if the overall shape of the distribution is similar between adjacent levels (Table 4.19). This non-parametric test is especially appropriate here because the data are metric, not normally distributed, and derived from two independent samples. The null hypothesis that the two samples are the same, can be rejected for all of the tests except that between levels CllN and C12N.

Since the probability level between levels CllN and C12N (0.007) lies below 0.01 the null hypothesis can be rejected. Levels CllN and C12N 114

Figure 4.8; Box plot of blade and bladelet tool widths by level.

50

40

I 30

S I 20

n A ^ R 10 u ^ n V 0

Level 115

Ficmre 4.9: Box plot of blade and bladelet tool thicknesses by iWel. (Outliers greater than 25 mm have been removed.)

20 r

- i;

0?

0 r i

^ W f n

J '-J u

Level 116

Tcible 4.16; Summary statistics for flake tool widths and thichnesses by level.

Width: C08N C09N CIO CllN C12N C13N C14 CIS N of Cases 24 25 23 38 17 17 44 2 MINIMUM 6.2 9.3 5.5 4.8 6.1 9.9 9.0 19.7 MAXIMUM 42.9 42.1 46.0 36.9 35.6 33 .5 39.1 24.2 RANGE 36.7 32. 8 40.5 32.1 29.5 23 .6 30.1 4.5 MEAN 24.7 21.2 20.9 20.5 17.6 17.5 19.2 22.0 MEDIAN 26.0 20.8 21.2 19 .8 14.0 15 .2 16.5 22.0

Thickness: C08N C09N CIO CllN C12N C13N C14 C15 MINIMUM 2.3 2.3 1.6 1.5 2.2 2.1 1.9 5.7 MAXIMUM 16.7 14 .4 20.7 49.0 14.0 12.2 15.3 6.6 RANGE 14.4 12.1 19.1 47.5 11.8 10.1 13 .4 0.9 MEAN 6.9 5.9 5.7 7.4 5.7 5.4 5.3 6.2 MEDIAN 6.6 5.0 5.6 6.1 4.5 4.3 4.8 6.2

Table 4.17: Summary statistics for blade and bladelet tool widths by level.

C08N C09N CIO CllN C12N C13N C14 C15 N OF CASES 83 57 78 69 87 145 86 8 MINIMUM 4.1 3.1 3.9 3.2 3.0 2.3 2.5 4.4 MAXIMUM 45.4 32.3 42.5 24.3 23.9 24.5 21. 9 17.9 RANGE 41.3 29.2 38.6 21.1 20.9 22.2 19.4 13 . 5 MEAN 12 .6 12.7 12.9 10.9 9.0 8.5 8.0 9.4 MEDIAN 11.1 11.4 11.4 10.6 7.6 6.6 6.0 7.9

Table 4.18: Summary statistics for bladelet tool widths by level.

C08N C09N CIO CllN C12N C13N C14 C15 N OF CASES 56 37 51 51 69 114 68 6 MINIMUM 4.1 3.1 3.9 3.2 3.0 2.3 2.5 4.4 MAXIMUM 15.0 14 .6 13 .1 16.3 23 .5 12 .6 13 .9 9.9 RANGE 10.9 11.5 9.2 13.1 20.5 10.3 11.4 5.5 MEAN 9.2 8.8 9.4 8.9 7.2 6.2 6.1 6.8 MEDIAN 9.4 9.1 10.4 9.2 6.7 5.4 5.2 5.9

Table 4.19: Kolmogorov-Smimov two sample probability test for bladelet tool widths by level.

LEVEL C08N C09N CIO CllN C12N C13N C14 C09N 0.411 CIO .... 0.273 CllN 0.373 C12N 0.007 C13N 0.073 C14 0.626 C15 0.459 117 probably do not come from the same population.

Blade and bladelet tool thicknesses exhibit a similar trend in that upper levels {C08N-C11N) have thicker blade and bladelet tools than lower levels (C12N-C14) (Figure 4.9). But, the confidence intervals for blade and bladelet tool thicknesses suggest that the null hypothesis, i.e., that the samples are from the same population, cannot be rejected.

Complete blade and bladelet tools occur in such low frec[uencies that a statistically significant sample was not availcible for determining variations between width and length.

In sum, blade/bladelet and flake tools all exhibit the same general trends in dimensions. Levels C08N-C11N have wider and thicker tools than levels C12N-C14 for both blade/bladelet smd flake tools.

This change is interesting because the interval of change does not correspond with the stratigraphic break between natural levels 5 and 7.

This suggests that lithic tool manufacture changed before the erosional event, as marked by natural level 6, an erosional cobble layer. This result differs from the result for unmodified blade/bladelet widths.

Unmodified blade/bladelet widths change significantly between levels CIO and CllN suggesting that wide blade/bladelet tools were manufactured before primary lithic manufacture produced significantly wide blade/bladelet blanks.

Histograms of blade and bladelet tool widths were generated in order to determine the distribution and possible modalities in the data

(Figures 4.10 and 4.11). It was already known that wider tools were manufactured in levels C08N-C11N than in levels C12N-C14 from the notched box plots. So what new information is gleaned from the histograms? Levels C08N-C11N have similar multi-modal distributions with modes occurring at roughly 5-6 mm and 11-12 mm. However, the height of each mode varies between levels. Generally, a higher 118

Figure 4.10: Histograms of blade and bladelet tool widths for levels C08N-C11N.

020

0.20

5 015 CD •X LU Q.

5 f- (T oQ.

§ Q05 LL

JE M 20 2G

C08N ClO

015 -I

tr < 03 0.10 -I CC 0 111 11 0

005

u 20 26

C09M 119

Figure 4.11: Histograms of blade and bladelet tool widths for levels C12N-C15.

015

020 0 ;c ^ X !0 0 !o ^

in x 0.10 0 i 005 Cl iX n a. iX - 005 ^ ' 1 I ir 10 U 20 U 20 26 C12N CU

n ! I ; I t~n

c: 0 :0 • iij 02-J y I— !i li n cr M i j05 - i i u i i T n !i f n: nii r!! : r ! i LL nj ^11 Ml

20 26 14 20

C13f-i C15 120 proportion of 11-12 mm wide bladelet tools than 5-6 mm wide bladelet tools are present in the upper levels (C08N, ClO and CllN). Levels

C12N-C14 all have right skewed distributions with the greatest mode at about 3 mm. Although Kolmogorov-Smimov two sample probability tests identified significant variation between the distribution of

blade/bladelet widths in levels C08N-C11N and C12N-C13N, the histograms

suggest an even more pronounced difference between central tendencies of blade/bladelet widths than is indicated by summary statistics and notched box plots. Bladelets from levels C08N-C11N are commonly 11-12

mm wide while bladelets from levels C12N-C14 are commonly 3 mm wide.

The lower mean aind median width values for levels C12N-C14 reflect the

increased proportion of very narrow microliths in these assemblages, rather than a decreased range in microlith widths.

The bimodal distribution of blade/bladelet tool widths is not due

to the original blank morphology. Unlike blade and bladelet tools,

unmodified blade/bladelet widths are not bimodally distributed but have

slightly right skewed distributions in all levels, except level C15

which is discounted in this analysis because of its small sample size

(Figures 4.12 and 4.13). This suggests that two major width classes of

blade and bladelet tools were preferentially manufactured in the lower

levels. The very narrow bladelet tools are either backed on one edge,

backed on two edges, or backed and retouched. To manufacture these very

narrow bladelet tools, multiple retouch episodes, i.e., heavy backing,

may have been necessary in order to narrow the width of the tool.

Since secondary lithic reduction (i.e., retouch) for retooling or

resharpening purposes may significantly alter the original blank

morphology, the number, distribution and orientation of dorsal flakes

scars on blade/bladelet and flake tools should help determine potential

differences in core reduction strategies and blank selection. The 121

Fiaure 4.12: Histograms of unmodified blade and bladelet blank widths for levels C08N-C11N.

026

020 tc lU 0. 015 - z 0 I-er 0.10- 0a. 0 J 005

~r~ -f=^ U 20 26 32 U 20 26 G2

C08N CIO

03

tr < CD tr 02 - lU Q. 2 Q

§cc a.

i~i —T" a JTL U 20 28 32 U 20 26 32 C09N C11N 122

Pi 4.13: Hiscograms of unmodified blade and bladelet blank widths for levels C12N-C15.

03 n

It < m tr 02

z 0 (t(- ai £

M=q- U 20 28 32 14 20 26 32

C12N CU

03 r a. < iD (T 02 - UJ (L

CE 2 0-1 0 rrG.

14 20 26 32 S 14 20 26 32

C13N C15 123 assumption is that blade and bladelet tools would have one cr r^.cre dorsal flake scars parallel to the striking cixis of the original blank.

Flake tools would be more likely to have one or more dorsal flake scars that were not parallel or a mix of parallel and non-parallel dorsal flake scars. If wide tools in levels C08N-C11N reflect differences in blank selection, this would be an important technological difference between the assemblages.

As expected, the majority of blade and bladelet tools had dorsal flake scars parallel to the striking axis of the flake (Table 4.21).

Interestingly, the number of dorsal scars parallel to the striking axis varied between levels. Levels C08N-C11N had about twice as many blades and bladelets with one dorsal flake scar parallel to the striking axis than levels C12N-C15. Although variable, tools with more than three dorsal flake scars parallel to the striking axis are the most common flake scar pattern in all levels. The second most common category is two flake scars parallel to the striking axis. Except for level C13N, which has many tools with more than three, parallel dorsal flake scars, little variability exists between levels. Not only did level C13N produce higher proportions of single platform bladelet cores than other levels (Table 4.11), it also has high frequencies of very narrow microlithic tools. These narrow microliths are too narrow, in most instances, to have three or more dorsal flake scars.

The proportion of blade and bladelet tools with three or more parallel, dorsal flake scars, and mixed parallel and non-parallel dorsal flake scars is substantially higher in levels C08N and C09N than in lower levels. These values may be related to the relatively wide blanks that are manufactured in the upper levels. Although parallel flake scars on dorsal surfaces are frequently present, their presence or absence is not the determining criterion. As sample sizes are 124

Table 4.20: Summary statistics for blade and bladelet blank v/idths by level.

C08N C09N CIO CllN C12N C13N C14 C15 N OF CASES 174 81 227 129 149 279 182 37 MINIMUM 3.6 3.6 3.7 3.2 2. 8 2.7 2.6 3.4 MAXIMUM 36.0 22.0 28 .6 33 .1 29.0 25.2 20.6 15.2 RANGE 32 .4 18.4 24.9 29.9 26.2 22.5 18. 0 11.0 MEAN 10.9 9.3 10 .5 9.5 8.1 8.0 8.0 8.3 MEDIAN 10. 0 8.7 9.8 8.3 7.3 7.4 7.1 8.9

Table 4.21: Distribution of dorsal flake scars on blade and bladelet tools by level.

C08N C09N CIO CllN C12N C13N C14 CIS TOTAL N/%

N/A • 00 .00 .00 .00 1.18 .00 .00 .00 .17

Cortex - 00 .00 .00 .00 1.18 .71 1.20 .00 .51

1 parallel 6 .67 7.27 5.13 9.22 20.48 .00 9.78 scar 2 parallel 26 .67 29.09 29.49 26.64 scars 3+parallel 52 . 00 45 .45 56.69 54.97 scars 3+ not 5 . 33 5.45 6.41 . 00 4 .26 .00 . 00 3.71 parallel Multiple 9 .33 12.73 1.28 1.42 3 .61 .00 4.22 mixed TOTAL % 100.0 100. 0 100.0 LOO.O .00.0 .00.0 100.0 NUMBER 75 55 78 68 85 141 83 8 593

Table 4.22: Distribution of dorsal flake scars on flake tools by level.

C08N C09N CIO CllN C12N C13N C14 C15 TOTAL N/A .00 .00 .00 4 .00 • 00 5 .88 .00 .00 1.05 Cortex .00 . 00 4 .00 2.63 5.88 • 00 6 .82 .00 3.16

1 parallel 20.83 8. 00 .00 7.89 5 .88 41.18 15 .91 .00 13 .16 scars 2 parallel 16.67 32. 00 30.43 21.05 11.76 5.88 15. 90 50.00 20.00 scars 3+parallel 12 . 50 20.00 34 .78 26.32 17. 65 23 . 53 27.27 .00 23.68 scars 1 not .00 . 00 .00 .00 00 00 4.55 .00 1.05 parallel Multiple 37.50 24.00 17.39 42.11 47 . 06 23 .53 18. 18 .00 28.95 not paral. Multiple 12 . 50 8.00 17 .39 .00 11.76 • 00 11.36 50. 00 8.95 mixed TOTAL % LOO . 0 100 . 0 100.0 100.0 100 .0 100.0 100.0 100.0 100. NUMBER 24 25 23 38 17 17 44 2 190 125 relatively small, these results should not be considered conclusive.

Flake tools comprise a greater proportion of the tool assemblage in the upper levels (C08N-C11N) than in the lower levels {C12N and C13N)

(Table 4.14). As expected, the majority of flake tools had multiple dorsal flake scars not parallel to the direction of the striking axis.

Many flake tools in level C13N had one dorsal flake scar parallel to the striking axis of the blank. Flake tools with two parallel, dorsal flake scars comprised a lower proportion of the flake tool assemblage in lower levels (C12N-C14) than in upper levels (C08N-C11N). Multiple (more than three) parallel dorsal flake scars are variable in the upper and lower levels exhibiting no clear trends between levels. Level C14 had the only occurrence of a single dorsal flake scar not parallel to the striking axis of the flake. This type of dorsal surface is most common in levels C11N-C12N and C08N. The occurrence of mixed dorsal flake scar patterns, i.e., parallel and non-parallel flake scars orientated with respect to the direction of the blow, is variable. When mixed dorsal flake scar patterns occur, they comprise similar proportions of the dorsal flake scar pattern on flake tools. However, mixed flake scar patterns are absent in levels CllN and C13N, perhaps reflecting differences in the relative amoiant of core reduction in these levels.

