'Isaac went out to the field': Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead 9781784918293, 9781784918309

‘Isaac went out to the field (Genesis 24:63)’ is a collection of 28 articles by 47 authors from research institutions in

386 21 48MB

English Pages 402 [422] Year 2019

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Cover
Title Page Hebrew
Title Page
Contents Page
List of Figures
About the Editors and Contributors
Preface and Acknowledgments
Isaac Gilead: An Appreciation
Shamir Yona on behalf of the editors
Nissim Amzallag
The Religious Dimension of Copper Metallurgy in the Southern Levant
Neolithic Cult Sites in the Southern Negev, Israel
U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel
Figure 1. Eastern ‘Uvda Valley, remains of an elongated cell with an in-situ triad of perforated maṣṣeboth, up to 32 cm high.
Figure 2. Map of Naḥal Roded area with the recorded Rodedian sites, in relation to lithology.
Figure 3. Map of currently recorded Rodedian sites in the Negev, with regions’ names mentioned in the text.
Figure 4. Har Assa, a typical pair of installations, the elongated cell points to the circle (pavement was covered by 1-3 cm of dust and arkose).
Figure 5. Har Argaman, plan of a typical pair of installations.
Figure 6. ‘Uvda Valley, a large elongated cell with maṣṣebah, pointing to a barely preserved circle.
Figure 8. Naḥal Roded, a pair of limestone maṣṣeboth, found tumbled and broken.
Figure 11. Naḥal Roded, perforated and naturally holed limestones, found scattered in and around the installations (recorded first as found).
Figure 9. Naḥal Roded, a triad of maṣṣeboth in a circular cell, as found (left) and after re-setting the central one.
Figure 12. Naḥal Roded, three examples of anthropomorphic stones: 1. Unshaped 2. With a hammered neck 3. Entirely hammered.
Figure 13. Unique stone images: 1. Eastern ‘Uvda Valley, a natural broad ‘female’-look cobble. 2. Nahal Paran, a broad ‘female’-look cobble partially shaped. 3. Eastern ‘Uvda Valley, basalt figurine entirely hammered.
Figure 15. Naḥal Roded, three examples of stones with elongated perforation, the middle one bears an engraving of meander or a snake.
Figure 16. Examples of stone bowls: 1. Naḥal Roded, natural bowl. 2. Naḥal Roded, a complete small bowl and fragments of large one. 3. Naḥal Botem, a well shaped bowl.
Figure 17. Naḥal Roded, a ‘vase-shaped’ installation, as found and after cleaning.
Figure 18. Naḥal Roded, a ‘vase-shaped’ installation with a head-down stone image: left- As found, middle- after cleaning, right- with minor restoration.
Figure 20. Naḥal Demama, two miniature houses built on the end of an elongated cell.
Figure 21. Naḥal Roded, perforated maṣṣeboth set with the perforation down, originally set into the ground to the level of the light/dark line.
Figure 22. Naḥal Roded, regular and perforated buried maṣṣeboth: 1. As found. 2. As originally set before being covered.
Figure 23. Naḥal Roded, a perforated buried maṣṣebah: 1. As found. 2. As originally set, before being covered.
Figure 24. Naḥal Roded, a natural, anthropomorphic copper nodule (two views).
Figure 25. Naḥal Roded, flint blades typical for the PPNB, an ‘Amuq arrow-head (center), on the and a Byblos arrow-head from a site on Har ‘Eshet (central ‘Araba, on the right).
Figure 26. Naḥal Roded, ad-hoc tools, typical for desert sites of the 6th-3rd millennia BC.
Figure 27. Naḥal Roded Site 110, radiocarbon dates from an ash spot. By L. Cummings, the PaleoResearch Institute, Golden Colorado. Calibrated with OxCal 4.2 (Ramsey 2013). Mean dates (red lines) are only approximate, based on the dominant peak in the cu
Figure 28. Jebel Ḥashem alTaref, eastern Sinai, a perforated maṣṣebah among a group of animal ‘stone-drawings’.
Figure 29. Naḥal Roded, a vertical view of installation with a stone alignment, 6.5 m long, connecting to a circle, and a limestone image on the right (enlarged on the lower right corner).
Uzi Avner, Moti Shem-Tov, Lior Enmar, Gideon Ragolski, Rachamim Shem-Tov, and Omry Barzilai
A Basket-Handled Teapot from Fazael 2
S. Bar: A Basket-Handled Teapot from Fazael 2
Figure 1. The location of the crushed basket-handled teapot near an entrance to a room (S. Bar).
Figure 2. The Fazael 2 basket-handled teapot.
Shay Bar
O. Bar-Yosef and F. Valla: The Contributions of Early French Scholars to Levantine Prehistory
Figure 1. Father Germer-Durand.
Figure 2. Father A. Mallon.
Figure 3a. R. Neuville when incorporated as a diplomat.
Figure 3b. R. Neuville in Jerusalem as the French Consul in Israel.
Ofer Bar-Yosef and François Valla
The Contributions of Early French Scholars to Levantine Prehistory
A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev
Figure 1. Map of the central Negev, with Nahal Neqarot south of Maktesh Ramon marked.
Figure 2. Nahal Neqarot rockshelter from the southeast. The Ramonian occupation layer underlies the massive fallen blocks.
Figure 4. Close-up of the two excavation areas and drawn section of grid squares Q/R along the 9 line. Note that the occupation layer is located directly on bedrock and is overlain by the massive roof collapse.
Figure 5. Radiocarbon dates from Nahal Neqarot.
Figure 6. Pestle and other groundstone tools from Nahal Neqarot. Both scales in cm.
Figure 7. Core types and raw materials at Nahal Neqarot.
Figure 11. Microburin indices at Nahal Neqarot.
Figure 8. Core types and raw materials by level at Nahal Neqarot.
Figure 9. Debitage counts at Nahal Neqarot.
Figure 12. Observations on microburins by level at Nahal Neqarot.
Figure 13. Scans of Microburins.
Figure 14. Types of notch on microburins by level at Nahal Neqarot.
Figure 15. Distal shapes of microburins by level at Nahal Neqarot.
Figure 16. Scans of Ramon points at Nahal Neqarot.
Figure 17. Typological counts by level at Nahal Neqarot (Type list after Goring-Morris 1987).
Figure 18. Scans of endscrapers on thick blades at Nahal Neqarot – note ‘Aurignacian-type’ retouch.
Figure 19. Thick endscrapers with ‘Aurignacian-type’ retouch at Nahal Neqarot.
Figure 20. Tools from Nahal Neqarot – burin and scrapers.
Figure 21. Tools from Nahal Neqarot – Ramon points, denticulates and chisel/retoucher.
Figure 22. Mean metric attributes of Ramon points by level at Nahal Neqarot (length, width and thickness in mm).
Figure 23. Retouch/backing types on Ramon points by level at Nahal Neqarot (frequencies in parentheses).
Figure 24. Shape of backed edge of Ramon points at Nahal Neqarot (frequencies in parentheses).
Figure 25. Location of Ramon point tips, retouch related to tip, evidence for presence of microburin scar (in parentheses) at Nahal Neqarot.
Figure 26. Correlation between between Ramon point and microburin alignments attributes at Nahal Neqarot.
Anna Belfer-Cohen and A. Nigel Goring-Morris
The Riddle of the ‘Aurignacian’ in the Negev: The Lithic Assemblage from Nahal Neqarot in the Central Negev, Israel
E. Braun: Forging A link: Evidence for a ‘Lost Horizon’
Figure 1. Map of southern Levant with principal LC to EB 1 transition deposits.
Figure 2. Radiocarbon dates from the Ashqelon Cluster of sites.
Figure 3. A plan of the excavation locales at Barnea, based on Golani and Nagar 2011, Fig. 7.1.
Figure 4. Aerial view of sand dune covering a portion of the Afridar neighborhood ca. 1960. Photo courtesy of Professor Ram Gophna’s personal archive.
Figure 5. Inspector for the Israel Department of Antiquities and Museums, Dov Meiron, stands well below the modern surface in a bulldozer trench cut into a kurkar ridge in the Afridar neighborhood of Ashqelon, 1968. Note the modern ground level next to t
Figure 6. Afridar, Area G during excavation. Note the modern bulldozed surface (‘ground zero’) and low-lying excavated trenches in which early EB 1 remains were found at the base of a truncated kurkar prominence. Photo courtesy of the Israel Antiquities
Figure 7. View facing south from the excavation area of Barnea of the newly flattened littoral. The high-rise buildings in the distance are at the border of the Afridar neighborhood. Numbers 1-3 indicate three locales excavated at the Barnea site. Note r
Figure 10. A partial view of the Barnea locale during excavation, facing south-southwest. Note remains of a kurkar prominence above ‘ground zero’ (Fig. 6) in the center background. The buildings of Afridar are visible in the left background below the shad
Figure 8. A view of the bulldozed littoral from the Barnea locale during excavation, facing north. Note hills in the background, remains of the natural topography of the region.
Figure 9. Remains of a kurkar prominence at the northwest extent of ‘ground zero’, Barnea. Beyond it is the sea. The numbers in the foreground mark different excavation locales. Details of archaeological deposits in it are documented in Figs. 12-14.
Figure 11. View of a bulldozer sectioned segment of a kurkar ridge that parallels the beach at the western border of the Barnea excavation locale (see also Fig. 4), It is covered by grass growing in a layer of sand, below which is a thick layer of soil (t
Figure 12. View of a segment of a bulldozer section at the western edge of the excavation area in Barnea (see Fig. 12). The arrow points to what was apparently a complete pot, in situ, sectioned by the bulldozer.
Figure 13. Closeup view of an in situ pot sheared off by a bulldozer in an archaeological deposit ca. 1.5 m above ‘ground zero’ in Barnea.
Figure 15 Pottery and a limestone vessel (n) from tumuli superimposed on ladder-like cist burials at Adeimeh, Jordan. New renderings after Stekelis 1935: Fig. 19.
Figure 16. An early Chalcolithic vessel in the Amman Jordan Citadel Museum, probably from the Adeimeh Cemetery (photo by the author; no scale but approximately 15 cm high).
Figure 17. A jar burial in Barnea with remains of a wall of kurkar stone just below the modern bulldozed surface; photo by the author, courtesy of A. Golani, Israel Antiquities Authority.
Figure 18. Select Pottery from the Ashqelon Cluster and Chalcolithic Sites: 1. Straight-sided bowl (so-called ‘v-shaped’) from Afridar, Area J, after Baumgarten 2004: Fig. 10:1. 2. Straight-sided bowl (so-called ‘v-shaped’) from Afridar, Area J, after Bau
Figure 19. Select Pottery from the Ashqelon Cluster and Chalcolithic Sites: 1. Jar rim from Afridar Area J, after Baumgarten 2004: Fig. 16:3. 2. Jar rim from Afridar Area J, after Baumgarten 2004: Fig. 10:7. 3. Incised, decorated jar neck from Afridar, Ar