In sum, the upper levels have higher percentages of blades and bladelets with non-parallel or mixed (parallel and non-parallel) dorsal flake scars than lower levels. This suggests that core reduction strategies varied somewhat between the upper and lower levels. In the upper levels, blade, bladelet, and flake tools are more likely to have multiple non-parallel or mixed dorsal flake scars which may reflect the increased emphasis on manufacturing wide blanks. Blade or bladelet technology in the upper levels was still important but perhaps more variability existed in lithic reduction strategies. Cores were more 126 likely to be used for both flake and blade mamufacture in the upper levels. The lower levels were more likely to have prepared cores that enabled multiple blades and bladelets to be removed from a single platform, as seen in the high frequencies of single platform cores and tools with multiple parallel, dorsal flake scars.

Next, I wanted to determine if this variability in blade/bladelet and flake tool widths, thicknesses and lengths was reflected in other lithic manufacturing attributes such as striking platforms and retouch types. The results were disappointing. Both the type of striking platform and type of retouch do not seem to vary significantly between levels (Tables 4.23 and 4.24). The striking platform on tools was recorded in order to identify potential differences in blank manufacture. As might be expected, the majority of tools did not have the striking platforms preserved as the proximal end of the blank was often removed by retouch or broken. Large broken tools were not selected out with the tool fragments because they were sufficiently large and complete to be categorized.

Some variability between levels is apparent (Table 4.23). There seems to be a very slight increase in the proportion of plain platforms in levels C08N-C11N. In contrast, punctiform platforms are more common in lower levels, especially level C13N, but infrequent in levels CIO,

CllN, and C14. This reflects an increase in the number of bladelets removed with the punch technique and an emphasis on narrow blank manufacture mentioned earlier. The manufacture of narrow bladelets may require more controlled blank removal than the removal of wide blanks.

Crushed platforms were also somewhat more abxandant in levels C09N-C10 than in other levels. There were few dihedral platforms in these assemblages Again, level CIS is discounted because of its small sample size. 127

Table 4.23: Percentages of tool platform types by level.

C08N C09N CIO CllN C12N C13N C14 C15 TOTAL

Absent 66 .28 67 .14 73 .91 67 39 73 97 75 44 74 51 75 00 71.59

Cortex 1.16 .00 2.17 1 09 4 11 88 98 00 1.41

Plain 12 .79 10.00 15 .22 17 39 6 85 7 02 10 78 25 .00 11.62

Faceted 1.16 .00 1. 09 1 09 1 37 88 98 .00 .94

Pxinctiform 8. 14 10.00 1.09 5 43 8 22 11.40 3 92 .00 6 .75

Crushed 2.33 10 .00 6.52 3 26 2 74 2 63 2 94 .00 4.08

Broken 2 .33 2 .86 .00 3 25 1 37 1 75 5 88 .00 2 .51

Dihedral 5 .81 .00 .00 1 09 00 00 00 .00 94

Other 00 00 • • - 00 00 1 37 00 00 00 • 16

TOTAL % 100.0 100.0 100. 0 100.0 100.0 100.0 100.0 100.0 100.0 NUMBER 86 70 92 92 73 114 102 8 537

Table 4.24: Percentages of tool retouch types by level.

C08N C09N CIO CllN C12N C13N C14 C15 TOTAL

N/A 00 00 1. 09 2 17 00 88 .98 .00 .78

Very fine 1 16 1 43 1. 09 2 17 1 37 00 1.96 .00 1.25 or fine Semi- 17 44 22 86 29.35 25 00 10 96 28 95 30.39 25.00 24.33 abrupt Abrupt or 43 02 31 43 38.04 29 35 50 68 37 72 29.41 .00 36.26 backing Mixed ret. 18 60 31 43 14. 13 23 91 21 92 18 42 18 .63 62.50 21.04 types Side 12 79 4 29 10.87 7 61 9 59 7 02 7. 84 12. 50 8.63 scraper Alternate 00 00 .00 00 00 00 1.96 .00 .31

Marginal 6 98 8 57 4.35 9 78 1 37 7 02 8 .82 .00 6.75 or util. Burin blow 00 00 1. 09 00 4 11 00 .00 .00 .63

TOTAL % 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 NUMBER 86 70 92 92 73 114 102 8 537 128

There was little variability in retouch types between levels

(Table 4.24). In general, abrupt retouch occurs frequently (35-55%).

Semi-abrupt retouch (12-25%), and mixed retouch, a combination of semi- abrupt, abrupt, and fine retouch (13-35%), occur in similar proportions.

Here, retouch type is more likely to vary between major tool classes than within them. For example, microliths frequently have short, semi- abrupt or abrupt retouch along a lateral edge or tnmcation whereas side scrapers have long, steep retouch. Except for microliths, major tool classes did not occur in sufficient quantities to enable comparisons of retouch type within one tool class. Therefore, any variability in frequencies and percents of retouch type relate more to the frequency of specific tool categories than technological variability.

Microliths

Microliths comprise the highest proportion of tools (20.22-35.14%) in all levels. In previous analyses, some discemable changes between natural levels 5 and 7 were noted in the proportions of geometric and non-geometric microliths, and differences in the microburin index. To assess the frequency of geometric microliths, two categories of geometric microliths were identified (Clark et al. 1987, 1988; Donaldson

1986; Neeley ec al. 1995). Geometric A microliths represent geometric forms typically associated with Geometric Kebaran assemblages (Henry

1989b). These include "true" geometric forms such as lunates, trapezes, rectangles, and triangles. Geometric B microliths include any microlith that is truncated (oblique or straight) and backed, and are commonly associated with non-geometric Kebaran assemblages (Table 4.25).

For comparative purposes, microliths are grouped into general categories, similar to those used in earlier studies, based on the presence or absence of backing and the angle of truncation (i.e., straight or oblique). These categories are used because they (1) 129 facilitate comparisons with in-field analyses and between levels by increasing the number within each category; and (2) reflect general typological and secondary technological attributes. These techno- typological characteristics are better indicators of variation between levels, as fine typological divisions are more apt to be based on subtle variations in tool morphology, even a specialist may choose different typological categories if asked to reanalyze a microlithic assemblage.

In addition, some morphological differences seem to reflect varying degrees of utilization, a point which will be discussed in greater detail below.

Backed microliths are the most abundant microliths in all levels

(Table 4.25). But, they comprise roughly 5-16% more of the microlith assemblages in lower levels {C12N-C14) than in upper levels (C08N-C11N).

Some variability within specific types of backed microliths exists between levels (Table 4.26). Backed, backed and retouched, backed and notched, and backed and curved microliths exhibit no consistent trends between levels (Table 4.26). However, pointed varieties are consistently more abundant in levels C12N-C14 than in upper levels.

Level CIO has several pointed microliths with backing and/or semi-abrupt retouch along both lateral edges, and straight, backed pointed microliths. Morphological variations of backed and retouched microliths, that do not clearly "fit" into any category, comprise 2-12% of the microlith assemblage. They are most abundeuit in levels CllN,

C12N and C14.

Few non-backed trimcated microliths are present; they comprise less than 6% of the microliths in each level. Backed and truncated microliths comprise slightly higher proportions of the assemblage (2-

10%) than unbacked truncated microliths. In general, backed and truncated microliths are more abxindant in upper levels (C08N, C09N, 130

Table 4.25: Percentages of microliths by level.

C08N C09N CIO CllN C12N C13N C14 C15 TOTAL

Backed/Curved 39 .5 50.0 38 .1 48.0 55.0 66.7 65.7 50.0 52.38 Bladelets Truncated 2 .6 5.0 2.4 16.0 5.0 6.2 2.9 0.0 5.16 Bladelets Oblq. Trune. 13 .1 5.0 26.2 8.0 7.5 4.1 8.6 25.0 11.11 Bladelets Backed Trune. 7 .9 10.0 0.0 8.0 0.0 2.1 8.6 0.0 4.37 Bladelets Backed Oblq. 28. 9 15.0 28.6 16.0 30.0 16.7 11.4 0.0 21.43 Trune. Bldt Bi-trunc. 0 .0 5.0 0.0 4.0 0.0 0.00 2.8 0.0 1.19 Backed Bldt Atypical 0.0 0.0 0.0 0.0 0.0 2.1 0.0 0.0 0.40 Trap/Reet. Atypical 7 .9 5.0 0.0 0.0 0.0 0.0 0.0 0.0 1.19 Triangle Atypical 0.0 5.0 0.0 0.0 2.5 2.1 0.0 25.0 1.59 Lunate Other 0.0 0.0 4.7 0.0 0.0 0.0 0.0 0.0 0.79

TOTAL 100 100 100 100 100 100 100 100 100 N 35 23 36 25 40 48 31 4 242

C08N C09N CIO CllN C12N C13N C14 CIS % Geometries A 7.9 15.0 4.7 4.0 2.5 4.2 2.8 25.0 % Geometries B 44.7 40.0 33.3 28.0 32.5 23.0 22.8 25.0

Geometries A = Backed and bi-truneated bladelets, trapeze/rectangles, triangles and lunates. Geometries B = Geometries A plus backed and truncated bladelets, and backed and obliquely truncated bladelets. 131

CllN) than in lower levels. Unbacked, obliquely truncated microliths are most abundant in levels C08N and CIO. Backed, obliquely truncated microliths are the second most abundcint microlith category in most levels and have high percentages in levels C08N, CIO, and C12N.

Although wide microliths were already known to be associated with increased proportions of geometric microliths from this study, both obliquely truncated, backed microliths and atypical lunates tend to be manufactured on wide blanks and comprise a greater proportion of tools in upper levels (C08N-C11N) than in lower levels (C12N-C14).

"True" geometric forms occur in very low frequencies (less than

8%). As previously indicated, many blade and bladelet tools are

fragments. While small microlithic fragments have been removed from the sample, larger incomplete tools were included in this analysis. Since recognition of geometric forms almost always recjuires a complete specimen, backed and truncated fragments may represent one end of a geometric microlith. Therefore, the proportion of truncated microliths

may be a useful proxy measure for increased proportions of geometries in

very fragmented assemblages.

Analyses of blade/bladelet tool widths indicated a bimodal distribution of widths in both the upper (C08N-C11N) and lower (C12N-

C14) levels. The separation between modes occurs at about 8 mm (Figures

4.10 and 4.11). The majority of backed microliths (84-100%) in all levels are less than 8 tran wide. Only two types of backed microliths

(curved and backed, and backed and retouched varieties) have equal

proportions of microlith widths greater than and less than 8 mm.

Interestingly, 75% or more of curved and backed, and backed and retouched varieties of microliths in levels C08N-C11N are wider than 8 mm. Although some truncated, obliquely truncated, backed and truncated, and backed and obliquely truncated microliths have widths less than 8 132

Table 4.26: Frequencies of specific microlithic t^'pes by level.

C08N C09N CIO CllN C12N C13N C14 C15 TOTAL Backed Backed |4 0 7 0 1 9 l 0l22 Backed & ret. [2 5 l 4 7 6 6 l|32 Backed & notch.\ 2 0 0 0 0 2 0 0|4 Curved & backed{ 4 3 1 3 1 1 5 OjlS Bck.&ret. variaj 2 1 1 3 4 2 4 0|l7 Pointed: 2 edges backed [0 0 3 0 5 3 5 l|l7 Straight backed', 0 1 3 1 4 9 2 0|20 Microperforator I 1 0 0 1 0 0 0 0|2 Trxincated (no backing) Truncated |l l i 4 2 3 1 0|l3 Oblique truncation (no backing) Oblq. trunc. l4 1 5 2 3 2 1 0jl8 Bitriinc .oblq. |l 0 3 0 0 0 0 0|4 Oblq.tr\mc.&ret| 0 0 3 0 0 0 2 1|6 Backed & truncated Bck.& trunc. |3 2 0 2 0 1 3 Ojll Backed & obliquely truncated Bck.&oblq. tr\in| 6 2 5 4 12 8 3 0[40 w/ notch ]5 1 7 0 0 0 1 0|l4 Bi-truncated & backed Bi-trunc,& bckj 0 1 0 l 0 0 1 0l3 Geometries Proto-trapeze |0 0 0 0 0 1 0 0|l Atyp. lunate jO 1 0 0 1 1 0 1|4 Atyp. trianglej 3 1 0 0 0 0 0 0|4 Other |0 0 2 0 0 0 0 0j2

TOTAL NUMBER 38 20 42 25 40 48 35 4 252 133 mm, 50% or more of these microliths in all levels are wider than 8 mm.

Therefore, trxincated and obliquely truncated microliths, including both backed and unbacked forms, comprise the majority of the wide (11-12 mm) microliths in these assemblages. Usually "true" geometries are variations of backed and truncated, and backed and obliquely truncated microliths; so, the predominance of unbacked truncated and unbacked obliquely truncated microliths with wide forms may also indicate an intermediary form between what is considered a non-geometric and a geometric microlith.

In general, microliths are considered interchangeable tool components (e.g., Henry 1989b; Henry and Garrard 1988). An important question is whether variability in microlithic widths between the

"spiky", very narrow microliths (narrower than 8 mm and commonly 3 mm wide) and wide microliths (generally wider than 8 mm) in the Step C lithic assemblages indicate fionctional differences. Such functional differentiation would be useful in determining the type of activities conducted at this site. In addition, it would help explain the modality in the distribution of microlithic tool widths. Some general differences in utilization and/or retouch on the unbacked edge exist between wide and narrow forms in levels C08N-C15. Many wide forms

(i.e., backed and obliquely truncated, and obliquely truncated microliths) are either heavily utilized and/or notched on the unbacked edge. Although many narrow forms (e.g., backed microliths) have retouch on the unbacked edge, very few narrow forms are notched. In addition, the notch on narrow forms is not as deep as the notch on wide forms.

Furthermore, many backed microliths (approximately 30%) are pointed varieties. Narrow pointed varieties, especially backed and retouched microliths, and microliths backed on two edges, may have been more easily hafted in a manner similar to that suggested by Henry and 134

Garrard (1988:12) for mounting points at Tor Hatnar. The notch associated with wide truncated and obliquely trimcated microliths may indicate that these microliths were not hafted at the tip of the haft nor aligned longitudinally along the haft's length. Rather, they may have been hand-held or hafted singly, along the length of a haft. The wide tools with backing on one lateral edge would enable these microliths to be more easily held. It is unlikely that these microliths were hafted in an alignment because the placement of the notch in the center of the unbacked edge suggests somewhat more controlled use of the microlith. I would expect that microliths hafted longitudinally might produce irregular evidence of utilization on the unbacked edge.