Figure 20. Chalcolithic ground stone vessel fragments from the Ashqelon Cluster and Grar: 1. From Afridar Area E, after Rowan 2004: Fig. 4:3. 2. From Grar, after Gilead 1995: Fig. 7.2:9. 3. From Afridar Area E, after Rowan 2004: Fig. 4:2. 4. From Grar, af
Figure 21. Select Pottery Types from Jebel Mutawwaq: 1. After PAJM photo posted 2014: 7/4/14. 2. After Álvarez et al. 2014: Fig. 11. 3. After PAJM photo posted 2014: 10/3/14. 4. After Álvarez et al. 2014: Fig. 12. 5. After Álvarez et al. 2014: Fig. 11. 6.
Eliot Braun
Forging A link: Evidence for a ‘Lost Horizon’ – The Late Chalcolithic to EB 1 Transition in the Southern Levant
A. Gopher: Unresolved Pottery Neolithic Chrono-Stratigraphic and Crhrono-Cultural issues
Figure 1. Schematic timeline and geography of PN cultures.
Avi Gopher
Unresolved Pottery Neolithic Chrono-Stratigraphic and Crhrono-Cultural issues: comments on the beginning and the end of the Pottery Neolithic period
Going through customs: changing rituals of the Ghassulian culture of the Southern Levant, ca. 4500-3900 BC
M. Gošić: Going through customs: changing rituals of the Ghassulian culture of the Southern Levant
Figure 1. 14C dates from Gilat and Teleilat Ghassul (Bourke et al. 2001; Bourke et al. 2004; Levy and Burton 2006; Weinstein 1984). Calibrated by OxCal v4.2.4 (Bronk Ramsey et al. 2013, Reimer et al. 2013) and shown here with 2σ probability range.
Figure 2. Plan of the temple in Area E at Teleilat Ghassul, classic courtyard phase. Courtesy of Peta Seaton (reproduced from Seaton 2008, Plate 8).
Figure 3. 14C dates from Horvat Beter, Bir es-Safadi, Shiqmim and Giv’at ha-Oranim (Burton and Levy 2011, 179; Gilead 1994, 2)(Carmi and Boaretto 2004). Calibrated by OxCal v4.2.4 (Bronk Ramsey et al. 2013, Reimer et al. 2013) and shown here with 2σ proba
Figure 4. 14C dates from Horvat Qarqar South, Nahal Qanah, Shoham (North) and Peqi’in (Carmi 1998, Carmi and Segal 2005; Segal et al. 1998; Shalem, Gal and Smithline 2013, 413-415). Calibrated by OxCal v4.2.4 (Bronk Ramsey et al. 2013, Reimer et al. 2013)
Figure 5. Selected ossuaries from Peqi’in: A. female B. male C. zoomorphic. Courtesy of Israel Antiquates Authority and Dina Shalem, Zvi Gal and Howard Smithline.
Figure 6. 14C dates from Nahal Mishmar cave of the Treasure (Bar-Adon 1980, 199; Davidovich 2008, Table 3). Calibrated by OxCal v4.2.4 (Bronk Ramsey et al. 2013, Reimer et al. 2013) and shown here with 2σ probability range.
Figure 7. Selected Ghassulian copper artifacts: A. anthropomorphic standard (redrawn after Bar-Adon 1980: 49, no. 21), B. ‘crown’ featuring architectural and zoomorphic motifs (redrawn after Bar-Adon 1980: 28, no. 7), C. standard featuring ibexes and moti
Milena Gošić
Animal offerings from the Nawamis fields and Coeval Habitation Sites in Southern Sinai
L. Kolska Horwitz: Animal offerings from the Nawamis fields and Coeval Habitation Sites
Figure 2. Photograph showing a section of the nawamis field at Ein Huderah (Photograph: Uzi Avner).
Figure 3. Map showing location of the nine nawamis in southern Sinai excavated by Israeli teams (based on Bar-Yosef Mayer 2002: Fig. 1).
Figure 4. Collapsed nawamis photographed ca. 1990-1920 by G. Eric and Edith Matson. Open access print, Library of Congress, Prints & Photographs Division, Photograph Collection – Reproduction number: LC-DIG-matpc-01976.
Table 2. Faunal remains from the nawamis tombs (given as NISP counts).
Table 3. Identifications of faunal species from the nawamis habitation sites.
Table 4. Species list for the EBII Southern Sinai faunal assemblages.
Liora Kolska Horwitz
An Early Pottery Neolithic (Jericho IX) site east of Tel Nagila
H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) site east of Tel Nagila
Figure 1. Location map.
Figure 3. Main section showing the main archaeological deposits.
Figure 4. Archaeological layer overlaid the sterile loess.
Figure 5. Plan of the excavation squares and the main loci.
Figure 6. Pit 101 in Sq. X-3.
Figure 7. Photo of pit 102 with molded mud-bricks.
Figure 8. Plan of Sq. Y-9 including L 206.
Figure 9. Two phases of Living surfaces in Area B.
Figure 10. Walls 2007 and 208 in Sq. X-6, 7.
Figure 11. Cores. 1-4 single platform cores.
Figure 12. Cores. 1 Amorphous core; 2-3 two platform cores.
Figure 13. Tools. 1-2 Arrowheads; 3-4 Sickle blades; 5-6 fan-scrapers.
Figure 14. Tools. 1, 4 Awls; 2 denticulation; 3 massive borer; 4 bifacial in preparation.
Figure 15. Bowls.
Figure 16. Jars.
Figure 17. Handles and Bases.
Figure 18. Photomicrograph of sample 3001.
Figure 19. 1-2 Grinding slab; 3-4 Mobil cup-mark.
Figure 20. 1-2 Grooved stones.
Figure 21. Limestone artifacts. 1-2 abraders; 3-4 cores.
Figure 22. Limestone artifacts. 1-3 massive scrapers.
Figure 23. Limestone artifacts. 1-2 side-scrapers; 3 awls; 4 denticulated artifacts.
Table 7. Frequencies of knapped limestone tools.
Table 8. Dimensions of several limestone categories.
Hamoudi Khalaily, Anastasia Shapiro and Ofer Marder
Clothes Maketh Man: Textile Production in the Southern Levant in the Chalcolithic Period
J. Levy: Clothes Maketh Man: Textile Production in the Southern Levant in the Chalcolithic Period
Figure 1. Map of main sites mentioned in the chapter, courtesy of Isaac Gilead.
Figure 2. Harvesting flax, author’s garden, near the Beersheva Valley, March 2013.
Figure 3. Flax stem anatomy A=cuticle, B=epidermis, C=cortex, D=fibre bundles, E=woody core, F=pith, G=lumen after Baines 1989 Fig. 2, courtesy of Yuval Shach.
Figure 4. Supported on the thigh spinning, Navajo, from Wilson 1896 PL.22.
Figure 5. Drop spinning, high whorl spindle, Iran after Hochberg 1980, 63 courtesy of Yuval Shach.
Figure 6. Horizontal ground loom on ceramic bowl Badari, Egypt from Brunton and Caton-Thompson 1928 PL. XLVIII/70k.
Figure 7. Representative spinning whorls from the Southern Levant (1) photography by author.
Figure 8. Representative spinning whorls from the Southern Levant (2) photography by author.
Figure 9. Spinning bowl, Neve Ur from Perrot, Zori and Reich 1967, photography by Alter Fogel.
Figure 10. Schematic of linen shroud, Cave of the Warrior, after Schick 1998 Fig. 3.2, courtesy of Yuval Shach.
Figure 11. Schematic of weft fringe, shroud, Cave of the Warrior, after Schick 1998 Fig. 3.26 courtesy of Natanel Levy.
Figure 13. Sash, detail of countered weft twining and warp tassels, Cave of the Warrior, from Schick 1998 PL. 3.9, courtesy of Orit Shamir and the Israel Antiquities Authority, photography by Clara Amit.
Figure 15. Fresco, The Procession, Teleilat Ghassul after Cameron 1981, frontispiece, courtesy of Yuval Shach.
Figure 16. Leather sandals, Cave of the Warrior, from Schick 2003, 16 courtesy of the Israel Antiquities Authority, photography by Olga Negnevitsky.
Figure 17. Fresco, The Notables, Teleilat Ghassul from Mallon et al. 1934 PL. 66.
Figure 18. Fresco, The Notables, Teleilat Ghassul, detail of footwear, from Mallon et al. 1934 PL. 56.
Janet Levy
Atar Livneh
Stylistic Devices and Exegetical Techniques in ‘Rewritten Bible’ Compositions
Conus ornaments from Tel Bareqet in an Early Bronze Age Near East context
D. E. Bar-Yosef Mayer et al.: Conus ornaments from Tel Bareqet in an Early Bronze Age Near East context
Figure 1: The Conus apex beads from Bareqet. Each bead is viewed from two sides. Photography and layout: Oz Rittner.
Daniella E. Bar-Yosef Mayer, Sarit Paz and Yitzhak Paz
D. Nadel et al.: Flint knapping in a brush hut
Figure 1. a) Location map.
Figure 1. b) plan of the Ohalo II site.
Figure 2. The distribution map of the flint assemblage (debitage and tools) on the floor of Brush Hut 2 (N=4,536). Note the two distinct clusters and the many low-density units around the wall.
Figure 3. The distribution maps of a) blade/lets (N=2,410) and cores; b) primary elements (N=463) and cores; c) core trimming elements (N=182) and cores.
Figure 4. The distribution maps of a) tools (N=132) and b) microliths (N=60).
Figure 5. The distribution map of the minute remains (N=2,943).
Figure 6. Spatially explicit regression results for pairs of flint products. a) primary elements – core trimming elements; b) blade/lets – core trimming elements; c) blade/lets – primary elements; d) blade/lets – flakes; e) tools – blade/lets; f) tools
Figure 7. The spatial distribution of the raw material 2 refits: a) Set #2 with the fragments composing a flake, Set #2a and four sets of fragments; b) sets *2b-*2d are refitted fragments.
Figure 8. The spatial distribution of the raw material 3 refits: Set #3 with four specimens and three additional flakes / primary elements composed of refitted fragments.
Figure 9. The refitted set of raw material 4. 1) primary element; 2-5) first series of blade/lets; 6) core tablet; 7-11) second series of blade/lets; 12) core. Scale bar 2 cm, photo is enlarged for details.
Figure 10. The spatial distribution of the raw material 4 refits: Set #4 with 12 specimens and three flakes / primary elements composed of fragments. Two possible spatial reconstructions are presented, (a) as found and (b) with the core moved to its postu
Figure 11. The spatial distribution of the raw material 10 refits: Set #10b and four sets of fragments.
Figure 12. The spatial distribution of the raw material 14 refits: Set #14 with three specimens and two smaller sets.
Figure 13. The locations of the major knapping events on the floor of Brush Hut 2 reconstructed according to the spatial distribution of the flint assemblage. The dashed contours mark the two main flint clusters.
Table 5. Contingency table comparing the northern and southern knapping areas with macroliths and microliths considered separately. Significant adjusted residuals (AR) are shown in bold.
Table 6. Contingency table comparing macroliths and microliths between the northern and southern knapping areas. Significant adjusted residuals (AR) are shown in bold.
Dani Nadel, Daniel Kaufman, Udi Grinburg and Dan Malkinson
Flint knapping in a brush hut: a case study from Ohalo II, a 23000 year-old camp in the Sea of Galilee
A. Ronen: Middle Palaeolithic Humans in the Levant
Figure 1. Tentative site distribution by Middle Palaeolithic populations in the Galilee and northern Levant (area removed shown in Fig. 2).
Figure 2. Tentative site distribution by Middle Palaeolithic populations in Mount Carmel.
Avraham Ronen
Middle Palaeolithic Humans in the Levant: An Archaeological Perspective
S. A. Rosen: The Time-Space Discontinuum
Figure 1. Basic culture-chronological framework for the Negev. Absolute chronologies are approximate and the dating of some periods (e.g., Iron Age IIa, Chalcolithic-Early Bronze Age I transition) is debated (e.g., Regev et al. 2012).
Figure 2. Map of the Negev in the Near East. Modern borders are indicated by a dashed line. Vegetation zones are as follows: 1. Mediterranean vegetation, 2. Irano-Turanian steppe vegetation, 3. Sudano-Deccanian tropical vegetation (in patches in the ind
Figure 3. Site frequency graph from the Makhtesh Ramon, Map 204 (Rosen 1994) with periods and chronology. Note that the x-axis is actually nominal, and not to a linear scale due to the non-linea nature of periodization. Arrows indicate stratigraphic bre
Figure 4. Negev settlement frequency graphs by period from selected monographs of the Archaeological Survey of Israel. Survey numbers refer to the Israel Grid and to the official map publication number. Data from Gazit 1996, Govrin 1981, and Haiman 1981
Figure 5. Location of published Negev Survey blocks, in special references to Figures 4 and 5. Each square is 10 x 10 km.
Figure 6. East section of the Ramon I Rockshelter with calibrated radiocarbon dates.
Steven A. Rosen
The Time-Space Discontinuum: Scale in the Geography and Chronology of Negev Archaeology
D. Rosenberg: The incised flint assemblage from Neolithic Beisamoun and its significance
Figure 1. Map of Beisamoun and other Pre-Pottery Neolithic and Early Pottery Neolithic Yarmukian sites.
Figure 2. Beisamoun past excavations and the 2007 salvage excavation (enlarged area). The area enclosed in a broken line at the top of the drawing marks the dispersal of finds as noted by the French delegation. Hatched squares mark other excavations in th
Figure 3. Cores bearing incised pattern and grooves.
Figure 4. Various grooved and incised items – 1-9: Flakes, 10: over-shot, 11-12: End-scrapers.
Figure 5. Characteristics of the assemblage: metrics and weights. * Broken.
Figure 6. Characteristics of grooves: Metrics. * Broken.
Figure 7. Characteristics of the incised flint assemblage * Broken.
Danny Rosenberg
The incised flint assemblage from Neolithic Beisamoun and its significance
Ceramic Connections and Regional Entities: The Petrography of Late Chalcolithic Pottery from Sites in the Galilee (Israel)
D. Shalem et al.: Ceramic Connections and Regional Entities
Figure 1. Location map of sites mentioned in the text. For the sites of the Golan see Figure 5.
Figure 2. Pottery examples presented in the text. 1. Golan Ware, Tel Te’o (Eisenberg et al. 2001: Fig. 6.4:8); 2. Golan Ware, Beer Tzunam (Shalem 2003: Fig. 28:4); 3. Ḥula Ware, Tel Te’o (Eisenberg et al. 2001: Fig. 6.1:7); 4. Ḥula Ware, Beer Tzunam (Shal
Figure 3. Photomicrographs of Golanite pottery? 1: Peqi’in, Table 1:4. Fine-grained basalt fragment imbedded in ferruginous silty matrix. The basalt has a pilotaxitic and partly fludial microtexture of the groundmass and microlites of olivine and ore mine
Figure 4. Photomicrographs of pottery of several sites? 1. Assawir (Table 2:9), Morzovella sp. PPL; 2. Assawir (Table 2: 7), Acarinina sp. PPL; 3. Beer Tzunam (Table 2:13). Chiloguembelina. PPL; 4. Kaukab (Table 2:16). Globotruncana sp. PPL.
Figure 5. Map of volcanic units in the Galilee and the Golan after Weinstein and Garfunkel (2014) and the sites of the sampled Golanite vessels. Sites: 1. Beer Tzunam 2. Peqi’in 3. Asherat 5. Kaukab Springs 7. Agam Dalton. 8 Horbat Duvshan. 9. Tel Turmus
Table 1. Inventory and results of the petrographically analyzed Golanian jars.
Table 2. The inventory of the petrographicaly analyzed Painted Ware and Ḥula Ware vessels and cornets.
Dina Shalem, Anat Cohen-Weinberger, Bernardo Gandulla and Ianir Milevski
Communal Bison Hunting in Western North America: Current Understandings and Unresolved Issues
J. D. Speth: Communal Bison Hunting in Western North America
Figure 1. American bison (Bison bison). Photo by Jack Dykinga, released in the public domain by the Agricultural Research Service, United States Department of Agriculture (Image ID K5680-1). Available on Wikipedia http://en.wikipedia.org/wiki/File:Americ
Figure 2. Map of the extermination of the American bison to 1889. Intermediate gray, original range; dark gray, range 1870; black, range 1889; black numbers, estimated number of bison in 1889; white numbers, date of local extermination. Adapted from a dra
Figure 3. Pile of bison skulls awaiting processing into fertilizer and bone char used for refining sugar and bone black, Michigan Carbon Works, Detroit, Michigan, late 19th-century. Image (ID DPA4901600jpg) courtesy of the Burton Historical Collection, De
John D. Speth
Job 40:30-31 and the first whalers1
O. Tammuz: Job 40:30-31 and the first whalers
Figure 1. A Vertebra of a sperm whale with a bronze tip. Museo  Whitaker di Mozia (Marsala).
Oded Tammuz
F. R. Valla: More on Early Natufian building 131 at Eynan (Ain Mallaha)
Figure 1. Row of fallen stones in the lower fill of wall 51 (squares O-N-6-7-8). (1973 season, Photo F. Valla).
Figure 2. Map of the ashy floor at the base of wall 51. Note hearths 128 and 130. The floor is cut by the negative of wall 62 (removed at the time of this drawing). In squares O/4-5 and P/7: windows open in earlier deposits. In squares Q-R-S/6-7-8 ‘soundi
Figure 3. Map of floor 131a (1975 season. Drawing A. Dagand and M. Barazani).
Figure 4. Floor 131a: curves indicating the movement of floor 131a in its densely packed area (1975 season. Drawing A. Dagand and M. Barazani).
Figure 5. Floor 131a: Map and section of posthole 141 (1975 season, drawing M. Barazani).
Figure 6. Floor 131a: Map and section of posthole 140 (1975 season, drawing M. Barazani).
Figure 7. Floor 131a: Map and section of posthole 139 (1975 season, drawing M. Barazani).
Figure 8. Floor 131a: Map and section of posthole 145 (1975 season, drawing M. Barazani).
Figure 9. Floor 131a: Map and section of posthole 144 (1975 season, drawing M. Barazani).
Figure 10. Floor 131a: Map and section of posthole 143 (1975 season, drawing M. Barazani).
Figure 11. Floor 131a: Map and section of post-hole 142 (1975 season, drawing M. Barazani).
Figure 12. Floor 131a: relationship of the floor with walls 131 and 51 (behind) in squares N/6-7. (1975 season. Photo F. Valla).
Figure 13. Floor 131a: fragment of a soft limestone sculpture in square P/5. (1975 season, photo J. Perrot).
Figure 14. Floor 131a: Worked roe deer horn found broken into pieces in hearth 147 (1975 season, photo. M. Barazani).
Figure 15. Attributed to floor 131a: Calcite figurine (square P/9). (1975 season, photo L. Davin).
Figure 16. Attributed to floor 131a: Calcite figurine (square P/9). (1975 season, drawing F. Valla and M. Barazani).
Figure 17. Floor 131b: fragment of Iron ore, probably hematite (square O/6). (1976 season, photo M. Barazani).
Figure 18. Grave Homo 104: an old woman buried with a puppy dog. (1976 season, photo A. Dagan and F. Valla).
François R. Valla
More on Early Natufian building 131 at Eynan (Ain Mallaha), Israel
Hatrurim Quarry – A Newly Discovered Quarry and Production of Larnite Bifacial Tools
J. Vardi: Hatrurim Quarry – A Newly Discovered Quarry and Production of Larnite Bifacial Tools
Figure 1. The Hatrurim Basin. The largest quarries are marked in the map: HJQ (Hatrurim Junction Quarry), and HP (Har Parsa).
Figure 3. Larnite nodules from the Hatrurim quarry.
Figure 4. A fused lump of knapped Larnite bearing rock debitage.
Figure 6. A knapping area.
Figure 7. The excavation area (view to the east).
Figure 9. The excavated area, note the bedrock Larnite nodules and the layer of waste at the section (horizontal scale 50 cm).
Figure 10. The Waste and tools frequencies from the excavation at the Hatrurim quarry.
Figure 11. The flakes metric attributes in mm.
Figure 12. A flake.
Figure 13. A large flake.
Figure 14. Length vs. width dimensions of the flakes from the Hatrurim quarry.
Figure 15. Roughout of an axe.
Figure 16. Roughout (of an adze?).
Figure 17. An unfinished adze.
Figure 18. An unfinished adze.
Figure 19. An unfinished adze.
Figure 20. A bi-convex roughout.
Figure 22. Metric and weight attributes of the roughouts from the Hatrurim quarry (excluding initially prepared).
Figure 23. A list of the roughouts of the bifacial tools from the Hatrurim quarry site.
Jacob Vardi
A. Witztum et al.: The Origin of the Hebrew Word ṣîṣit ‘Fringe’
Figure 1. The source: Alan Gardiner. 1957. Egyptian Grammar, Oxford, Griffith Institute, Ashmolean Museum, p. 507. Reproduced with permission of the Griffith Institute at Oxford University.
Figure 1. Cactus hedges at Tel Erani. A look from the High terrace north, towards the acropolis (Courtesy: Marcin Czarnowicz).
Allan Witztum, Avi Gold and Mayer I. Gruber
The Origin of the Hebrew Word ṣîṣit ‘Fringe’1
Tel Erani – Reassessing old published records
Yuval Yekutieli
Figure 2. Cactus hedges and parcellation at Tel Erani. A. Cadastral outlines projected on 2006 aerial photograph of Tel Erani. B. Zalman Lifshitz’s cadastral map of ‘Iraq el-Manshiye. (The figure is compiled by the author on the basis of: A. Kiryat-Gat
Figure 3. Tel Erani – Settlement size in ancient periods and location of excavation areas (Basic map layer adapted from Kempinski and Gilead 1991).
Figure 4. Area D, location of excavation areas.
Figure 6. Area D, Deep probes (in grey shades) – Yeivin’s strata VII-XIII, Kempinski and Gilead’s layers C2, D and E.
Figure 7. Yeivin’s stratum VI, Kempinski and Gilead’s layer C1.
Figure 8. Yeivin’s stratum V.
Figure 10. Yeivin’s stratum IV superimposed on stratum V (shaded).
Figure 9. Yeivin’s stratum IV.
Figure 12. Yeivin’s stratum III superimposed on stratum IV (shaded).
Figure 13. Yeivin’s stratum IIc.
Figure 14. Yeivin’s stratum IIb superimposed on IIc (shaded).
Figure 15. Yeivin’s stratum IIa superimposed on IIb and IIc (shaded).
Figure 16. Mudbricks exposed on top of Yeivin’s stratum II, interpreted as remains of a city-wall by Brandl (1989: 383). Note the orientation of the brick-lines.
Elad Filler
‘Isaac went out to meditate in the field’ (Genesis 24:63) in the Exegesis of Philo of Alexandria and Naphtali Zvi Yehuda Berlin
N. Getzov: The Nahal Zippori Horizon, between the Lodian and Wadi Rabah Cultures
Figure 1.
Figure 2.
Figure 3.
Figure 4.
Figure 5.
Figure 6.
Nimrod Getzov, Ianir Milevski and Hamoudi Khalaily
The Nahal Zippori Horizon, between the Lodian and Wadi Rabah Cultures
Eran Viezel
‘In Green Pastures’ [bi-ne’ot deshshe]: Psalm 23 and Song of Songs 5 in the Poetry of Agi Mishol
Shamir Yona and Rafael Furman
The Varieties of Vengeance in the Book of Jeremiah
Recommend Papers