Microburin Indices

The microburin technique is a method of trvincating or segmenting a bladelet whereby a bladelet is notched and snapped. The notching serves to control the position of the snap break. Together, the notch and snap produce a characteristic scar on the microburin. Bladelets segmented with this technique are usually an intermediate stage in microlith production. Later, the scar on the segmented bladelet may be removed as further modification (usually trioncation) of the bladelet segment occurs. Microburins, segmented bladelets that retain the characteristic snap scar, are the waste or final end products of this technique (Henry

1974; Tixier 1963, 1974).

Various methods have been proposed to calculate the frequency of the microburin technique (mbt) (Table 4.27). One index (Imbt) reflects the use of the technique in the total tool assemblage (Bar-Yosef 1970) .

The reduced Imbt or rimbt (Henry 1974) and the adjusted Imbt or adjimbt

(Marks and Larson 1977) reflect the use of the technique among all microliths and truncations (pieces where the technique was most likely to have been used). The adjImbt reflects the use of the technique only 135 on those items that probably were produced by the technique.

The number of microburins and microburin indices in each level show a steady increase between levels C08N and C13N (Table 4.28 and

4.29). Clear "jumps" occur in all microburin indices between levels

C08N and C09N, CIO and CllN, C12N and C13N, and C14 and CIS. The rimbt and adj Imbt indicate that the microburin technique was used more in levels C11N-C14 than in levels C08N-C10. This division corresponds with natural levels 5 and 7. These microburin indices are higher than those presented in earlier studies (Clark et al. 1987, 1988; Donaldson 1986).

This difference most likely reflects the greater number of microburins identified in this analysis than in in-field analyses. Only 60 microburins were recorded during in-field analyses of the lithic assemblage from all of natural level 7 (Clark et al. 1987, 1988;

Donaldson 1986). In this study, 8 5 microburins were recorded for only a portion of natural level 7, i.e., the northern half of levels C11-C13, and all of levels C14 and C15.

The use of these indices "as cultural markers assumes that manufacturing debris (microburins) and tools (microliths) were regularly discarded together in consistent frequencies" (Neeley and Barton

1994:278). The high microburin indices in the lower levels are interesting because this technique is not commonly associated with non- geometric assemblages in the early Epipaleolithic. This suggests that this technique is more common in the early Epipaleolithic than previously indicated. The association of low microburin indices and geometric microliths in the upper levels is also found in other

Epipaleolithic assemblages (Henry 1989b). One plausible explanation is that microburin scars were removed when segmented bladelets were truncated during the manufacture of geometric microliths. In this situation, microburin scars would no longer be identifiable. 136

Table 4.27: Formulas for microburin indices.

Imbt = # of microburins # of microburins + tools

rimbt = # of microburins # of microburins + microliths + truncations

adjImbt = # of microburins # of microburins + truncations

Table 4.28: Percentages of microburins.

C08N C09N CIO CllN C12N C13N C14 C15 Total

True 0.0 66.7 75.0 90.9 83 .3 94.2 90.0 100. 0 89.4 iCnikowski 100.0 0.0 0.0 9.1 8.3 2.9 10.0 0.0 5.9

Piquant 0.0 33.3 25.0 0.0 8.3 2.9 0.0 0.0 4.7 tiedre Total % 100 100 100 100 100 100 100 100 100.0 Number 1 3 4 11 12 34 19 1 85

Table 4.29: Microburin indices by level.

C08N C09N CIO CllN C12N C13N C14 C15 Imbt 1.2 5.0 4.8 13.9 18.2 31.2 20.2 12.5 rImbt 1.6 9.1 5.5 22.4 17.1 34.7 28.8 14.3 adjImbt 4.1 23.1 13.3 45.8 40.0 68.0 61.3 33.3 137

Some have suggested that low microburin indices in Geometric

Kebaran assemblages are one response to a need to use lithic material more conservatively (Neeley and Barton 1994:280). This argument is based on the difference between the mecin microlith length and the mean blank length in the Geometric Kebaran, Mushabian, and Natufian industries. It is suggested that the difference is more pronounced in the Geometric Kebaran because two microliths were being generated from each bladelet while in the Mushabian one microlith was manufactured from each bladelet (Neeley and Barton 1994). In the Mushabian, the unused portion of the bladelet was discarded with the characteristic microburin scar. In the Geometric Kebaran, both bladelet segments were manufactured into microliths (Neeley and Barton 1994) .

If this argument is true, I would expect the mean microlith length in the Geometric Kebaran to be 50% or less of the mean bladelet length.

However, Geometric Kebaran microliths account for 61% of the average length of unretouched bladelets (Neeley and Barton 1994:280). In the

Mushabian, backed bladelets account for 80% of the unmodified blank length (Neeley and Barton 1994:280). Neeley and Barton's argument seems to hold for the Mushabian which has about a 1:1 ratio of microburins to microliths (this would give an rlmbt of about 50). However, they seem to argue from negative evidence for the Geometric Kebaran, i.e., fewer microburins in those assemblages. They do not offer an explanation why lithic sources may have been used more conservatively. I think the argument could be turned around to suggest just the opposite conclusion, i.e., lithic material was used more wastefully. The microburin technique may not have been used because it was not as effective for segmenting the wide bladelet blanks used in the manufacture of geometric microliths. Before either interpretation is accepted, additional studies linking changes in land use patterns, subsistence strategies. 138 and conservative or wasteful use of lithic rav.' ^.atsrial CIATQ nscssssiTy'. 139

CHAPTER 5: DISCUSSION AND CONCLUSIONS

In chis study, techno-typological variability in the debitage and tool components of the lithic assemblage from Tor al-Tareeq (WHS 1065),

Step C, arbitrary levels C08N-C15 was analyzed. This variability can be used to determine differences in the chaine operatoire, or operational sequence of lithic manufacture in these levels. Some of the variability identified in these levels is cyclic and appears to represent functional changes in the activities conducted, or at least deposited, in these levels. Most of the cyclic changes relate in some way to location of core reduction activities and the intensiveness of core reduction in each level. Since much of the assemblage is fragmentary, additional analyses were conducted in order to identify relative amounts of depositional and post-depositional breakage and to determine how these breakage patterns might influence the completeness of the assemblage.

Variability in the operational sequence, specifically core reduction, and tool manufacture, use and discard, relates to temporal changes in the lithic assemblage. Three areas of variability in the lithic assemblage (site formation processes, site function, and operational sequence) will be discussed below. Then, comparisons of the operational sequences identified in this study will be compared to others in the region.

Site formation processes

In this study, the frequency of completeness categories, and debitage and debris size categories, and fragment to debitage ratios suggest the amount and size of fragmentary debitage and debris in this assemblage is influenced by the artifact's original size and morphology.

The lowest levels (C12N-C15) have at least 10-20% more very small debris

(lengths less than l cm) and higher percentages of small flakes than the upper levels (CG8N-C11N). Therefore, the narrow debitage and retouched 140 cools (particularly microliths) in the lower levels (C12w-C15) may have been more susceptible to post-depositional stress than wide artifacts in upper levels (C08N-C11N). This may account for the relatively high breakage rates (fragment to debitage ratios) and abundant small debris in the lowest levels (C13N-C15) of this site.

The abundance of small debitage in the lower levels may also be a result of downslope movement from upper areas. Previous analyses of surface and subsurface deposits at WHS 1065 were conducted in order to study Che site boundaries, artifact class frequencies, site function, surface-subsurface congruity, and site formation and disturbance processes (Coinman et al. 1989). Structural and compositional characteristics of a 95% sample of Che site's surface assemblage

(collected in I x 1 m units) were compared to excavated, subsurface deposics from Steps A-I. Results from this study suggest chat much of the spatial patterning at the site can be attributed to downslope movement of materials (Coinman et al. 1989). Dense surface deposits are generally located downslope from dense subsurface deposics. However, it is unlikely that this would completely account for the relatively high percentages of very small debris and flakes in levels C11N-C15. Levels

C11N-C15 are considered to be in situ deposics, an interpretation that is strongly supported by the presence of two hearth features in level

C13. The abundance of small artifacts recovered from lower levels

(C13N-C15) might also be related to the size effect which tends to transport large artifacts up in a deposit (Baker and Schiffer 1975;

Baker 1978). Trampling, in combination with the permeability and texture of the sediments, can also sort surface artifacts into size classes (Coinman ec al. 1989; Gifford 1978:81-83).

Clark (et al. 1988:261) suggests that lower frequencies of shatter relative to debitage on the surface and in Step C, natural level 4 141 indicates that these surfaces may hgve been exposed fcr Icngsr periods of time. The underlying deposits may have been exposed to chemical and physical weathering, and other forms of erosion and disturbance. Since the shatter to debitage ratio was greater than one for all subsurface deposits except Step C, natural level 4, Clark (et al. 1988:261) suggests that this level may have been exposed to the surface for longer periods of time. However, Clark (et al. 1988:261) includes debris

(medial and distal flake fragments) in his shatter category. With a separation of these categories, it appears that different processes may be patterning the proportion of debris and shatter in these assemblages.

Levels C08N, C13N, C14, and C15 have higher fragment to debitage ratios suggesting more breakage in these levels than in other levels. The higher percentages of shatter in the lower levels than in the upper levels may be related either to chemical and physical weathering (Clark et al. 1987, 1988) or to primary core reduction activities (Sullivan and

Rozen 1985) . Therefore, although the size effect, collection strategies, and weathering may influence the size patterning of artifacts in a deposit (Clark et al. 1987, 1988; Coinman et al. 1989) artifact size also affects the susceptibility of artifacts to breakage causing narrow blanks to break into smaller fragments than wide blanks.

Intra-site Functional Variability

Some of the variability identified in levels C08N-C15 represents functional changes in the activities conducted, or at least deposited, at the site. Most of these changes are cyclic, cross-cutting temporal techno-typological variability at the site, and relate in some way to the staging and intensiveness of core reduction. Functional interpretations of artifact assemblages are based on evidence of primary- core reduction and tool manufacture, and general tool type frequencies such as the frequency of large, non-microlithic tools and microlithic 142 tools. Evidence for primary core reduction activities include: core to debitage, tool to debitage, and shatter to debitage ratios, cortex frequencies, and to a lesser extent core weights.

Many initial core reduction activities probably occurred off-site, as evidenced by the relatively low percentages (usually less than 25%) of debitage and cores with some cortex. Level C13N differs from other levels in that approximately 10-15% more of its debitage has some amount of cortex. Debitage with completely cortical dorsal surfaces was very- rare, comprising less than two percent of the entire chipped stone assemblage from this site (Clark et al. 1987, 1988). The frequency of shatter is higher in lower levels (C12N-C15) than in upper levels (C08N-

CllN) suggesting either more primary core reduction (Sullivan and Rozen

1985) or more intensive weathering of the lower than of the upper deposits. However, an increase in the amount of primary core reduction is not supported by the frequency of cortex on debitage and cores in these levels. In this regard, this study and previous studies of this assemblage (i.e., Clark et al. 1987, 1988; Donaldson 1986; Neeley ec al.

1995) produced similar results.

Interpretations of the intensiveness of core reduction are based on debitage to core ratios, the relative frequencies of cores and core trimming elements, and to a lesser extent core weights. Earlier analyses already remarked that the frequency of cores and core trimming elements is very low throughout all excavated portions of WHS 1065

(Clark et al. 1987, 1988). In levels C08N-C15, core and core trimming elements each comprise less than two percent of the assemblage. In previous analyses, debitage to core ratios indicate slightly more emphasis on knapping in Step C, natural levels 5 and 5a than in natural level 7 (Clark et al. 1987, 1988). However, this analysis indicates higher debitage to core ratios in levels C08N and C14 suggesting that 143

Che intensity of core reduction activities v.'as probably not constant at this portion of the site throughout the deposition of natural levels 5 and 7. More intensive core reduction may have occurred in level C08N than in levels C09N-C10, and in level C14 than in levels C11N-C13N.

Although the median core weights are higher in the upper levels (C08N-

ClO) than in the lower levels (C11N-C13N), a notched box plot of core

weights indicates that this difference is not statistically significant.

There is little evidence for change in the frequency of tool

manufacture and the frequency of major tool types in this study.

Generally, retouched tools comprise less than 10% of each assemblage.

The proportion of non-microlithic and microlithic tools varies little

between levels. In the non-microlithic tool category, retouched flakes

and pieces are slightly more abundant in levels C09N-C11N and C14 than

in other levels. While, notches and denticulates are more abundant in

levels C08N-C09N than in lower levels. Still, the proportions of these tool classes are not markedly different between levels. Therefore,

activity differences cannot be inferred from these data. Although the combined results from these analyses do not differ significantly from

earlier studies (i.e., Clark et al. 1987, 1988; Donaldson 1986), this study does show more variability in staging and intensity of core reduction in natural levels 5 and 7 than reported in earlier studies

(i.e., Donaldson 1986; Clark et al. 1987, 1988).

Intra-site Variability in Operational Sequences

The operational sequence for the manufacture of chipped stone

begins with the acquisition of raw material. The majority of artifacts deposited at Tor al-Tareeq are manufactured from fine-grained grayish

brown chert. A very coarse grained, fossiliferous limestone is also present, but is rarely used for tool manufacture. Although compositional analyses have not been conducted on the lithic material. 144

it seems that similar lithic sources were used during all occupations of

the site. Fossiliferous limestone outcrops in several locations in the

immediate vicinity of Tor al-Tareeq and is almost certainly the source

area for this material. Although the ubiquity of chert throughout the

Levant has been questioned, fine-grained chert is found within a

kilometer radius of the site. Therefore, it is unlikely that lithic raw

material was scarce. Thus, from the lithic assemblages analyzed here,

similar lithic procurement strategies were employed throughout the occupation of the site.