'Isaac went out to the field': Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead
 9781784918293, 9781784918309

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

Isaac went out ... to the field (Genesis 24:63)

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead edited by

Haim Goldfus, Mayer I. Gruber, Shamir Yona and Peter Fabian

Isaac went out ... to the field (Genesis 24:63)

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead edited by

Haim Goldfus, Mayer I. Gruber, Shamir Yona and Peter Fabian

Archaeopress Archaeology

Archaeopress Publishing Ltd Summertown Pavilion 18-24 Middle Way Summertown Oxford OX2 7LG www.archaeopress.com

ISBN 978-1-78491-829-3 ISBN 978-1-78491-830-9 (e-Pdf)

© Authors and Archaeopress 2019 Cover photograph by Sky Baloon, Israel

All rights reserved. No part of this book may be reproduced, or transmitted, in any form or by any means, electronic, mechanical, photocopying or otherwise, without the prior written permission of the copyright owners. Printed in England by Severn, Goucester This book is available direct from Archaeopress or from our website www.archaeopress.com

Contents

List of Figures and Tables........................................................................................................................................................... iii Authors’ and Editors’ Biographies........................................................................................................................................... viii Preface and Acknowledgments............................................................................................................................................... xiii Isaac Gilead: An Appreciation............................................................................................................................................... xv The Editors The Religious Dimension of Copper Metallurgy in the Southern Levant....................................................................1 Nissim Amzallag Neolithic Cult Sites in the Southern Negev, Israel...........................................................................................................14 Uzi Avner, Moti Shem-Tov, Lior Enmar, Gideon Ragolski, Rachamim Shem-Tov, and Omry Barzilai A Basket-Handled Teapot from Fazael 2.............................................................................................................................36 Shay Bar The Contributions of Early French Scholars to Levantine Prehistory........................................................................40 Ofer Bar-Yosef and François Valla The Riddle of the ‘Aurignacian’ in the Negev: The Lithic Assemblage from Nahal Neqarot in the Central Negev, Israel....................................................................................................................................................48 Anna Belfer-Cohen and A. Nigel Goring-Morris Forging a Link: Evidence for a ‘Lost Horizon’ – The Late Chalcolithic to EB 1 Transition in the Southern Levant.................................................................................................................................................................66 Eliot Braun Unresolved Pottery Neolithic Chrono-Stratigraphic and Chrono-Cultural issues: Comments on the Beginning and the End of the Pottery Neolithic Period..........................................................................................96 Avi Gopher Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant, ca. 4500-3900 BC.......................................................................................................................................................................109 Milena Gošić Animal Offerings from the Nawamis Fields and Coeval Habitation Sites in Southern Sinai...............................131 Liora Kolska Horwitz An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila...................................................................................145 Hamoudi Khalaily, Anastasia Shapiro and Ofer Marder Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period..................172 Janet Levy Stylistic Devices and Exegetical Techniques in ‘Rewritten Bible’ Compositions...................................................203 Atar Livneh Conus Ornaments from Tel Bareqet in an Early Bronze Age Near East Context....................................................210 Daniella E. Bar-Yosef Mayer, Sarit Paz and Yitzhak Paz i

Flint Knapping in a Brush Hut: A Case Study from Ohalo II, a 23000 Year-old Camp in the Sea of Galilee.................................................................................................................................................................216 Dani Nadel, Daniel Kaufman, Udi Grinburg and Dan Malkinson Middle Palaeolithic Humans in the Levant: An Archaeological Perspective..........................................................232 Avraham Ronen The Time-Space Discontinuum: Scale in the Geography and Chronology of Negev Archaeology....................240 Steven A. Rosen The Incised Flint Assemblage from Neolithic Beisamoun and its Significance.....................................................251 Danny Rosenberg Ceramic Connections and Regional Entities: The Petrography of Late Chalcolithic Pottery from Sites in the Galilee (Israel)..........................................................................................................................................262 Dina Shalem, Anat Cohen-Weinberger, Bernardo Gandulla and Ianir Milevski Communal Bison Hunting in Western North America: Current Understandings and Unresolved Issues.......................................................................................................................................... 278 John D. Speth Job 40:30-31 and the First Whalers.....................................................................................................................................295 Oded Tammuz More on Early Natufian Building 131 at Eynan (Ain Mallaha), Israel.......................................................................302 François R. Valla Hatrurim Quarry – A Newly Discovered Production Site of Larnite Bifacial Tools..............................................316 Jacob Vardi The Origin of the Hebrew Word ṣîṣit ‘Fringe’...................................................................................................................337 Allan Witztum, Avi Gold and Mayer I. Gruber Tel Erani – Reassessing Old Published Records...............................................................................................................341 Yuval Yekutieli ‘Isaac went out to meditate in the field’ (Genesis 24:63) in the Exegesis of Philo of Alexandria and Naphtali Zvi Yehuda Berlin..........................................................................................................................................355 Elad Filler The Nahal Zippori Horizon, Between the Lodian and Wadi Rabah Cultures..........................................................365 Nimrod Getzov, Ianir Milevski and Hamoudi Khalaily ‘In Green Pastures’ [bi-ne’ot deshshe]: Psalm 23 and Song of Songs 5 in the Poetry of Agi Mishol...................374 Eran Viezel The Varieties of Vengeance in the Book of Jeremiah....................................................................................................391 Shamir Yona and Rafael Furman

ii

List of Figures and Tables U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel

Figure 1. Eastern ‘Uvda Valley, remains of an elongated cell with an in-situ triad of perforated maṣṣeboth����������������������������������������������14 Figure 2. Map of Naḥal Roded area with the recorded Rodedian sites, in relation to lithology���������������������������������������������������������� 15 Figure 3. Map of currently recorded Rodedian sites in the Negev������������������������������������������������������������������������������������������������������������ 16 Figure 4. Har Assa, a typical pair of installations, the elongated cell points to the circle�������������������������������������������������������������������� 18 Figure 5. Har Argaman, plan of a typical pair of installations������������������������������������������������������������������������������������������������������������������� 18 Figure 6. ‘Uvda Valley, a large elongated cell with maṣṣebah, pointing to a barely preserved circle��������������������������������������������������� 19 Figure 7. Naḥal Roded, an elongated cell paved with small limestone slabs������������������������������������������������������������������������������������������� 19 Figure 8. Naḥal Roded, a pair of limestone maṣṣeboth, found tumbled and broken������������������������������������������������������������������������������� 19 Figure 9. Naḥal Roded, a triad of maṣṣeboth in a circular cell, as found (left) and after re-setting the central one�������������������������� 20 Figure 10. Naḥal Shehoret, a group of seven small maṣṣeboth as found��������������������������������������������������������������������������������������������������� 20 Figure 11. Naḥal Roded, perforated and naturally holed limestones, found scattered in and around the installations����������������� 20 Figure 12. Naḥal Roded, three examples of anthropomorphic stones: 1. Unshaped 2. With a hammered neck 3. Entirely hammered�������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 21 Figure 13. Unique stone images: 1. Eastern ‘Uvda Valley, a natural broad ‘female’-look cobble. 2. Naḥal Paran, a broad ‘female’-look cobble partially shaped. 3. Eastern ‘Uvda Valley, basalt figurine entirely hammered����������������������������� 21 Figure 14. Eastern ‘Uvda Valley, a pair of stone images at the end of an elongated cell����������������������������������������������������������������������� 22 Figure 15. Naḥal Roded, three examples of stones with elongated perforation������������������������������������������������������������������������������������� 22 Figure 16. Examples of stone bowls: 1. Naḥal Roded, natural bowl. 2. Naḥal Roded, a complete small bowl and fragments of large one. 3. Naḥal Botem, a well shaped bowl�������������������������������������������������������������������������������������������������������������� 22 Figure 17. Naḥal Roded, a ‘vase-shaped’ installation, as found and after cleaning�������������������������������������������������������������������������������� 23 Figure 18. Naḥal Roded, a ‘vase-shaped’ installation with a head-down stone image�������������������������������������������������������������������������� 23 Figure 19. Naḥal ‘Eteq, a pair of miniature houses as found, with a natural stone image lying on the flagstone���������������������������� 24 Figure 20. Naḥal Demama, two miniature houses built on the end of an elongated cell���������������������������������������������������������������������� 24 Figure 21. Naḥal Roded, perforated maṣṣeboth set with the perforation down��������������������������������������������������������������������������������������� 24 Figure 22. Naḥal Roded, regular and perforated buried maṣṣeboth: 1. As found. 2. As originally set before being covered������������ 25 Figure 23. Naḥal Roded, a perforated buried maṣṣebah: 1. As found. 2. As originally set, before being covered������������������������������� 25 Figure 24. Naḥal Roded, a natural, anthropomorphic copper nodule������������������������������������������������������������������������������������������������������ 26 Figure 25. Naḥal Roded, flint blades typical for the PPNB, an ‘Amuq arrow-head, on the and a Byblos arrow-head from a site on Har ‘Eshet��������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 26 Figure 26. Naḥal Roded, ad-hoc tools, typical for desert sites of the 6th-3rd millennia BC������������������������������������������������������������������ 26 Figure 27. Naḥal Roded Site 110, radiocarbon dates from an ash spot����������������������������������������������������������������������������������������������������� 27 Figure 28. Jebel Ḥashem alTaref, eastern Sinai, a perforated maṣṣebah among a group of animal ‘stone-drawings’����������������������� 27 Figure 29. Naḥal Roded, a vertical view of installation with a stone alignment������������������������������������������������������������������������������������ 30

S. Bar: A Basket-Handled Teapot from Fazael 2

Figure 1. The location of the crushed basket-handled teapot near an entrance to a room������������������������������������������������������������������ 37 Figure 2. The Fazael 2 basket-handled teapot���������������������������������������������������������������������������������������������������������������������������������������������� 38

O. Bar-Yosef and F. Valla: The Contributions of Early French Scholars to Levantine Prehistory

Figure 1. Father Germer-Durand�������������������������������������������������������������������������������������������������������������������������������������������������������������������� 41 Figure 2. Father A. Mallon�������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 42 Figure 3a. R. Neuville when incorporated as a diplomat���������������������������������������������������������������������������������������������������������������������������� 43 Figure 3b. R. Neuville in Jerusalem as the French Consul in Israel����������������������������������������������������������������������������������������������������������� 45

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev

Figure 1. Map of the central Negev, with Nahal Neqarot south of Maktesh Ramon marked���������������������������������������������������������������� 49 Figure 2. Nahal Neqarot rockshelter from the southeast. The Ramonian occupation layer underlies the massive fallen blocks� 50 Figure 3. Plan and general sections of Nahal Neqarot rockshelter����������������������������������������������������������������������������������������������������������� 50 Figure 4. Close-up of the two excavation areas and drawn section of grid squares Q/R along the 9 line������������������������������������������ 51 Figure 5. Radiocarbon dates from Nahal Neqarot���������������������������������������������������������������������������������������������������������������������������������������� 52 Figure 6. Pestle and other groundstone tools from Nahal Neqarot���������������������������������������������������������������������������������������������������������� 52 Figure 7. Core types and raw materials at Nahal Neqarot�������������������������������������������������������������������������������������������������������������������������� 53 Figure 8. Core types and raw materials by level at Nahal Neqarot����������������������������������������������������������������������������������������������������������� 54 Figure 9. Debitage counts at Nahal Neqarot������������������������������������������������������������������������������������������������������������������������������������������������� 54 Figure 10. Artefact counts and ratios at Nahal Neqarot����������������������������������������������������������������������������������������������������������������������������� 54 Figure 11. Microburin indices at Nahal Neqarot������������������������������������������������������������������������������������������������������������������������������������������ 54 Figure 12. Observations on microburins by level at Nahal Neqarot��������������������������������������������������������������������������������������������������������� 55 Figure 13. Scans of Microburins���������������������������������������������������������������������������������������������������������������������������������������������������������������������� 55

iii

Figure 14. Types of notch on microburins by level at Nahal Neqarot������������������������������������������������������������������������������������������������������ 55 Figure 15. Distal shapes of microburins by level at Nahal Neqarot���������������������������������������������������������������������������������������������������������� 55 Figure 16. Scans of Ramon points at Nahal Neqarot����������������������������������������������������������������������������������������������������������������������������������� 56 Figure 17. Typological counts by level at Nahal Neqarot��������������������������������������������������������������������������������������������������������������������������� 57 Figure 18. Scans of endscrapers on thick blades at Nahal Neqarot – note ‘Aurignacian-type’ retouch���������������������������������������������� 59 Figure 19. Thick endscrapers with ‘Aurignacian-type’ retouch at Nahal Neqarot��������������������������������������������������������������������������������� 60 Figure 20. Tools from Nahal Neqarot – a burin and scrapers��������������������������������������������������������������������������������������������������������������������� 61 Figure 21. Tools from Nahal Neqarot – Ramon points, denticulates and a chisel/retoucher��������������������������������������������������������������� 62 Figure 22. Mean metric attributes of Ramon points by level at Nahal Neqarot������������������������������������������������������������������������������������� 63 Figure 23. Retouch/backing types of Ramon points by level at Nahal Neqarot������������������������������������������������������������������������������������� 63 Figure 24. Shape of backed edge of Ramon points at Nahal Neqarot������������������������������������������������������������������������������������������������������� 63 Figure 25. Location of Ramon point tips, retouch related to tip, evidence for presence of microburin scar at Nahal Neqarot���������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 63 Figure 26. Correlation between Ramon point and microburin alignments attributes at Nahal Neqarot������������������������������������������� 63

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’