Core and platform types, distribution of dorsal flake scars, debitage morphometric data, and flake to blade ratios are used to

determine variability in core reduction strategies. Some technological changes in core reduction occur between levels CIO and CllN, following

the natural levels; others occur between levels CllN and C12N. The

upper levels (C08N-C10) have higher portions of bi-directional blade,

bi-directional flake and multiple platform flake cores than lower levels

(C12N-C15). Dorsal flake scar patterns on debitage from upper levels

(C08N-C11N) show higher frequencies of non-parallel and mixed (parallel

and non-parallel) dorsal flake scars on blades and bladelets, and higher

proportions of flakes with two or more parallel flake scars than lower

levels. This may reflect greater emphasis on manufacturing wide blanks

and higher percentages of bi- or multi-directional cores in these

assemblages. In contrast, cores from the lower levels (C11N-C14) are commonly single platform blade and multiple platform flake and blade cores. Some bi-directional flake cores are also present. The lower

levels have higher frequencies of single platform cores and bladelet

tools with multiple (more than three) parallel flake scars than the

upper levels suggesting a somewhat different manner of blank detachment

may have been employed in the lower levels that enabled multiple blades 145 and bladelets to be removed from a single platform. However, an analysis of platform types showed little variability between levels.

The flake to blade ratios did not change significantly between levels. The slightly higher ratios for the lower levels (C12N-C13N) may be influenced by the classification of debitage and the susceptibility of the narrow bladelet blanks in the lower levels to breakage. Previous analyses suggested a one to one relationship for the manufacture of flake and blade blanks (Clark et al. 1987, 1988). Certainly, the high proportion of flake, and flake and blade cores in this assemblage supports this.

Median widths of unmodified blade and bladelet blanks from upper levels (C08N-C10) were variable and could not be statistically separated into different populations. Median widths from the upper levels were statistically different, however, from the lower levels (C12N-C14). The median widths from level CllN are intermediate between the upper (C08N-

ClO) and lower (C12N-C14) levels and could not be statistically separated from either population, suggesting a change in lithic manufacture towards wide bladelets occurred at this time. Despite the transitional nature of level CllN, the shape of the distribution of unmodified blade and bladelet widths is statistically different from the upper (CIO), but not the lower level (C12N).

The next stage in the operational sequence approach is to identify the primary and secondary technologies used in the manufacture of retouched tools. Evidence for these technologies include blank selection criterion, the frequencies of retouch type, and use of the microburin technique. No significant differences in blank selection criterion for major tool types were noted between levels. In a combined analysis of all levels, approximately one-half of all non-microlithic tools are manufactured on blades suggesting blades and flakes were 146 selected in equal proportions for non-mlcrolichic tool manufacture. The median widths of flake, blade and bladelet tools show similar trends to those already observed in the debitage. Flake tools in the upper levels

{C08N-C11N) have wider, thicker, and longer median values than those in lower levels (C12N-C15) but upper and lower levels cannot be separated into statistically different populations. Blade and bladelet median tool widths, however, are significantly wider (ca. 3-5 mm) in upper

(C08N-C11N) than in lower levels (CI2N-C15). In addition, the shape of the width distribution in upper levels statistically differs from that in lower levels. These data suggest that the trend towards the manufacture of wide tools occurred before the trend towards the manufacture of wide blanks.

Typological differences in the proportion of geometric to non- geometric microliths between natural levels 5 and 7 were already known to exist from in-field analyses (Clark et al. 1987, 1988,- Donaldson

1986). In this study, more geometric microliths are present in levels

C08N-C09N than in all of the lower levels suggesting the trend towards manufacturing geometric microliths occurred slightly later than previous analyses indicated. Therefore, relatively wide blanks and tools were being manufactured before an increase in the percentages of geometric microlith forms occurred.

Secondary technological attributes of tool manufacture are represented in retouch type and use of the microburin technique. In this study, retouch type did not vary significantly within major tool types. Differences in retouch type between levels seems to be influenced by the proportion of major tool classes suggesting that variability in retouch type within a major tool class requires a larger sample than the one used in this analysis.

The frequency of the microburin technique changes markedly between 147

lower levels (C11N-C14) and upper levels (C08N-C10). The microburin

indices reported here are higher than those presented in previous

studies, reflecting the higher number of microburins identified in this analysis. During in-field analyses, only 60 microburins were recorded

in the north and south portions of levels C11-C15. In this study, 85

microburins were recorded in a sample approximately one-half the size of the sample used in earlier studies (calculated by the total number of artifacts recorded during in-field analyses). As a result, microburin

indices in this study indicate that this technique is much more common than earlier studies suggested.

In sum, the important features of the operational sequence used to

manufacture chipped stone debitage and tools in the lower levels (C12N-

C15) includes: (1) frequent use of single platform blade cores,- (2) the

manufacture of flake and blade blanks in similar proportions; (3) the

manufacture of narrow (ca. 7.1-7.4 mm) bladelet blanks; (4) the

manufacture of narrow, retouched microliths which are usually backed or doubly backed and pointed; (5) the manufacture of slightly wide

microliths with more variable widths; and (6) the frequent use of the

microburin technique. In contrast, the operational sequence in the upper levels {C08N-C10) is characterized by (1) higher frequencies of

bi- and multi-directional flake and blade cores; (2) the manufacture of equal proportions of flakes and blades; (3) the manufacture of wide, blade and bladelet blanks; (4) the manufacture of wide, obliquely bi- truncated and obliquely bi-truncated and backed microliths (ca. 12 mm); and (5) the infrequent use of the microburin technique.

The lack of synchrony in the technological trend towards the

manufacture of wide tools and wide blanks suggests that primary lithic technology which is associated with blank production is more resistant to change or conservative than secondary lithic technologies which is 140 associated with tool manufacture. Therefore, prirr.ar'/ lithic technology should be a more reliable indicator of prehistoric culture groups than secondary lithic technology. Since flake to blade ratios are similar in all levels, the techno-typological differences analyzed in this study probably do not reflect different culture groups. Therefore, regional cultural continuity probably exists in this area.

The techno-typological differences between upper and lower levels probably reflect both diachronic and adaptive changes. The wider blanks and tools, and higher frequencies of bi- and multi-direction flake and blade cores in the upper levels may reflect a more variable lithic technology and indicate a decrease in mobility in the upper levels. The removal of wide blanks would exhaust the core more readily than the removal of narrow blanks. Although this trend is not reflected in the debitage to core ratios, the size effect may influence the frequency of cores in these deposits. Also, original core sizes may have varied

between levels which would have enabled more blanks to be manufactured

in some levels. The bi- and multi-directional flake and blade cores and the distribution of dorsal flake scars also suggest that more variability in blank removal exists in the upper levels. Since the upper levels are associated with more mesic climatic conditions than the

lower levels, subsistence resources may have been more abundant and/or diverse enabling a reduction in mobility during the occupation of the upper levels. The wide blanks and tools in the upper levels may also

reflect the need for a slightly more robust, durable blank in mesic

environments.

Regional Comparisons of Operational Sequences

Variability in the proportion of non-microlithic and microlithic tools may relate more to differences in site activities, than to technological differences in lithic manufacturing activities. 149

Technological differences, specifically core tiype^ flake to blade ratios, blade and bladelet dimensions, proportion of geometric and non- geometric microliths, and the use of the microburin technique, are important for identifying similarities in operational sequences between sites. Proportional differences in retouched bladelet types and dimensions are considered diagnostic of the Kebaran, Geometric Kebaran, early Hamran, Qalkhan and Natufian Epipaleolithic Complexes (Bar-Yosef

1984, 1987; Henry 1983, 1986).

The operational sequence identified in Step C, levels C11N-C15 of

WHS 1065 has some technological attributes that resemble those used in other early Epipaleolithic assemblages. Several stratified sites in the eastern Levant have evidence for the use of the microburin technique and are found in the Azraq Basin at Uwaynid 18 (trench 1, upper phase),

Uwaynid 14 (late and middle phases), and Jilat 6 (lower, middle, and upper phases) (Byrd 1980; Garrard et al. 1985, 1986, 1987; Garrard and

Byrd 1992) and in southern Jordan at J405, Wadi Humeima (J406b), J407, and Tor Hamar (J431) (Henry 1989b, 1995) (see Figure 1.1).

The adjusted microburin index for the Azraq Basin sites ranges between 20.8 and 54.9 with a mean of 34.8. Geometric tools comprise a very small percentage (mean 0.5%) of retouched tool assemblages while non-geometric microliths dominate them (mean 84%). Single platform bladelet cores consistently comprise 60% or more of the core sample

(Byrd 1988:259). Non-geometric microliths forms include arched backed, curved pointed pieces, la Mouillah points, and double truncated pieces.

Temporally, there is a typological change in these sites from small, narrow microliths with arched and backed, curved pointed bladelets at

Uwaynid 14 (Middle Phase) and Jilat 6 (Lower Phase) which are dated to ca. 19,800 to 18,400 + 350/250 BP) to stratigraphically later assemblages with longer, thicker, backed bladelets such as robust la 150

Mouillah points or double truncated, backed bladelets (e.g., Uv/a^'nid 14

Upper Phase and Jilat 6 Middle Phase dated to ca. 18,900 to 18,4 00 + 250

BP)(Byrd 1988:260). Metric data on these assemblages have not yet been published.

At WHS 1065, the lithic assemblages from levels C11N-C15 differ from early Azraq assemblages in that they have higher flake to blade ratios, perhaps early evidence for the trend towards high flake to blade ratios and wide microliths, and very few triangular microlithic forms.

The use of the microburin technique is often associated with the manufacture of triangle or pointed microlithic forms (Henry 1995).

Chronometrically, the upper phase of Wadi Jilat 6, dating to ca. 16,700 to 15,470 + 14 0 BP, is contemporaneous with Step C, level C13 at WHS

1065 (Garrard et al. 1994). Although the assemblage from Jilat 6 has a high adjusted microburin index, it differs typologically, especially in the frequency of small asymmetric triangles and microgravette points in the assemblage (Byrd 1988 ; Garrard et al. 1994). However, the higher flake to blade ratios at Jilat 6 are more similar to those found in the basal layers of Step C at WHS 1065. Therefore, based on chronomecric and techno-typological criteria, levels C11N-C15 at WHS 1065 resemble early Epipaleolithic assemblages in the Azraq Basin.

The restricted microburin index for the southern Jordan sites ranges between 20.3 and 50.0 with a mean of 33.3 (Henry 1995) . Cores in these assemblages also have single, unfaceted platforms on wedge shaped cores. Microburins are relatively large compared to those found in other Epipaleolithic industries (Henry 1995:229). Typologically,

Qalkhan points, manufactured by the microburin technique, are unique to

Qalkhan industry sites. Other microliths include narrow (ca. 4.1-8.1 mm) forms with arched backed bladelets, narrow arched backed and pointed bladelets, and la Mouillah points (Henry 1995). While there are many 151 techno-typological similarities between the southern Jordan sites and levels C11N-C15 at WHS 1065, Qalkhan points are not found in the WHS

1065 assemblage.

Henry (1995) considers the Azraq Basin sites and Sabra 3 and Adh

Daman (Schyle and LJerpmann 1988) to be Qalkhan Complex sites based on the present of Qalkhan points and use of the microburin technique.

Sabra 3 and Adh Daman are surface sites and no chronometric dates have been obtained from them (Schyle and Uerpmann 1988). The Azraq Basin sites have many similarities with southern Jordan sites. However,

Qalkhan style points were only identified by Henry (1995) in the Middle

Phases of Jilat 5. In Syria, layers 4-7 at Yabrud (Rust 1950) and three sites in the El-Kowm Oasis (Cauvin ec al. 1979; Cauvin 1981, Cauvin

Coqueugniot 1990) have many techno-typological similarities to Azraq

Basin and southern Jordan sites including narrow, arched backed, pointed bladelets, scalene triangles, la Mouillah points, possible Qalkhan points, and use of the microburin technique (Henry 1995:234). However, the chronological and typological relationships between all of these sites are not clear.

Another southern Jordan site, Wadi Madamagh, also has narrow arched backed and truncated microliths and microburins (Byrd 1994;

Kirkbride 1958) . Although some radiocarbon determinations have been obtained from this site (Schyle and Uerpmann 1988:47-52), it is not clear how these dated deposits relate to Kirkbride's excavation. Henry

(1989b, 1995) suggests that Wadi Madamagh represents a regional variant of the Mushabian. However, this classification does not appear to be completely accepted (e.g., Byrd 1994)

The assemblages from levels C11N-C15 somewhat resemble "classic"

Kebaran assemblages which are characterized by the presence of abundant backed bladelets and very narrow curved microliths (Bar-Yosef 1970) . 152

However, the microburin technique which is present in levels C11N-C14 is not present at Kebaran sites in the western Levant. Some suggest that levels C11N-C15 resemble Wadi Hammeh 26 in the northern part of the Rift

Valley based on the high proportions of non-geometric microliths (Neeley ec a.1. 1995:48). However, the assemblages from Wadi Hammeh 26 do not have evidence for use of the microburin technique at this time (Edwards

1987, 1990). The microburin technique is not consistently used in the western Levant until the Mushabian (ca. 14,000-13,000 BP). The

Mushabian, defined in the Gebel Maghara, Northern Sinai and some other sites in the Negev, is characterized by arched backed bladelets and la

Mouillah points manufactured with the microburin technique (Phillips and

Mintz 1977). Chronometrically, this industry post-dates level C13 at

WHS 1065 by at least 2,500 years. The microburin technique, therefore, appears to be associated with arid and semi-arid environmental adaptations.

The use of the microburin technique in the early Epipaleolithic

(ca. 19,000-15,000 BP) is unique to the eastern Levant. Unfortunately, the only reliable radiocarbon determinations come from the Azraq Basin and Wadi Hasa. Although many techno-typological similarities are found with other sites in southern Jordan, the lack of reliable radiocarbon dates and deeply stratified deposits at many southern sites make it difficult to strongly tie these sites chronologically to others in the region.

Levels C08N-C10 differ significantly from lower levels (C11N-C15)

The appearance of wide obliquely, bi-truncated and backed geometric microliths in C08N-C10 (termed the broad "Hasa" lunate by Neeley et al.