Figure 1. Map of southern Levant with principal LC to EB 1 transition deposits����������������������������������������������������������������������������������� 68 Figure 2. Radiocarbon dates from the Ashqelon Cluster of sites��������������������������������������������������������������������������������������������������������������� 70 Figure 3. A plan of the excavation locales at Barnea����������������������������������������������������������������������������������������������������������������������������������� 72 Figure 4. Aerial view of sand dune covering a portion of the Afridar neighborhood c. 1960��������������������������������������������������������������� 73 Figure 5. Inspector for the Israel Department of Antiquities and Museums, Dov Meiron, stands well below the modern surface in a bulldozer trench cut into a kurkar ridge in the Afridar neighborhood of Ashqelon, 1968�������������������������������������� 74 Figure 6. Afridar, Area G during excavation�������������������������������������������������������������������������������������������������������������������������������������������������� 75 Figure 7. View facing south from the excavation area of Barnea of the newly flattened littoral�������������������������������������������������������� 75 Figure 8. A view of the bulldozed littoral from the Barnea locale during excavation, facing north��������������������������������������������������� 76 Figure 9. Remains of a kurkar prominence at the northwest extent of ‘ground zero’, Barnea������������������������������������������������������������� 76 Figure 10. A partial view of the Barnea locale during excavation, facing south-southwest����������������������������������������������������������������� 76 Figure 11. View of a bulldozer sectioned segment of a kurkar ridge that parallels the beach at the western border of the Barnea excavation locale��������������������������������������������������������������������������������������������������������������������������������������������������������������� 77 Figure 12. View of a segment of a bulldozer section at the western edge of the excavation area in Barnea������������������������������������ 77 Figure 13. Closeup view of an in situ pot sheared off by a bulldozer in an archaeological deposit ca. 1.5 m above ‘ground zero’ in Barnea����������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 78 Figure 14. Comparison of ladder-like arrangements of cist burials from two sites������������������������������������������������������������������������������� 78 Figure 15. Pottery and a limestone vessel (n) from tumuli superimposed on ladder-like cist burials at Adeimeh, Jordan������������ 80 Figure 16. An early Chalcolithic vessel in the Amman Jordan Citadel Museum, probably from the Adeimeh Cemetery���������������� 80 Figure 17. A jar burial in Barnea with remains of a wall of kurkar stone just below the modern bulldozed surface������������������������ 81 Figure 18. Select Pottery from the Ashqelon Cluster and Chalcolithic Sites������������������������������������������������������������������������������������������� 83 Figure 19. Select Pottery from the Ashqelon Cluster and Chalcolithic Sites������������������������������������������������������������������������������������������� 84 Figure 20. Chalcolithic ground stone vessel fragments from the Ashqelon Cluster and Grar������������������������������������������������������������� 85 Figure 21. Select Pottery Types from Jebel Mutawwaq������������������������������������������������������������������������������������������������������������������������������ 86

A. Gopher: Unresolved Pottery Neolithic Chrono-Stratigraphic and Chrono-Cultural issues

Figure 1. Schematic timeline and geography of PN cultures��������������������������������������������������������������������������������������������������������������������� 97

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant

Figure 1. 14C dates from Gilat and Teleilat Ghassul������������������������������������������������������������������������������������������������������������������������������������ 111 Figure 2. Plan of the temple in Area E at Teleilat Ghassul, ‘Classic Courtyard’ phase������������������������������������������������������������������������� 113 Figure 3. 14C dates from Horvat Beter, Bir es-Safadi, Shiqmim and Giv’at ha-Oranim������������������������������������������������������������������������� 117 Figure 4. 14C dates from Horvat Qarqar South, Nahal Qanah, Shoham (North) and Peqi’in���������������������������������������������������������������� 118 Figure 5. Selected ossuaries from Peqi’in: A. female B. male C. zoomorphic���������������������������������������������������������������������������������������� 119 Figure 6. 14C dates from Nahal Mishmar cave of the Treasure����������������������������������������������������������������������������������������������������������������� 121 Figure 7. Selected Ghassulian copper artifacts: A. anthropomorphic standard, B. ‘crown’ featuring architectural and zoomorphic motifs, C. standard featuring ibexes and motifs of tools, D. skeuomorphic axe�����������������������������������������������������������122

L. Kolska Horwitz: Animal Offerings from the Nawamis Fields and Coeval Habitation Sites

Figure 1. Example of a nawamis from Ein Huderah������������������������������������������������������������������������������������������������������������������������������������ 132 Figure 2. Photograph showing a section of the nawamis field at Ein Huderah�������������������������������������������������������������������������������������� 132 Figure 3. Map showing location of the nine nawamis fields in southern Sinai excavated by Israeli teams�������������������������������������� 133 Figure 4. Collapsed nawamis photographed ca. 1990-1920 by G. Eric and Edith Matson��������������������������������������������������������������������� 134 Table 1. Radiocarbon dates for the nawamis����������������������������������������������������������������������������������������������������������������������������������������������� 134 Table 2. Faunal remains from the nawamis tombs (given as NISP counts)��������������������������������������������������������������������������������������������� 135 Table 3. Identifications of faunal species from the nawamis habitation sites���������������������������������������������������������������������������������������� 138 Table 4. Species list for the EBII Southern Sinai faunal assemblages����������������������������������������������������������������������������������������������������� 140

iv

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 1. Location map����������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 146 Figure 2. Plan of excavation areas���������������������������������������������������������������������������������������������������������������������������������������������������������������� 146 Figure 3. Main section showing the main archaeological deposits�������������������������������������������������������������������������������������������������������� 147 Figure 4. Archaeological layer overlaid the sterile loess�������������������������������������������������������������������������������������������������������������������������� 147 Figure 5. Plan of the excavation squares and the main loci��������������������������������������������������������������������������������������������������������������������� 148 Figure 6. Pit 101 in Sq. X-3����������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 149 Figure 7. Photo of pit 102 with molded mud-bricks���������������������������������������������������������������������������������������������������������������������������������� 149 Figure 8. Plan of Sq. Y-9 including L 206����������������������������������������������������������������������������������������������������������������������������������������������������� 150 Figure 9. Two phases of Living surfaces in Area B������������������������������������������������������������������������������������������������������������������������������������� 151 Figure 10. Walls 2007 and 208 in Sq. X-6, 7�������������������������������������������������������������������������������������������������������������������������������������������������� 152 Figure 11. Cores. 1-4 single platform cores������������������������������������������������������������������������������������������������������������������������������������������������� 153 Figure 12. Cores. 1 Amorphous core; 2-3 two platform cores������������������������������������������������������������������������������������������������������������������ 154 Figure 13. Tools. 1-2 Arrowheads; 3-4 Sickle blades; 5-6 fan-scrapers��������������������������������������������������������������������������������������������������� 155 Figure 14. Tools. 1, 4 Awls; 2 denticulation; 3 massive borer; 4 bifacial in preparation���������������������������������������������������������������������� 156 Figure 15. Bowls����������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 158 Figure 16. Jars�������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 159 Figure 17. Handles and Bases������������������������������������������������������������������������������������������������������������������������������������������������������������������������ 161 Figure 18. Photomicrograph of sample 3001���������������������������������������������������������������������������������������������������������������������������������������������� 162 Figure 19. 1-2 Grinding slab; 3-4 Mobil cup-mark�������������������������������������������������������������������������������������������������������������������������������������� 163 Figure 20. 1-2 Grooved stones����������������������������������������������������������������������������������������������������������������������������������������������������������������������� 164 Figure 21. Limestone artifacts. 1-2 abraders; 3-4 cores����������������������������������������������������������������������������������������������������������������������������� 165 Figure 22. Limestone artifacts. 1-3 massive scrapers�������������������������������������������������������������������������������������������������������������������������������� 166 Figure 23. Limestone artifacts. 1-2 side-scrapers; 3 awls; 4 denticulated artifacts������������������������������������������������������������������������������ 167 Table 1. Breakdown of Flint Artifacts According to Type and Area of Excavation������������������������������������������������������������������������������� 151 Table 2. Flint Core frequencies���������������������������������������������������������������������������������������������������������������������������������������������������������������������� 160 Table 3. Flint Tool frequencies���������������������������������������������������������������������������������������������������������������������������������������������������������������������� 160 Table 4. Inventory of the thin-sections������������������������������������������������������������������������������������������������������������������������������������������������������� 161 Table 5. Ground stone and knapped limestone artifacts according to raw material��������������������������������������������������������������������������� 162 Table 6. Cores on limestone artifacts����������������������������������������������������������������������������������������������������������������������������������������������������������� 164 Table 7. Frequencies of knapped limestone tools�������������������������������������������������������������������������������������������������������������������������������������� 168 Table 8. Dimensions of several limestone categories�������������������������������������������������������������������������������������������������������������������������������� 168

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period

Figure 1. Map of main sites mentioned in the chapter����������������������������������������������������������������������������������������������������������������������������� 172 Figure 2. Harvesting flax, author’s garden, near the Beersheva Valley������������������������������������������������������������������������������������������������� 175 Figure 3. Flax stem anatomy A=cuticle, B=epidermis, C=cortex, D=fibre bundles, E=woody core, F=pith, G=lumen���������������������� 175 Figure 4. Supported on the thigh spinning, Navajo����������������������������������������������������������������������������������������������������������������������������������� 177 Figure 5. Drop spinning, high whorl spindle, Iran������������������������������������������������������������������������������������������������������������������������������������� 178 Figure 6. Horizontal ground loom on ceramic bowl Badari, Egypt��������������������������������������������������������������������������������������������������������� 179 Figure 7. Representative spinning whorls from the Southern Levant (1)��������������������������������������������������������������������������������������������� 181 Figure 8. Representative spinning whorls from the Southern Levant (2)��������������������������������������������������������������������������������������������� 182 Figure 9. Spinning bowl, Neve Ur from Perrot������������������������������������������������������������������������������������������������������������������������������������������� 184 Figure 10. Schematic of linen shroud, Cave of the Warrior���������������������������������������������������������������������������������������������������������������������� 186 Figure 11. Schematic of weft fringe, shroud, Cave of the Warrior���������������������������������������������������������������������������������������������������������� 187 Figure 12. Shroud, detail of decorative band and warp fringe���������������������������������������������������������������������������������������������������������������� 187 Figure 13. Sash, detail of countered weft twining and warp tassels, Cave of the Warrior������������������������������������������������������������������ 187 Figure 14. Fresco, The Procession, Teleilat Ghassul���������������������������������������������������������������������������������������������������������������������������������� 190 Figure 15. Leather sandals, Cave of the Warrior���������������������������������������������������������������������������������������������������������������������������������������� 192 Figure 16. Fresco, The Notables, Teleilat Ghassul�������������������������������������������������������������������������������������������������������������������������������������� 193 Figure 17. Fresco, The Notables, Teleilat Ghassul�������������������������������������������������������������������������������������������������������������������������������������� 193 Table 1. Person-power labour days in shroud manufacture, Cave of the Warrior������������������������������������������������������������������������������� 188