1995) accompanied by the virtual absence of narrow arched/curved backed, backed and pointed narrow microliths, higher percentages of geometric microliths, and little evidence for the use of the microburin technique. 153 indicate different operational sequences for lithic manufacture were used in the upper levels. Unfortunately, these levels are undated but obviously are later stratigraphically than lower levels.

Levels C08N-C10 have more geometric microliths than earlier deposits. In the western Levant, Geometric Kebaran microlithic assemblages are dominated by triangles, trapezes, rectangularly shaped pieces, and low microburin indices. Many regional variants of the

Geometric Kebaran have been identified, e.g., the Falitan (Besancon et al. 1977) and the Hamran (Henry 1983). Bar-Yosef (n.d. cited in Goring-

Morris 1987:18) has subdivided this period into Al, a group represented by trapeze-rectangles (most sites), and A2, those sites represented by triangles (only three sites). The Al grouping has been further subdivided into an early phase with narrow bladelets and a late phase with wide bladelets. There is a tendency for wide bladelets and trapeze-rectangles to dominate in southern sites (Bar-Yosef and Phillips

1977). Although there are some general typological similarities, these

Geometric Kebaran assemblages especially in the Negev and Sinai have much higher percentages of geometric microliths than WHS 1065, levels

C08N-C10 (Marks et al. 1976, 1977, 1978; Goring-Morris 1987; Neeley et al. 1995).

Based solely on the relative frequency of geometric microliths, it has been suggested that levels (C08N-C10) most closely resemble Middle and Late Hamran assemblages, regional variants of the Geometric Kebaran, from southern Jordan (Henry 1995; Neeley et al. 1995:47). Middle Hamran assemblages are characterized by relatively narrow bladelets, relatively narrow trapeze/rectangles which were manufactured without the use of the microburin technique while Late Hamran assemblages have comparatively short bladelets, intentional use of the microburin technique, and the appearance of lunates (Henry 1989b, 1995). In the Final Hamran, 154 trapeze/rectangles are replaced by lunates ar.d there is an increase in the use of the microburin technique (Henry 1989b, 1995). Although there are similarities in the percentages of geometric microliths between

Hamran assemblages and levels C08N-C10 at WHS 1065, the Middle-Final

Hamran has consistent use of the microburin technique and comparatively narrow microlith forms. These techno-typological differences suggest that levels C08N-C10 are not Hamran industry sites. Unfortunately, none of these Hamran assemblages have reliable radiocarbon dates.

Typologically, the wide, obliquely, bi-truncated and obliquely bi- truncated and backed microliths resembles some of the wide, geometric, truncated and/or backed microliths and atypically wide trapezes recovered from phase D at Kharaneh IV in eastern Jordan (Byrd 1994;

Muheisen 1985, 1988; Neeley ec al. 1995). There is also little use of the microburin technique in Phase D at Kharaneh IV. Two apparently reliable radiocarbon determinations of 15,200 + 450 BP and 15,700 + 160

BP have been obtained in Phase D (Byrd 1994:219). Based on techno- typological considerations, the upper levels (C08N-C10) most closely

resemble Kharaneh IV in northeastern Levant, not other variants of the

Geometric Kebaran which are found in southwestern and southeastern

Levantine areas.

Conclusions

Some technological attributes (e.g., flake to blade ratios)

between levels C08N-C15 at WHS 1065 indicate regional cultural continuity in the west-central Levant. However, substantial variability

in the operational sequences used to manufacture chipped stone exists

between the upper (C08N-C10) and the lower levels (C11N-C15). This variability reflects both diachronic and adaptive changes. Not all technological and typological changes between the upper and lower levels

are synchronous. This study suggests that some primary technological 155 change (specifically the manufacture of wide tools) occurred before sctr.e secondary technological change (specifically the manufacture of wide unmodified blanks). Therefore, primary lithic technology should be a more reliable indicator of prehistoric culture groups than secondary lithic technology. The manufacture of wide blanks and tools, and bi- and multi-directional flake and blade cores in the upper levels may indicate a more variable, lithic technology and a decrease in mobility in the upper levels. These technological changes appear to be associated with more mesic climatic conditions at that time. The manufacture of wide blanks and tools in the upper levels may also reflect the need for more robust, durable blanks in mesic environments.

Changes in the frequency of the microburin technique between the upper and the lower levels may also be associated with changing adaptations to different environmental conditions, since use of the microburin technique is frequently associated with arid and semi-arid environmental adaptations.

The assemblages from both the upper (C08N-C10) and the lower levels (C11N-C15) at WHS 1055 seem to most closely resemble contemporary sites in the Azraq Basin. The lower levels (C11N-C15) resemble Uwaynid

18, Uwaynid 14 and Jilat 6 while the upper levels (C08N-C10) most closely resemble Kharaneh IV in northeastern Levant. It would be interesting to determine if the timing of the techno-typological trends identified in this study are also found at other contemporary sites in the eastern Levant, specifically the early Epipaleolithic sites in the

Azraq Basin. Although others suggest similarities to southern and western Levantine sites, these techno-typological and temporal associations are problematic.

Local and regional environmental conditions around Pleistocene

Lake Hasa and the Levant changed from xeric to relatively mesic 156 conditions at about this time. Some of the techno-t^'pclcgical changes and changes in the operational sequence of lithic manufacture at Tor al-

Tareeq reflect general regional trends in lithic technology and appear to be associated with more mesic climatic conditions during the middle

Epipaleolithic. It has not yet been determined to what extent environmental change is related to techno-typological changes in lithic manufacture in other assemblages. Additional studies that more closely link lithic, environmental, and subsistence data are necessary in order to determine if these changes occur in similar paleoenvironmental and temporal contexts. APPENDIX A: WHS 1065 DEBITAGE ANALYSTS CODING LIST

VARIABLE NAME (VARIABLE) UNIT/LEVEL/DIVISION (UNITLEVEL$) N - North S - South B - Not divided # - Feature

BLANK TYPE (BLANK) Flake (Blank Group [BG]) =1) 1 - Flake <1 cm 2 - Flake 1-2 cm 3 - Flake 2-3 cm 4 - Flake >3 cm Blade (BG = 2) 5 - Blade >3cm - IDC 6 - Blade >3cm - 2DC 7 - Bladelet L<3 cm, W < 1.2 cm Core rejuventation (BG = 3) 8 - Core rejuvenation blade 9 - Core rejuventation bladelet 10 - Core rejuvenation flake Microburin (BG = 4) 11 - Regular microburin 12 - Piquant tiedre 13 - Krukowski microburin Burin spall (BG = 5) 14 - Burin spall Shatter (BG = 6) 15 - Shatter Cores (BG = 7 and 8) 16 - Cores ((BG = 7) 17 - Core fragments (BG = 8) •Debris (BG = 9) (medial and distal flake fragments) Tools (BG = 10 and 11) 18 - Tool (BG = 10) 19 - Tool fragment (BG = 11)

CORTEX (CORTEX) 0 - Not applicable 1 - Absent 2 - Present

COMPLETENESS (COMPLETE) 0 - Not Applicable 1 - Complete 2 - Bulb present, distal end absent 3 - Bulb absent, distal end pre^rent 4 - Medial segment

COUNT (NUMBER)

FOR BLADES LENGTH (L) - Measured to the nearest 0.1 millimeter. WIDTH (W) - Measured to the nearest 0.1 millimeter. THICKNESS (TH) - Measured to the nearest 0.1 millimeter. WEIGHT (WT) - Measured to the nearest 0.1 gram. 158

APPENDIX B: WHS 10<55 TOOL AND CORE ANALYSIS CODING LIST

VARIABLE NAME (VARIABLE) UNIT/LEVEL/DIVISION (UNITLEVEL$) N - North S - South B - Not divided # - Feature

BLANK TYPE (BLANK) Blades (L > 3 cm, W > 1.2 cm) 1 - 1st order (Blank group [BG] = 1) 2 - 2nd order (BG = 1) 3 - Core rejuvenation 4 - Crested 5 - Laminar tablet Bladelets (L < 3 cm, W < 1.2 cm) 6 - 1st order (BG = 2) 7 - 2nd order (BG = 2) 8 - Core rejuvenation 9 - Laminar tablet Flajces 10 - Flake 11 - Core rejuvenation 12 - Core tablets Shatter 13 - Shatter Microburins 14 - True microburin 15 - Krukowski microburin 16 - Piquant triedre Burin spalls 17 - Burin spalls Cores 18 - Single platform, blade 19 - Opposed (perpendicular) platform, blade 20 - Single platform, flake 21 - Opposed platform, flake or flake and blade 22 - Multiple platform, flake and blade 23 - Multiple platform, flake 24 - Bipolar blade 25 - Amorphous, exhausted core 26 - Bipolar, flake 27 - Two platforms, not opposed, flake 28 - Two platforms, not opposed, blade 29 - Single platform on large flake, flake 30 - Other

CORTEX (CORTEX) 0 - 0 % No cortex 1 - 100 % Cortex present 2 - 51-99 % Cortex present 3 - 1-50 % Cortex present 4 - Cortex present on platform only TOOL TYPES (TYPE) 0 - Not appliciable Scrapers (Standard Tool Type [ST] = 1 1 - On a flake 2 - On a retouched flake 3 - Rounded or circular 4 - Thumbnail 5 - Transversal 6 - Sidescraper 7 - On a blade or bladelet 8 - On a retouched blade or bladelet 9 - Ogival 10 - Denticulate 11 - Double Carinated (ST = 2) 12 - Shouldered or nosed 13 - Broad carinated 14 - Narrow carinated 15 - Lateral carinated 15 - Core scraper 17 - Double carinated Burins (ST = 3) 18 - Dihedral 19 - Dihedral angle 20 - Angle burin on a break or natural surface 21 - Multiple dihedral burin 22 - On straight truncation 23 - On oblique truncation 24 - On concave truncation 25 - On convex truncation 26 - Multiple on truncation 27 - Multiple on mixed 28 - Beaked 29 - Carinated 30 - Flat faced 31 - Transverse on lateral notch Multiple tools (ST = 4) 32 - Burin/scraper 33 - Other Retouched/backed blades (ST = 5) 34 - Partially retouched 35 - Completely retouched 36 - Helwan blade 37 - Backed knife 38 - Curved backed knife 39 - Retouched/back varia 40 - Backed/retouched fragments (ST = 18) Truncations (ST = 6) 41 - Straight 42 - Concave 43 - Oblique 44 - Backed and truncated or retouched and truncated Microliths (ST = 7) 45 - Backed 46 - Retouched and backed 47 - Backed and notched 48 - Truncated 49 - Oblique truncation 160

50 - Backeci and truncated 51 - Backed and oblique truncation 52 - Bi-truncated and backed 53 - Curved/arched and backed 54 - Curved/arched and backed with basal modification 55 - Pointed - semi-abrupt Sc/or backed retouch on 2 edges 56 - Pointed straight backed microlith (micropoint) 57 - Microperforator 58 - Retouched/backed bladelet varia 59 - Retouched/backed fragment 60 - Humped backed bladelet 61 - Backed obliquely truncated and retouched or notched on oppposite edge 62 - Bitruncated (obliquely) and not backed 63 - Obliquely truncated & retouched Geometries (ST = 8) 64 - Trapeze/rectangle 65 - Proto-trapeze 66 - Trapeze 67 - Assymmetrical trapeze 68 - Trapeze with one convex end 69 - Lunate 70 - Atypical lunate 71 - Isoceles triangle 72 - Atypical triangle 73 - Other Notches and Denticulates (ST = 9) 74 - Retouched notch 75 - Two or more notches 76 - Denticulate Various (ST = 10) 77 - Retouched flake or piece (ST = 11 for complete flake, ST = 16 for flake fragments) 78 - Utilized blank (ST = 10) 79 - Other (ST = 12) 8 0 - Microburin (ST = 15) 81 - Core (ST = 14)

PLACEMENT OF RETOUCH (RET_PLACE) 0 - Not applicable 1 - Lateral obverse 2 - Lateral inverse 3 - Distal obverse (or microburin scar) 4 - Distal inverse (or microburin scar) 5 - Proximal obverse (or microburin scar) 6 - Proximal inverse (or microburin scar) 7 - Obverse/inverse 8 - Distal and lateral obverse 9 - Distal and lateral inverse 10 - Distal/proximal 11 - Lateral inverse and distal obverse 12 - Proximal and lateral inverse 13 - Proximal and lateral obverse 14 - Top ridge 15 - Inverse 161

RETOUCH ORIENTATION (RET_ORIENT) 0 - Noc applicable 1 - Left lateral edge 2 - Right lateral edge 3 - Both lateral edges 4 - Left distal (or microburin scar) 5 - Right distal (or microburin scar) 6 - Left proximal 7 - Right proximal 8 - Complete distal 9 - Lateral and distal 10 - Proximal and distal 11 - Complete proximal 12 - Proximal and lateral 13 - Lateral, proximal, and distal 14 - Top ridge

RETOUCH TYPE (RET_TYPE) 0 - Not applicable 1 - Very fine and fine (ouchata) 2 - Semiabrupt 3 - Abrupt (backing) 4 - Mixed 5 - Side scraper retouch 6 - Bipolar 7 - Inverse 8 - Alternate 9 - Helwan 10 - Marginal (utilized, crushed or battered) 11 - Burin blow

DISTRIBUTION OF FLAKE SCARS(D_FLK_SCAR) 0 - Not applicable/no platform 1 - Cortex/no flake scars 2 - No cortex/flake scars parallel to direction of the blow (I I to dir. of blow) 3 - Two flake scars (11 to dir. of blow) 4 - Multiple flake scars (|| to dir. of blow) 5 - One flake scar (not || to dir. of blow) S - Multiple flake scars (not |1 to dir. of blow) 7 - Mixed flake scars (|j and not jj to dir. of blow)

PLATFORM TYPE (PLATFORM) 0 - Not applicable or missing 1 - Cortex, plain, not prepared 2 - Non-cortical - plain, prepared by single blow 3 - Faceted, several steep splits or facets determines the angle, not the point of percussion 4 - Punctiform (pointed) 5 - Crushed 6 - Broken 7 - Dihedral

SHAPE OF DISTAL END (DISTAL_END) 0 - Not applicable or missing 1 - Pointed 2 - Blunt and thick 3 - Convex/arched 162

4 - Straight 5 - Oblique 6 - Irregular/wavy 7 - Concave 8 - Microburin scar

SHAPE OF LATERAL EDGES (LATERAL_ED) 0 - Not applicable 1 - Straight 2 - Convex 3 - Concave 4 - Irregular 5 - Straight/convex 6 - Straight/concave 7 - Straight/irregular 8 - Convex/concave 9 - Convex/irregular

MATERIAL TYPE (MATERIAL) 1 - Fine grained chert 2 - Coarse grained chert 3 - Mixed - fine and coarse grained chert or with inclusions 4 - Fossiliferous limestone 5 - Other

LENGTH (LENGTH) - Measured to the nearest O.l millimeter.

WIDTH (WIDTH) - Measured to the nearest 0.1 millimeter.

THICKNESS (THICKNESS) - Measured to the nearest 0.1 millimeter.

WEIGHT (WEIGHT) - Measured to the nearest O.l gram.

COMMENTS (COMMENTS $) 163

REFERENCES

Anderson-Geraud, P.C. 1983 A consideration of the uses of certain backed and "lustered" stone tools from late Mesolithic and Natufian levels of Abu Hureyra and (Syria). In Traces d'Utilization sur les Outlis Neolithiques du Proche Orient, edited by M.-C. Cauvin, pp. 77-106. Travaux de la Maison de 1'Orient, no. 5. Maison de I'Orient Mediterranean, Lyon.

Aurenche, 0., J. Cauvin, M.-C. Cauvin, L. Copeland, F. Hours, and P. Sanlaville 1981 Chronologie et organization de I'espace dans le Proche Orient de 12000 a 5600 avant J.-C. In Prehistoire de Levant, edited by J. Cauvin and P. Sanlaville, pp. 571-602. Colloques Internationaux de CNRS 598, Paris.

Banning, E.B., R.R. Dods, J. Field, I. Kuiut, J. McCorriston, J. Siggers, H. Taani and J, Triggs 1992 Tabaqat al-Bumma: 1990 excavations at a Kebaran and Late Neolithic site in Wadi Ziqlab. Annual of the Department of Antiquities 36:43-62.

Baker, C. 1978 The size effect: an explanation of variability in surface artifact assemblage content. American Antiquity 43:288-293.

Baker, C., and M.B. Schiffer 1975 Archaeological evidence for the size effect. In Arkansas Eastman Archaeological Project, edited by C. Baker, pp. 117- 122. Arkansas Archaeological Survey, Research Report No. 6. Arkansas Archaeological Society, Fayetteville, Arkansas. Baruch, U. 1994 The Late Quaternary pollen record of the Near East. In Late Quaternary Chronology and Palaeoclimates of the Eastern Mediterranean, edited by 0. Bar-Yosef and R.S. Kra, pp. 103- 120. Supplement to Radiocarbon, Department of Geosciences, University of Arizona, Tucson, Arizona.

Baruch, U., and S. Bottema 1991 Palynological evidence for climatic changes in the Levant ca. 17,000-9,000 B.P. In The Natufian Culture in the Levant, edited by 0. Bar-Yosef and F. Valla, pp. 11-20. International Monographs in , Ann Arbor, Michigan. Bar-Yosef, 0. 1970 The Epipaleolithic Cultures of Palestine. Unpublished Ph.D. dissertation. Department of Archaeology, Hebrew University, .

1975 The Epipaleolithic in Palestine and Sinai. In Problems in Prehistory: North Africa and the Levant, edited by F. Wendorf and A.E. Marks, pp. 313-371. SMU Press, Dallas.

1977 Section A, Mushabi IV. In Prehistoric Investigations in the Gebel Maghara, Northern Sinai, edited by 0. Bar-Yosef and J.L. Phillips, pp. 184-193. Qedem 7. Monographs of the 164

Institute of Archaeology, Hebrew University, Jerusalem.

1981 The Epipaleolithic complexes in the southern Levant. In Prehistoire de Levant, edited by J. Cauvin and P. Sanlaville, pp. 389-408. Colloques Internationaux de CNRS 598, Paris.

1987 Late Pleistocene adaptations in the Levant. In The Pleistocene Old World: Regional Perspectives, edited by 0. Soffer, pp. 219-236. Plenum Press, New York.

1990 The last glacial maximum in the Mediterranean Levant. In The World at 18,000 BP, edited by C. Gamble and 0. Soffer, pp. 58-77. Unwin Hyman Ltd., London.

1991a Stone tools and social context in Levantine prehistory. In Perspectives on the Past: Theoretical Biases in Mediterranean Hunter-Gatherer Research, edited by G.A. Clark., pp. 371-395. University of Pennsylvania Press, Philadelphia.

1991b The search for lithic variability among Levantine Epi- Palaeolithic industries. In 25 Ans D'Etudes Technologiques en Prehistoire, pp. 319-355. XI Rencontres Internationales d'Archeologie et d'Histoire d'Antibes. Editions APDC, Juan- Les-Pins.

Bar-Yosef, O., and A. Belfer-Cohen 1977 The Lagaman industry. In Prehistoric Investigations in Gebel Maghara, Northern Sinai, edited by 0. Bar-Yosef and J.L. Phillips, pp. 42-88. Qedem 7. Monographs of the Institute of Archaeology, Hebrew University, Jerusalem.

1989 The origins of sedentism and farming communities. Journal of World Prehistory 3(4):447-498.

1991 From sedentary hunter-gatherers to territorial farmers in the Levant. In Between Bands and States, edited by S. Gregg, pp. 181-202. Center for Archaeological Investigations Occasional Paper No. 9. Southern Illinois University at Carbondale, Illinois.

Bar-Yosef, 0., and J.L. Phillips (editors) 1977 Prehistoric Investigations in Gebel Maghara, Northern Sinai, edited by 0. Bar-Yosef and J.L. Phillips. Qedem 7. Monographs of the Institute of Archaeology, Hebrew University, Jerusalem.

Bar-Yosef, O., and A. Killebrew 1985 Wadi Sayakh - a Geometric Kebaran site in Southern Sinai. Paleorient 10:95-102.

Bar-Yosef, 0., and A.N. Goring-Morris 1977 Geometric Kebaran A occurrences. In Prehistoric Investigations in Gebel Maghara, Northern Sinai, edited by 0. Bar-Yosef and J.L. Phillips, pp. 115-148. Qedem 7. Monographs of the Institute of Archaeology, Hebrew University, Jerusalem. 165

Bar-Yosef, 0., and J.C. Vogel 1987 Relative and absolute chronology of the Epi-Paleolithic in the Southern Levant. In Chronologies in the Near Ease, edited by 0. Aurenche, J. Evin and F. Hours, pp. 219-245. BAR International Series 379, Oxford.

Begin, Z.B., A. Ehrlich, and Y. Nathan 1974 Lake Lisan, the Pleistocene precursor of the Dead Sea. Bulletin of Geological Survey, Min. Comm. Ind., State of Israel, No. 63:30.

Begin, Z.B., W. Broeker, B. Buchbinder, Y. Druckman, A. Kaufman, M. Magaritz, and D. Neev 1985 Dead Sea and Lake Lisan levels in the last 30,000 years: a preliminary report. Geological Survey of Israel Report 29:1-30.

Belfer-Cohen, A., and P. Goldberg 1982 An Upper Palaeolithic site in South Central Sinai. Israel Exploration Journal 32:185-189.

Besancon J., L. Copeland, and F. Hours 1975-7 Tableaux de prehistoire Libanaise. Paleorient 3:5-45.

Binford, L.R. 1973 Interassemblage variability - the Mousterian and the "functional" argument. In The Explanation of Culture Change, edited by C. Renfrew, pp. 227-254. Duckworth, London.

1983 Working at Archaeology. Academic Press, New York.

Binford, L.R., and S.R. Binford 1966 A preliminary analysis of functional variability in the Mousterian of Levallois facies. American Anthropologist 68 ( :2) :238-295.

Bordes, F. 1961 Mousterian cultures in France. Science 134 (3482):803-810.

1973 On the chronology and contemporaneity of different Paleolithic cultures in France. In The Explanation of Culture Change, edited by C. Renfrew, pp. 217-226. Duckworth, London.

1978 Typological variability in the Mousterian layers at Peche de I'Aze I, II, and IV. Journal of Anthropological Research 34 (2) : 181-193.

Bottema, S. 1987 Chronology and climatic phases in the Near East from 16,000- 10,000 B.P. In Chronologies in the Near East. Relative Chronologies and Absolute Chronology 16,000-4,000 B.P., edited by 0. Aurenche, J. Evin, and F. Hours, pp. 295-310. BAR International Series 379, Oxford.

Bottema, S., and W. Van Zeist 1981 Palynological evidence for the climatic history of the Near 166

Ease 50.000-6,000 B.P. In Prshlstclrs ds Levant, edited by J. Cauvin and P. Sanlaville, pp. 111-132. Colloques Internationaux de CNRS 598, Paris.

Buchbinder B., Z.B. Begin and G.M. Friedmand 1974 Pleistocene algal tufa of Lake Lisan, Dead Sea area, Israel. Israel Journal Earth Science 23:131-138.

Buller, H. 1983 Methodological problems in the microwear analysis of tools selected from the Natufian sites of el Wad and Ain Mallaha. In Trace d'Utilization sur les Outlis Neolithiques due Proche Orient, edited by M. Cauvin, pp. 107-126. Travaux de la Maison de 1'Orient, Maison de 1' Orient, Lyon.

Byrd, B.F. 1987 Beidha and the Natufian: Variability in Levantine Settlement and Subsistence. Unpublished Ph.D. dissertation, Department of Anthropology, University of Arizona, Tucson, Arizona.

1988 Late Pleistocene settlement diversity in the Azraq Basin. Paleorient 14:257-264.

1990 Late Pleistocene assemblage diversity in the Azraq Basin. Paleorient 14 (2) :257-265 .

1994 Late Quaternary hunter-gatherer complexes in the Levant between 20,000 and 10,000 BP. In Late Quaternary Chronology and Paleocliwates of the Eastern Mediterranean, edited by 0. Bar-Yosef and R.S. Kra, pp. 205-226. Supplement to Radiocarbon, Department of Geosciences, University of Arizona, Tucson, Arizona.

Byrd, B.F., and A.N. Garrard 1990 The last glacial maximum in the Jordanian desert. In The World at 18,000 BP, Vol. 1, edited by C. Gamble and 0. Soffer, pp. 78-96. Unwin Hyman Ltd., London.

Cauvin, J., M.-C. Cauvin, and 0. Stordeur 1979 Recherches prehistorique a El Kowm (Syrie), premiere campagne 1978. Chiers de I'Euphrate 2:80-117. CNRS, Paris.

Cauvin, M.-C. 1981 L'Epipaleolithique du Levant. In Prehistoire du Levant, edited by J. Cauvin and P. Sanlaville, pp. 439-444. Editions du CNRS, Paris.

Cauvin, M.-C., and E. Coqueugniot 1990 L'oasis d'El Kowm et le Kebarxen Geometrique. In Paleorient 14(2) :270-2ai.

Childe, V. Gordon 193 9 Man Makes Himself. Oxford University Press, New York.

Clark, G.A. 1989 Romancing the stones: biases, style and lithics at La Riera. In Alternative Approaches to Lithic Analysis, edited 167

by D.O. Henry and G.K. Cdell, pp. 27-50. Archeological Papers of the American Anthropological Association Number 1.

1991 Paradigms, realism, adaptation, and evolution. In Perspectives on the Past: Theoretical Biases in Mediterranean Hunter-Gatherer Research, edited by G.A. Clark, pp. 441-439. University of Pennsylvania Press, Philadelphia.

1992 Wadi al'Hasa Paleolithic settlement patterns: Negev and South Jordan models compared. In Studies in the History and Archaeology of Jordan, Vol. IV, edited by M. Zaghloul et al., pp. 89-96. Department of Antiquities, Amman, Jordan.

Clark, G.A., N.R. Coinman, J. Lindly, and M. Donaldson 1987 A diachronic study of Paleolithic and aceramic Neolithic settlement patterns in the Wadi Hasa, West-central Jordan. In Studies in the History and Archaeology of Jordan III, edited by A. Hadidi, pp. 215-223. Department of Antiquities, Amman, Jordan and Routledge & Kegan Paul, London.

Clark, G.A., J. Lindly, M. Donaldson, A. Garrard, N. Coinman, J. Schuldenrein, S. Fish, and D. Olszewski 1987 Paleolithic archaeology in the southern Levant: a preliminary report of excavations at Middle, Upper, and Epipaleolithic sites in Wadi El-Hasa, West-central Jordan. Annual of the Department of Antiquities of Jordan 31:19-78.

1988 Excavations at Middle, Upper and Epipaleolithic sites in the Wadi al-Hasa, West-central Jordan. In The Prehistory of Jordan, edited by A.N. Garrard and H.G. Gebel, pp. 209-285. BAR International Series 396, Oxford.

Clark, G.A., M. Neeley, B. MacDonald, J. Schuldenrein, and K. 'Amr 1992 Wadi al-Hasa Paleolithic project 1992: preliminary report. Annual of the Department of Antiquities of Jordan 36:13-21.

Clark, J.G.D. 1954 Excavations at Star Carr.- An Early Mesolithic Site at Seasmer near Scarborough, Yorkshire. Cambridge University Press, Cambridge, England.

1975-77 Interpretation of from ancient Egyptian and other sources, part 2. Paleorient 3:127-150.

Clark, J.G.D., J.L. Phillips, and P.S. Stanley 1974 Interpretation of prehistoric technology from ancient Egyptian and other sources, part l. Paleorient 2:323-388.

Clarke, D.L. 1968 Analytical Archaeology: Collected Papers of David L. Clarke, edited by his colleagues. Academic Press, London.