D. E. Bar-Yosef Mayer et al.: Conus Ornaments from Tel Bareqet in an Early Bronze Age Near East Context

Figure 1. The Conus apex beads from Bareqet������������������������������������������������������������������������������������������������������������������������������������������� 211

D. Nadel et al.: Flint Knapping in a Brush Hut

Figure 1. a) Location map������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 216 Figure 1. b) plan of the Ohalo II site������������������������������������������������������������������������������������������������������������������������������������������������������������� 217 Figure 2. The distribution map of the flint assemblage (debitage and tools) on the floor of Brush Hut 2��������������������������������������� 219 Figure 3. The distribution maps of a) blade/lets and cores; b) primary elements and cores; c) core trimming elements and cores���������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 221 Figure 4. The distribution maps of a) tools and b) microliths����������������������������������������������������������������������������������������������������������������� 222 Figure 5. The distribution map of the minute remains���������������������������������������������������������������������������������������������������������������������������� 222

v

Figure 6. Spatially explicit regression results for pairs of flint products. a) primary elements – core trimming elements; b) blade/lets – core trimming elements; c) blade/lets – primary elements; d) blade/lets – flakes; e) tools – blade/lets; f) tools – core trimming elements��������������������������������������������������������������������������������������������������������������������������������������������������������� 223 Figure 7. The spatial distribution of the raw material 2 refits����������������������������������������������������������������������������������������������������������������� 224 Figure 8. The spatial distribution of the raw material 3 refits����������������������������������������������������������������������������������������������������������������� 224 Figure 9. The refitted set of raw material 4������������������������������������������������������������������������������������������������������������������������������������������������ 225 Figure 10. The spatial distribution of the raw material 4 refits��������������������������������������������������������������������������������������������������������������� 226 Figure 11. The spatial distribution of the raw material 10 refits������������������������������������������������������������������������������������������������������������ 227 Figure 12. The spatial distribution of the raw material 14 refits������������������������������������������������������������������������������������������������������������ 227 Figure 13. The locations of the major knapping events on the floor of Brush Hut 2 reconstructed according to the spatial distribution of the flint assemblage���������������������������������������������������������������������������������������������������������������������������������� 227 Table 1. The flint assemblages from six brush huts���������������������������������������������������������������������������������������������������������������������������������� 218 Table 2. The Brush Hut 2 refitted products, by raw material������������������������������������������������������������������������������������������������������������������� 220 Table 3. Pearson’s r correlation tests for the debitage, cores, minute debitage, and tools on the floor of Brush Hut 2���������������� 220 Table 4. Contingency table comparing the north and south knapping areas with macroliths and microliths combined������������ 226 Table 5. Contingency table comparing the northern and southern knapping areas with macroliths and microliths considered separately���������������������������������������������������������������������������������������������������������������������������������������������������������� 228 Table 6. Contingency table comparing macroliths and microliths between the northern and southern knapping areas����������� 228

A. Ronen: Middle Palaeolithic Humans in the Levant

Figure 1. Tentative site distribution by Middle Palaeolithic populations in the Galilee and northern Levant������������������������������� 235 Figure 2. Tentative site distribution by Middle Palaeolithic populations in Mount Carmel�������������������������������������������������������������� 236

S. A. Rosen: The Time-Space Discontinuum

Figure 1. Basic culture-chronological framework for the Negev������������������������������������������������������������������������������������������������������������ 241 Figure 2. Map of the Negev in the Near East���������������������������������������������������������������������������������������������������������������������������������������������� 242 Figure 3. Site frequency graph from the Makhtesh Ramon, Map 204 (Rosen 1994) with periods and chronology.����������������������� 243 Figure 4. Negev settlement frequency graphs by period from selected monographs of the Archaeological Survey of Israel����� 244 Figure 5. Location of published Negev Survey blocks������������������������������������������������������������������������������������������������������������������������������� 245 Figure 6. East section of the Ramon I Rockshelter with calibrated radiocarbon dates����������������������������������������������������������������������� 246

D. Rosenberg: The Incised Flint Assemblage from Neolithic Beisamoun and its Significance

Figure 1. Map of Beisamoun and other Pre-Pottery Neolithic and Early Pottery Neolithic Yarmukian sites��������������������������������� 252 Figure 2. Beisamoun past excavations and the 2007 salvage excavation����������������������������������������������������������������������������������������������� 253 Figure 3. Cores bearing incised pattern and grooves�������������������������������������������������������������������������������������������������������������������������������� 255 Figure 4. Various grooved and incised items – 1-9: Flakes, 10: over-shot, 11-12: End-scrapers��������������������������������������������������������� 256 Table 1. Characteristics of the assemblage: metrics and weights����������������������������������������������������������������������������������������������������������� 257 Table 2. Characteristics of grooves: Metrics����������������������������������������������������������������������������������������������������������������������������������������������� 258 Table 3. Characteristics of the incised flint assemblage��������������������������������������������������������������������������������������������������������������������������� 259

D. Shalem et al.: Ceramic Connections and Regional Entities

Figure 1. Location map of sites mentioned in the text����������������������������������������������������������������������������������������������������������������������������� 263 Figure 2. Pottery examples presented in the text�������������������������������������������������������������������������������������������������������������������������������������� 264 Figure 3. Photomicrographs of Golanian pottery�������������������������������������������������������������������������������������������������������������������������������������� 267 Figure 4. Photomicrographs of pottery of several sites���������������������������������������������������������������������������������������������������������������������������� 268 Figure 5. Map of volcanic units in the Galilee and the Golan and the sites of the sampled Golanite vessels���������������������������������� 270 Table 1. Inventory and results of the petrographically analyzed Golanian jars����������������������������������������������������������������������������������� 271 Table 2. Inventory of the petrographicaly analyzed Painted Ware and Ḥula Ware vessels and cornets������������������������������������������ 273

J. D. Speth: Communal Bison Hunting in Western North America

Figure 1. American bison (Bison bison)��������������������������������������������������������������������������������������������������������������������������������������������������������� 279 Figure 2. Map of the extermination of the American bison to 1889������������������������������������������������������������������������������������������������������� 281 Figure 3. Pile of bison skulls awaiting processing into fertilizer and bone char used for refining sugar and bone black������������� 282

O. Tammuz: Job 40:30-31 and the First Whalers

Figure 1. A Vertebra of a sperm whale with a bronze tip������������������������������������������������������������������������������������������������������������������������� 296

F. R. Valla: More on Early Natufian Building 131 at Eynan (Ain Mallaha)

Figure 1. Row of fallen stones in the lower fill of wall 51������������������������������������������������������������������������������������������������������������������������� 304 Figure 2. Map of the ashy floor at the base of wall 51������������������������������������������������������������������������������������������������������������������������������� 305 Figure 3. Map of floor 131a���������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 306 Figure 4. Floor 131a: curves indicating the movement of floor 131a in its densely packed area������������������������������������������������������� 307 Figure 5. Floor 131a: Map and section of posthole 141����������������������������������������������������������������������������������������������������������������������������� 308 Figure 6. Floor 131a: Map and section of posthole 140����������������������������������������������������������������������������������������������������������������������������� 308 Figure 7. Floor 131a: Map and section of posthole 139����������������������������������������������������������������������������������������������������������������������������� 308

vi

Figure 8. Floor 131a: Map and section of posthole 145����������������������������������������������������������������������������������������������������������������������������� 309 Figure 9. Floor 131a: Map and section of posthole 144����������������������������������������������������������������������������������������������������������������������������� 309 Figure 10. Floor 131a: Map and section of posthole 143��������������������������������������������������������������������������������������������������������������������������� 310 Figure 11. Floor 131a: Map and section of post-hole 142�������������������������������������������������������������������������������������������������������������������������� 310 Figure 12. Floor 131a: relationship of the floor with walls 131 and 51��������������������������������������������������������������������������������������������������� 310 Figure 13. Floor 131a: fragment of a soft limestone sculpture in square P/5��������������������������������������������������������������������������������������� 311 Figure 14. Floor 131a: Worked roe deer horn found broken into pieces in hearth 147����������������������������������������������������������������������� 311 Figure 15. Attributed to floor 131a: Calcite figurine (square P/9)���������������������������������������������������������������������������������������������������������� 312 Figure 16. Attributed to floor 131a: Calcite figurine (square P/9)���������������������������������������������������������������������������������������������������������� 312 Figure 17. Floor 131b: fragment of Iron ore, probably hematite (square O/6)�������������������������������������������������������������������������������������� 313 Figure 18. Grave Homo 104: an old woman buried with a puppy dog���������������������������������������������������������������������������������������������������� 313