1976 Mesolithic Europe: the economic basis. In Problems in Economic and Social Archaeology, edited by G. Sleveking, I.H. Longworth and K.E. Wilson, pp. 449-481. Duckworth, 163

London.

1979 Analytical Archaeology. Mechuen & Co. Ltd., London.

Close, A.E. 1978 The identification of style in lithic artifacts. World Archaeology 10:223-237.

1989 Identifying style in stone artifacts: a case study from the Nile Valley. In Alternative Approaches to Lithic Analysis, edited by D.O. Henry and G.H. Odell, pp. 3-26. Archeological Papers of the American Anthropological Association Number 1.

Coinman, N.R., G.A. Clark, and M.L. Donaldson 1989 Aspects of structure in an Epipaleolithic occupation site in West-central Jordan. In Alternative Approaches to Lithic Analysis, edited by D.O. Henry and G.H. Odell, pp. 213-235. Archeological Papers of the American Anthropological Association Number 1.

Copeland, L., and C. Vita-Finzi 1978 Archaeological dating of geological deposits in Jordan. Levant 10:10-25.

Curwen, W. 1941 Some food gathering implements: a study of Mesolithic tradition. Antiquity 15:320-336.

Darmon, F. 1987 Analyses polliniques de trois sites Natoufiens (Ancien, Recent, Final) dans la Region de Salibiya-Fazael. Paleorient 13:121-129.

Dibble, H.L. 1984 Interpreting typological variation of Middle Paleolithic scrapers: function, style, or sequence of reduction? Journal of Field Archaeology 11(4):431-436.

1985 Raw material variability in Levallois flake manufacture. Current Anthropology 26(3):391-393.

1987 The interpretation of Middle Paleolithic scrapers morphology. American Antiquity 52:109-117. Donaldson, M.L. 1986 Preliminary analysis of lithics from WHS site 1065: an Epipaleolithic site in the Wadi Hasa, Jordan. Unpublished seminar paper. University of Arizona, Tucson, Arizona.

1991 Historic biases in modern perceptions of the Levantine Epipaleolithic. In Perspectives on the Past: Theoretical Biases in Mediterranean Hunter-Gatherer Research, edited by G.A. Clark, pp. 341-352. University of Pennsylvania Press, Philadelphia. 169

Donaldson, M.L., and G.A. Clark 1906 WHS site 1065: a Geometric Kebaran site in the Wadi Hasa (West-central Jordan). Unpublished manuscript, Arizona State University, Tempe.

Druckman, Y., M. Magaritz, and A. Sneh 1987 The shrinking of Lake Lisan, as reflected by the diagenesis of its marginal oolithic deposits. Israel Journal of Earth Sciences 36:101-106.

Edwards, P.C. 1987 Late Pleistocene occupation in the Wadi al-Hammeh, Jordan Valley. Unpublished Ph.D. dissertation, University of Sydney.

1990 Kebaran occupation at the last glacial maximum in Wadi al- Hammeh, Jordan Valley. In The World at 18,000 BP, Vol. 2, edited by C. Gamble and 0. Soffer, pp. 97-118. Unwin Hyman Ltd., London.

Edwards, P., S. Bourke, S. Colledge, J. Head, and P. Macumber 1988 Late Pleistocene prehistory in the Wadi el-Hammeh, Jordan Valley. In The Prehistory of Jordan. The State of Research in 1986, edited by A.N. Garrard and H.G. Gebel, pp. 525-565. BAR International Series 396, Oxford.

Ewing, J.F. 1949 The treasures of Ksar Akil. Thought 24:255-288.

Ferring, C.R. 1977 The late Upper Paleolithic site of Ein Aqev East. In Prehistory and Paleoenvironments in the Central Negev, Isreal. Volume II: The Advat Aqev Area, Part 2 and the Harif, edited by A.E. Marks, pp. 81-118. SMU Press, Dallas.

1988 Technological change in the Upper Paleolithic of the Negev. In Upper Pleistocene Prehistory of Western Eurasia, edited by H. Dibble and A. Montet-White, pp. 333-348. University Museum, Philadelphia.

Garrard, A.N., D. Baird, and B.F. Byrd 1994 The chronological basis and significance of the Late Paleolithic and Neolithic sequence in the Azraq Basin, Jordan. In Late Quaternary Chronology and Paleoclimates of the Eastern Mediterranean, edited by 0. Bar-Yosef and R.S. Kra, pp. 177-200. Supplement to Radiocarbon, Department of Geosciences, University of Arizona, Tucson, Arizona.

Garrard, A.N., A. Betts, B.F. Byrd, and C. Hunt 1987 Prehistoric environment and settlement in the Azraq Basin: an interim report on the 1985 excavation season. Levant 19:5-25.

Garrard, A.N., A. Betts, B.F. Byrd, S. Colledge, and C. Hunt 1988a Summary of paleoenvironmental and prehistoric investigations in the Azraq Basin. In The Prehistory of Jordan; The State of Research in 1986, edited by A.N. Garrard and H.G. Gebel, pp. 311-337. BAR International Series 396, Oxford. 170

Garrard, A.N.. and B.F. Byrd 1992 New dimensions to the Epipaleolithic of the Wadi el-Jilat in Central Jordan. Paleorient 18(l):47-62.

Garrard, A.N., B.F. Byrd, and A. Betts 1986 Prehistoric environment and settlement in the Azraq Basin, an interim report on the 1984 excavation season. Levant 18 :5-24 .

Garrard, A.N., B.F. Byrd, P. Harvey, and F. Hivernel 1985 Prehistoric environment and settlement in the Azraq Basin, a report on the 1982 survey season. Levant 17:1-28.

Garrard, A.N., and Hans Georg Gebel (editors) 1988 The Prehistory of Jordan: The State of Research in 1986. BAR International Series 396, Oxford.

Garrard, A.M., S. Colledge, C. Hunt and R. Montague 1988b Late Quaternary environmental history during the late Pleistocene and early Holocene in the Azraq Basin. Paleorient 14(2):40-49.

Garrod, D.A.E. 1932 A new mesolithic industry: the Natufian of Palestine. Journal of the Royal Anthropological Society 62:257-269.

1942 Excavations at the cave of Shukbah, Palestine, 1928. Proceedings of the Prehistoric Society 8:1-20.

1954 Excavations at the Mugharet El-Kebara, Mount Carmel, 1931: The Aurignacian industries. Proceedings of the Prehistoric Society 43:155-192.

1958 The Natufian culture: the life and economy of a Mesolithic people in the Near East. Proceedings of the British Academy 43:155-192.

Garrod, D.A.E., and D.M.A. Bate 1937 The Stone Age of Mount Carmel, Vol. 1. Claredon Press, Oxford.

Gat, J., and M. Magaritz 1980 Climatic variations in the eastern Mediterranean area. Naturwissenschaften 67:80-87.

Gifford, D. 1978 Ethnoarchaeological observations of natural processes affecting cultural materials. In Explorations in Ethnoarchaeology, edited by R. Gould, pp. 51-62. Academic Press, New York.

Gilead, I., and 0. Marder 1989 Geometric Kebaran sites in the Nahal Rut area, western Negev, Israel. Paleorient 15:123-137.

Gilead, I. 1901a The Upper Paleolithic tool assemblages from the Negev and Sinai. In Prehistorie du Levant: Chronologie et 171

1'organization de I'espace depuis Iss crlginss jiisqu'3.iL. No. 598, pp. 331-342. Maison de L'Orient, Lyon.

1981b The Upper Paleolithic in Sinai and the Negev. Unpublished Ph.D. dissertation, Hebrew University, Jerusalem.

1983 Upper Paleolithic occurrences in Sinai and the transition to the Epi-Paleolithic. Paleorienc 9:30-53.

1984 Is the term "Epipaleolithic" relevant to Levantine prehistory? Current Anthropology 25)(2):227-228.

1988 The Upper Paleolithic to Epi-PaleoliChic transition in the Levant. Paleorient 14(2): 177-182.

1991 The Upper Paleolithic period in the Levant. Journal of World Prehistory 5(2):105-154.

Goldberg, P. 1976 Upper Pleistocene geology of the Advat/Agev area. In Prehistory and Paleoenviroaments in the Central Negev, Israel, Vol. I, edited by A.E. Marks, pp. 25-56. SMU Press, Dallas.

1981 Late Quaternary stratigraphy of Israel: an eclectic view. In Prehistoire du Levant, edited by J. Cauvin and P. Sanlaville, pp. 55-66. CNRS, Paris.

1986 Late Quaternary environmental history of the southern Levant. Geoarchaeology 1(3):225-244.

Goldberg, P., and 0. Bar-Yosef 1982 Environmental and archaeological evidence for climatic change in the southern Levant and adjacent areas. British Archaeological Reports 3-133:399-414.

Goldberg, P., and B. Brimer 1983 Late Pleistocene geomorphic surfaces and environmental history of Avdat/Havarim area, Mahal Zin. In Prehistory and Paleoenvironments in the Central Negev, Israel, Vol. Ill, edited by A.E. Marks, pp. 1-13. SMU Press, Dallas.

Goodfriend, G., and M. Magaritz 1988 Paleosols and Late Pleistocene rainfall fluctuations in the Negev Desert. Nature 322:144-146.

Goring-Morris, A.N. 1987 At the Edge: Terminal Pleistocene Hunter-Gatherers in the Negev and Sinai. BAR International Series 361, Oxford.

1995 Complex hunter/gatherers at the end of the Paleolithic (20,000-10,000 BP). In The Archaeology of Society in the Holy Land, edited by Thomas E. Levy, pp. 141-168. Leicester University Press, London.

Gould, R.A., D.A. Koster, and A.H.L. Sontz 1971 The lithic assemblage of the Western Desert Aborigines of Australia. American Antiquity 36:41-66. 172

Hayden. B. 1977 Stone tool functions in the Western Desert. In Scone Tools as Cultural Markers: Change, Evolution, Complexity, edited by R.V.S. Wright, pp. 178-188. Australian Institute of Aborigine Study, Canberra, Australia.

1979 Snap, shatter and super-fractures: use wear of stone skin scrapers. In Lithic Use-VIear Analysis, edited by B. Hayden, pp. 207-230. Academic Press, New York.

Henry, D.O. 1973 The Natufian of Palestine: its material culture and ecology. Ph.D. dissertation. Southern Methodist University, University Microfilms, Ann Arbor.

1974 The utilization of the microburin technique in the Levant. Paleorient 2(2) :389-395.

1977 An examination of the artifact variability in the Natufian of Palestine. In Eretz-Israel (Vol. 13), edited by B. Arensburg and 0. Bar-Yosef, pp. 229-239. Hebrew University, Jerusalem.

1982 The prehistory of Southern Jordan and relationships with the Levant. Journal of Field Archaeology 9(4):417-444.

1983 Adaptive evolution within the Epipaleolithic of the Near East. In Advances in World Archaeology Volume 2, edited by F. Wendorf and A. Close, pp. 99-160. Academic Press, New York.

1986 The prehistory and paleoenvironments of Jordan: an overview. Paleorient 12(2):5-26.

1987 Topographic influences on Epipaleolithic land use patterns in Southern Jordan. In Studies in the History and Archaeology of Jordan III, edited by A. Hadidi, pp. 21-27. Department of Antiquities, Amman, Jordan and Routledge & Kegan Paul, London.

1988 The Epipaleolithic sequence within the Ras en Naqb - el Quweira area. Southern Jordan. Paleorient 14:245-256.

1989a Correlations between reduction strategies and settlement patterns. In Alternative Approaches to Lithic Analysis, edited by D.O. Henry and G.H. Odell, pp. 13 9-155. Archeological Papers of the American Anthropological Association Number 1.

1989b From Foraging to Agriculture: the Levant at the End of the Ice Age. University of Pennsylvania Press, Philadelphia.

1994 Prehistoric cultural ecology in Southern Jordan. Science 265:336-341.

1995 Prehistoric Cultural Ecology and Evolution: Insights from Southern Jordan. Plenum Press, New York. 173

HenrVj D.O., and A.N. Garrard 1988 Tor Hamar: an Epipaleolithic rockshelter in Southern Jordan. Palestine Exploration Quarterly 120:1-25.

Horowitz, A. 1976 Late Quaternary paleoenvironments of prehistoric settlements in the Avdat/Aqev area. In Prehistory and Paleoenvironments in the Central Negev, Israel I The Avdat/Aqev Area, edited by A.E. Marks, pp. 57-68. SMU Press, Dallas.

1979 The Quaternary of Israel. Academic Press, New York.

1989 Prehistoric cultures of Israel: correlation with the oxygen isotope scale. In Investigations in the Prehistory in South Levantine Prehistory, edited by 0. Bar-Yosef and B. Vandermeersch, pp. 5-18. BAR International Series 497, Oxford.

Hours, F. 1974 Remarques sur I'utilisation de listes types pour 1'etude du Paleolithique superieur et de 1'Epipaleolithique du Levant. Paleorient 2:3-18.

1976 L'Epipaleolithiquie au Liban. Resultats acquis en 1975. Preprint from Xleme Congres de I'Union Internationale des Sciences Prehistoriques et Prothistoriques, pp. 106-131. Nice.

Hovers, E., and 0. Marder 1991 Typo-chronology and absolute dating of the Kebaran Complex: implications from the second season of excavations at Urkan E-Rubb Ila. Journal of the Israel Prehistoric Society - Mitekufat Haeven. 24:34-58.

Issar, A., and H. Bruins 1983 Special climatological conditions in the deserts of Sinai and the Negev during the latest Pleistocene. Palaeogeography, Paleoclimatology, Palaeecology 43:63-72.

Issar, A., et al. 1980 The character of the climate in the Negev and Sinai between 40,000 and 20,000 Years B.P. In Bat Sheva Seminar on Approaches and Methods in Paleoclimatic Research, pp. 48-49 (abstract). Weizmann Institute, Rehovot.

Katsnelson, J. 1964 The variability of annual precipitation in Palestine. Arch. Meteorological Bioki B13:163-172.

Kaufman, A., Y. Yechieli, and M. Gardosh 1992 Reevaluation of the lake-sediment chronology in the Dead Sea Basin, Israel, based on new ^^"Th/U dates. Quaternary- Research 38:292-304.

Kelley, L.H. 1980 Experimental Determination of Stone Tool Uses, A Microwear Analysis. University of Chicago Press, Chicago. 174

Kenyon. K. 1959 Earliest Jericho. Antiquity 33:5-9.

Kirkbride, D. 1958 A Kebaran rockshelter in Wadi Madamagh, near Petra, Jordan. Man 63:55-58.

1966 Five seasons at the Pre-Pottery Neolithic village of Beidha in Jordan. Palestine Exploration Quarterly 98:8-72.

Kuhn, S.L. 1994 A formal approach to the design and assembly of mobile toolkits. American Antiquity 59(3):426-442.

Lemmonnier, P. 1983 L'etude des systemes techniques, une urgence de technologie culturelle. Techniques et Culture, nouvelle serie 1:11-34.

1990 Topsy turvy techniques: remarks on the social representation of techniques. Archaeological Review from Cambridge 9:27-37.

Leroi-Gourhan, A., and F. Darmon 1987 Analyses palynoloqiques de sites archeologiques de Pleistocene Final dans la Vallee du Jourdain. Israel Journal of Earth Sciences 36:65-72.

Lewarch, D., and M. O'Brien 1981 The expanding role of the surface assemblages in archaeological research. In Advances in Archaeological Method and Theory, Vol. 4, edited by M.3. Schiffer, pp. 297- 342. Academic Press, New York.

Lieberman, D.E. 1993 The rise and fall of seasonal mobility among hunter- gatherers : the case of the southern Levant. Current Anthropology 34:599-631.

Lubbock, J. 1865 Pre-historic Times, as Illustrated by Ancient Remains, and the Manners and Customs of Modem Savages. Williams and Norgate, London.

MacDonald, B. (editor) 1988 The Wadi el-Hasa Survey, 1979-1983, West-central Jordan. Wilfrid Laurier University Press, Waterloo, Ontario.

MacDonald, B., G. Rollefson, and D. Roller 1982 The Wadi el-Hasa Survey 1981: a preliminary report. Annual of the Department of Antiquities of Jordan 26:117-132.

MacDonald, B., E. Banning, and L. Pavlish 1980 The Wadi el-Hasa Survey 1979: a preliminary report. Annual of the Department of Antiquities of Jordan 24:169-184.

MacDonald, B., G. Rollefson, E. Banning, B. Byrd, and C. d'Annibale 1983 The Wadi el-Hasa Survey 1982: a preliminary report. Annual of the Department of Antiquities of Jordan 27:311-323. 175

Macumber.. P.G.; and M.J. Head 1991 Implications of the Wadi al-Hammeh sectuences for the terminal drying of Lake Lisan, Jordan. Palaeogeograpby, Palaeoclimatology, Palaeoecology 84:163-173.

Magaritz, M. 1986 Environmental changes recorded in the Upper Pleistocene along the desert boundary, southern Israel. Palaeogeograpby, Palaeoclimatology, Palaeoecology 53:213- 219.

Marks, A.E. (editor) 1976 Prehistory and Paleoenvironments in the Central Negev, Isreal. Volume I: The Advat Aqev Area, Part 1. SMU Press, Dallas.

1977 Prehistory and Paleoenvironments in the Central Negev, Isreal. Volume II: The Advat Aqev Area, Part 2 and the Harif. SMU Press, Dallas.

198 3 Prehistory and Paleoenvironments in the Central Negev, Isreal. Volume III: The Advat Aqev Area, Part 3. SMU Press, Dallas.

Marks, A.E., and D.A. Friedel 1977 Prehistoric settlement patterns in the Avdat/Aqev area. In Prehistory and Paleoenvironment in the Central Negev, Israel, Vol. II, edited by A.E. Marks, pp. 191-232. Sr^U Press, Dallas.

Marks, A.E., and P.A. Larson Jr. 1977 Test excavations at the Natufian site of Rosh Horesha. In Prehistory and Paleoenvironment in the Central Negev, Israel, Vol. Ill, edited by A.E. Marks, pp. 293-316. SMU Press, Dallas.

Marks, A.E., and A.H. Simmons 1977 The Negev Kebaran of the Har Harif. In Prehistory and Paleoenvironments in the Central Negev, Israel, Volume II: The Advat Aqev Area, Part 2 and the Harif, pp. 233-270. SMU Press, Dallas.

Marks, A.E., H. Crew, R. Ferring, and J. Phillips 1972 Prehistoric sites near Har Harif. Israel Exploration Journal 22(2-3) :73-85.

Marks, A.E., and P. Volkman 1983 Changing core reduction strategies: a technological shift from Middle to the Upper Paleolithic in the Southern Levant. In The Mousterian Legacy. Human Biocultural Change in the Upper Pleistocene, edited by E. Trinkaus. British Archaeological Reports, Oxford.

Mellars, P.A. 1969 The chronology of Mousterian industries in the Perigord region. Proceedings of the Prehistoric Society 35:134-171. 176

Moore.. A. M . T. 1985 The development of Neolithic societies in the Near East. In Advances in World Archaeology, Vol. 4, edited by F. Wendorf and A.E. Close, pp. 1-69. Academic Press, New York.

Mortensen, P. 1970 A preliminary study of the chipped stone industry from Beidha. Acta Archaeologica 41:1-54.

1972 Seasonal camps and early villages in the Zagros. In Man, Settlement and Urbanism, edited by P. Ucko, R. Tringham, and G.W. Dimbleby, pp. 293-298. Duckworth and Co., London.

Muheisan, M. 1983 La Prehistoire en Jourdanie, Recherches sur 1'Epipaleolithique, L'example de gisement Kharaneh IV. Unpublished Ph.D. dissertation, Universite de Bordeaux, Bordeaux.

1985 L'Epipaleolithique dans le gisement de Kharaneh IV. Paleorient 11(2):149-160.

1988 The Epipaleolithic phases of Kharaneh IV. In The Prehistory of Jordan, edited by A.N. Garrard and H.G. Gebel, pp. 353- 367. BAR International Series 396, Oxford.

Neeley, M.P., and C.M. Barton 1994 A new approach to interpreting Late Pleistocene microlith industries in Southwest Asia. Antiquity 68:275-288.

Neeley, M.P., J.D. Peterson, G.A. Clark, and S.K. Fish 1995 WHS 1065 (Tor al-Tareeq), an Epipaleolithic site in the Wadi el-Hasa, West-central Jordan. Unpublished manuscript. On file in the Department of Anthropology, Arizona State University, Tempe.

Neev, 0., and Emery K.O. 1967 The Dead Sea: depositional processes and environments of evaporites. Geologic Survey of Israel Bulletin 41:1-147.

Neuville, R. 1934 Le prehistorique de Palestine. Revue Bihlique 43:237.

1951 Le Paleolitbique et le Mesolithique du desert de Judee. Archives de 1'Institut de Paleontologie Humaine 24. Masson, Paris.

Newcomer, M.H. 1974 Study and replication of bone points from Ksar Akil (Lebanon). World Archaeology 6:138-153.

Nissen, K., and M. Dittemore. 1974 Ethnographic data and wear pattern analysis: a study of socketed Eskimo scrapers. Tebiwa 17:67-88.

Olszewski, D.I. 1989 Tool blank selection, debitage and cores from Abu Hureyra 1, Northern Syria. Paleorient 15(2):29-37. 177

Olszewski. D.T., M. Stevens, M. Glass, P..F. Heck, J. Cccper and G.A. Clark 1994 The 1993 excavations at Yutil al-Hasa (WHS 784), an Upper/Epipaleolithic site in West-central Jordan. Paleorieat 20:129-141.

Perrot, J. 1955 Le Paleolithique Superieur d'El-Quseir et de Masaraq an Na'aj. Bulletin de la Societe Prehistorique Fraxicaise. 52(8):493-506.

1962 Palestine-Syria Celicia. In Courses Toward Urban Life, edited by R.J. Braidwood and G.R. Willey, pp. 147-164. Aldine, Chicago.

1966 Le gisement Natufian de Mallaha CEyan), Israel. L'Anthropologie 70:437-484.

Phillips, J.L. 1991 Refitting, edge-wear and chaines operatoires: a case study from Sinai. In 25 Ans D'Etudes Technologiques en Prehistoirs. XI Rencontres Internationales d'Archeologie et d'Histoire d'Antibes, pp. 303-317. Editions APDC, Juan-Les- Pins.

1994 The Upper Paleolithic chronology of the Levant and the Nile Valley. In Late Quaternary Chronology and Paleoclimates of the Eastern Mediterranean, edited by 0. Bar-Yosef and R.S. Kra, pp. 169-176. Supplement to Radiocarbon, Department of Geosciences, University of Arizona, Tucson, Arizona.

Phillips, J.L., and B.G. Gladfelter 1991 The refitting of lithics and site formation processes in the Upper Paleolithic of Southern Sinai. Paper presented at the annual meeting of the Society for American Archaeology, New Orleans.

Phillips, J.L., and Mintz, E 1977 The Mushabian, In Investigations in Gebel Maghara, Northern Sinai, pp. 149-183. Qedem 7. Monographs of the Institute of Archaeology, Hebrew University, Jerusalem.

Phillips, J.L., and Yerkes 1979 Unpublished artifact master coding sheet (revised).

Rathje, W., and M.B. Schiffer 1982 Archaeology. Harcourt Brace Jovanovich, New York.

Rolland, N. 1977 New aspects of Middle Palaeolithic variability in western Europe. Wature 266(5599):251-252.

1981 The interpretation of Middle Paleolithic variability. Man 16(1):15-42.

Rosenan, N. 1970 Climatic regions. Atlas of Israel. Sheet IV/3. 178

Rossingol, M. 1969 CJne sequence climacique du Pleistocene dans la region de la Mer Morte (Israel). Pollen and Spores XI(3):604-614.

Rust, A. 1950 Die hohlenfunde von Jabrud (Syrian). Karl Wachholtz Verlag, Neumunster.

Schiffer, M.B. 1976 Behavioral Archaeology. Academic Press, New York.

Schuldenrein, J. 1987 Paleolithic archaeology in the Southern Levant: a preliminary report of excavations at Middle, Upper, and Epipaleolithic sites in Wadi El-Hasa, West-central Jordan. Annual of the Department of Antiquities of Jrdan 31:68-70.

Schuldenrein, J., and G.A. Clark 1994 Landscape and prehistoric chronology of West-central Jordan. Geoarchaeology 9(1): 31-55.

Schuldenrein, J., and P. Goldberg 1981 Late quaternary paleoenvironments and prehistoric site distributions in the lower Jordan Valley: a preliminary report. Paleorienc 7(l):57-7l.

Schyle, D., and H.P. Uerpmann 1988 Palaeolithic sites in the Petra-area. In The Prehistory of Jordan: The State of Research in 1986, edited by A.N. Garrard and H.G. Gebel, pp. 39-65. BAR International Series 396, Oxford.

Seminov, S.A. 1964 Prehistoric Technology. Adams and Mackay, Cory, England.

Shehadeh, N. 1985 The climate of Jordan in the past and present. In Studies in the History and Archaeology of Jordan II, edited by A. Hadidi, pp. 25-38. Department of Antiquities, Amman, Jordan and Routledge & Kegan Paul, London.

Shott, M. 1986 Technological organization and settlement mobility: an ethnographic examination. Journal of Anthropological Research 42:15-51.

Stekelis, M., and G. Haas 1942 Preliminary report on soundings in prehistoric caves in Palestine (Mugharet al-Watwat; Mugharet an-Nugtah; Mugharet Wadi-Fallah; Mugharet Abu-Usba; el-Barud). Bulletin of the Awercian Schools of Oriental Research 86:2-10.

1952 The Abu-Usba Cave. Israel Exploration Journal 13:1-12.

Stekelis, M., and T. Yizraeli 1963 Excavations at Nahal Oren: a preliminary report. Israel Exploration Journal 13:1-2. 179

Sullivan. A.P., and K.C. Rozen 1985 Debitage analysis and archaeological interpretation. Awericem Antiquity 50(4):755-779

Tchernov, E. 1976 Some Late Pleistocene faunal remains from the Avdat/Aqev area. In Prehistory and Paleoenvironments in the Central Negev, Israel, Volume I: The Avdat/Aqev Area, edited by A.E. Marks, pp. 69-74. SMU Press, Dallas.

Tixier, J. 1963 Typologie de I'Epipaleolithic du Maghreb. Memoires du Centre de Recherches Anthropologiques. Prehistoriques et Ethnologiques, Alger, Paris.

1974 Glossary for the description of stone tools with special reference to the Epipaleolithic of the Maghreb. Newsletter of Lithic Technology. Special Publication 1.

Turville-Petre, F. 1932 Excavations in the Mugharet el Kebarah. Journal of the Royal Anthropological Institute 62:270-276.

Valla, F.R. 1988a Epipaleolithic syntheses. Paleorient 14 (2) :316-320.

1988b La fin de 1'Epipaleolithique au Levant: Les industries a microlithes geometriques. L'Anthropologie 92(3):901-925.

Vita-Finzi, C. 1964 Observations on the Late Quaternary of Jordan. Palestine Exploration Quarterly 96:19-31.

1966 The Hasa Formation: an alluvial deposition in Jordan. Man 1:387-390.

1982 The prehistory and history of the Jordanian landscape. In Studies in the History and Archaeology of Jordan, Vol. I, edited by A. Hadidi, pp. 23-27.. Department of Antiquities, Amman, Jordan.

Wendorf, F. 1968 The Prehistory of Nubia, 2 Volumes. SMU Press, Dallas.

Wiessner, P. 1983 Style and social information in Kalahari San projectile points. American Antiquity 48:253-276.

Wobst, M. 1977 Stylistic behaviour and information exchange. In Papers for the Director, edited by C. Cleland, pp. 317-342. Anthropological Papers No. 61. University of Michigan Museum of Anthropology, Ann Arbor.

Yechieli, Y., M. Magaritz, Y. Levy, U. Weber, U. Kafri, W. Woelfli, and G. Bonani 1993 Late Quaternary geological history of the Dead Sea area, Israel. Quaternary Research 39:59-74.