J. Vardi: Hatrurim Quarry – A Newly Discovered Production Site of Larnite Bifacial Tools

Figure 1. The Hatrurim Basin. The largest quarries are marked in the Map: HJQ – Hatrurim (Junction) Quarry, and HP (Har Parsa)����������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 317 Figure 2. The Hatrurim quarry: a view from the bottom of the slope (looking northwards)������������������������������������������������������������� 318 Figure 3. Larnite nodules from the Hatrurim quarry�������������������������������������������������������������������������������������������������������������������������������� 318 Figure 4. A fused lump of knapped Larnite bearing rock debitage��������������������������������������������������������������������������������������������������������� 319 Figure 5. Lithic waste of Larnite bearing rock on the hill summit���������������������������������������������������������������������������������������������������������� 320 Figure 6. A knapping area������������������������������������������������������������������������������������������������������������������������������������������������������������������������������ 320 Figure 7. The excavation area (view to the east)��������������������������������������������������������������������������������������������������������������������������������������� 321 Figure 8. Debitage from the excavated area����������������������������������������������������������������������������������������������������������������������������������������������� 322 Figure 9. The excavated area, note the bedrock Larnite nodules and the layer of waste at the section������������������������������������������ 322 Figure 10. The Waste and tools frequencies from the excavation at the Hatrurim quarry���������������������������������������������������������������� 323 Figure 11. The flakes metric attributes in mm������������������������������������������������������������������������������������������������������������������������������������������� 323 Figure 12. A flake��������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 323 Figure 13. A large flake����������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 324 Figure 14. Length vs. width dimensions of the flakes from the Hatrurim quarry�������������������������������������������������������������������������������� 325 Figure 15. Roughout of an axe����������������������������������������������������������������������������������������������������������������������������������������������������������������������� 326 Figure 16. Roughout (of an adze?)���������������������������������������������������������������������������������������������������������������������������������������������������������������� 327 Figure 17. An unfinished adze����������������������������������������������������������������������������������������������������������������������������������������������������������������������� 328 Figure 18. An unfinished adze����������������������������������������������������������������������������������������������������������������������������������������������������������������������� 329 Figure 19. An unfinished adze����������������������������������������������������������������������������������������������������������������������������������������������������������������������� 330 Figure 20. A bi-convex roughout������������������������������������������������������������������������������������������������������������������������������������������������������������������ 331 Figure 21. An initially shaped adze�������������������������������������������������������������������������������������������������������������������������������������������������������������� 332 Figure 22. Metric and weight attributes of the roughouts from the Hatrurim quarry����������������������������������������������������������������������� 332 Figure 23. A list of the roughouts of the bifacial tools from the Hatrurim quarry site����������������������������������������������������������������������� 333

A. Witztum et al.: The Origin of the Hebrew Word ṣîṣit ‘Fringe’

Figure 1. The source: Alan Gardiner. 1957. Egyptian Grammar, Oxford, Griffith Institute, Ashmolean Museum, p. 507����������������� 338

Y. Yekutieli: Tel Erani – Reassessing Old Published Records

Figure 1. Cactus hedges at Tel Erani. A look from the High terrace north, towards the acropolis���������������������������������������������������� 341 Figure 2. Cactus hedges and parcellation at Tel Erani������������������������������������������������������������������������������������������������������������������������������ 343 Figure 3. Tel Erani – Settlement size in ancient periods and location of excavation areas���������������������������������������������������������������� 344 Figure 4. Area D, location of excavation areas�������������������������������������������������������������������������������������������������������������������������������������������� 345 Figure 5. Area D, all starta superimposed��������������������������������������������������������������������������������������������������������������������������������������������������� 346 Figure 6. Area D, Deep probes (in grey shades) – Yeivin’s strata VII-XIII, Kempinski and Gilead’s layers C2, D and E������������������� 346 Figure 7. Yeivin’s stratum VI, Kempinski and Gilead’s layer C1�������������������������������������������������������������������������������������������������������������� 347 Figure 8. Yeivin’s stratum V�������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 348 Figure 9. Yeivin’s stratum IV������������������������������������������������������������������������������������������������������������������������������������������������������������������������� 349 Figure 10. Yeivin’s stratum IV superimposed on stratum V�������������������������������������������������������������������������������������������������������������������� 349 Figure 11. The cut of Yeivin’s strata V and IV, and intact remains on the flat surface east of it������������������������������������������������������� 350 Figure 12. Yeivin’s stratum III superimposed on stratum IV������������������������������������������������������������������������������������������������������������������� 350 Figure 13. Yeivin’s stratum IIc���������������������������������������������������������������������������������������������������������������������������������������������������������������������� 351 Figure 14. Yeivin’s stratum IIb superimposed on IIc��������������������������������������������������������������������������������������������������������������������������������� 351 Figure 15. Yeivin’s stratum IIa superimposed on IIb and IIc�������������������������������������������������������������������������������������������������������������������� 352 Figure 16. Mudbricks exposed on top of Yeivin’s stratum II�������������������������������������������������������������������������������������������������������������������� 353

N. Getzov et al.: The Nahal Zippori Horizon, Between the Lodian and Wadi Rabah Cultures

Figure 1. Location map with sites of the Nahal Zippori Horizon������������������������������������������������������������������������������������������������������������ 365 Figure 2. Yiftahel, pottery from Area G, Stratum G4, the Lodian culture���������������������������������������������������������������������������������������������� 366 Figure 3. Yiftahel, pottery from Area I, Stratum I2, Nahal Zippori Horizon����������������������������������������������������������������������������������������� 367 Figure 4. Yiftahel, decorated pottery from Area I, Stratum I2, Nahal Zippori Horizon����������������������������������������������������������������������� 368 Figure 5. Ein Zippori, selected pottery from the Wadi Rabah culture���������������������������������������������������������������������������������������������������� 369 Figure 6. Ein Zippori, pottery from Locus 7727, Nahal Zippori Horizon������������������������������������������������������������������������������������������������ 370

vii

Authors’ and Editors’ Biographies

Nissim Amzallag is Research Fellow in the Department of Bible, Archaeology and Ancient Near Eastern Studies at Ben-Gurion University of the Negev in Beer Sheva. During the last decade, he investigated the metallurgical background of ancient Yahwism as well as the cultural dimension of ancient metallurgy. Amzallag summarized his findings in these areas in twenty-five published articles, mostly in biblical studies journals. In addition, he investigated the IsraelEdom relationship in the Bible and its expression in ancient biblical poetry. This was the subject of his PhD dissertation (Ben-Gurion University, 2015). This line of research led to the publication of a book and 19 articles. Nissim Amzallag also earned a PhD in Biology (The Hebrew University, Jerusalem, 1993). Between 1990 and 2006, he published seven chapters in books, and 33 articles on plant physiology, development and evolution. Amzallag is also a researcher in philosophy and the history of science. Between 2002 and 2010, he published in this field three books, three book chapters, and ten articles.

volumes, and he authored or co-authored some 380 papers and book chapters. Daniella E. Bar-Yosef Mayer is a zooarchaeologist specializing in mollusks from archaeological sites. She is Collections Manager for Paleontology and Archaeomalacology at the Steinhardt National Natural History Museum at Tel-Aviv University. Omri Barzilai of Tzur Hadassah is Head of the Prehistory Branch of the Israel Antiquities Authority. Anna Belfer-Cohen is Professor Moshe Stekelis Professor of Prehistoric Archaeology at the Institute of Archaeology at the Hebrew University of Jerusalem. Her research interests include the evolution, spread and characteristics of Upper Palaeolithic entities; burial customs of prehistoric societies; the transition from mobile hunter-gatherers to sedentary farmers; and the evolution of human cognition. Eliot Braun is a Senior Fellow WF Albright Institute of Archaeological Research and an Associate Researcher, Centre de Recherche Français de Jérusalem. He earned his MA cum laude from the Hebrew University Institute of Archaeology in 1991 and his PhD from the Tel Aviv University Institute of Archaeology in 1997. He specializes in Early Bronze I, but he has worked at sites of many periods. His publications include En Shadud: Salvage Excavations at a Farming Community in the Jezreel Valley, Israel (Oxford: British Archaeological Reports, 1985); Salvage and Rescue Excavations at the Late Prehistoric Site of Yiftah’el in Lower Galilee, Israel (Jerusalem: Israel Antiquities Authority, 1997), in which he authored 14 chapters and co-authored 2 chapters); Early Beth Shan (Strata XIX-XIII): G. M. FitzGerald’s Deep Cut on the Tell (University Museum Monograph 121; Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology, 2004); Early Megiddo: Early Megiddo on the East Slope (The ‘Megiddo Stages’): A Report on the Early Occupation of the East Slope of Megiddo--Results of the Oriental Institute’s Excavations, 1925-1933 (Chicago: Oriental Institute of the University of Chicago, 2013), and numerous articles in peer reviewed journals, collections of colloquia, Festschriften, and many articles in reference works.

Uzi Avner earned his PhD summa cum laude from the Hebrew University of Jerusalem for his dissertation Studies in the Material and Spiritual Culture of the Desert Population During the 6th to 3rd Millennia BCE. He is currently associated with the Dead Sea-Arava Science Center in Eilat, Israel. Shay Bar received his PhD in archaeology from the University of Haifa in 2009. He is a researcher at the Zinman Institute and teaches in the archaeology department at the University of Haifa. He currently heads the excavations of the Bronze and Iron Age site of Tel Esur in the Sharon plain, and the Fazael Valley proto-historic project in the Jordan Valley. He is also the director of the Manasseh Hill Country Survey. Ofer Bar-Yosef received his PhD in 1970 from the Hebrew University of Jerusalem, where he taught from 1968 to 1987. Later (1988-2013) he taught in the Department of Anthropology at Harvard University. Bar-Yosef was the joint excavator of Lower through late Paleolithic and early Neolithic open-air sites and caves in Israel, Sinai (Egypt), Turkey, the Czech Republic, the Republic of Georgia, and the People’s Republic of China. His field work targeted early and late dispersals from Africa, Neanderthals and modern humans, and the transition from foraging to farming. He is an Associate (from 2001) and Full Member (from 2010) of the National Academy of Sciences of the USA, and a Corresponding Fellow of the British Academy (2005). He co-edited 19

Anat Cohen-Weinberger is the Head of the Petrographic Research Laboratory at the Israel Antiquities Authority (IAA). She studies the technology and provenance of pottery and other clay objects, collaborating with colleagues from the IAA and other viii

research institutes in Israel and abroad. Her PhD dissertation deals with the ‘Petrography of Middle Bronze 2 Age Pottery: Implications to Understanding Egypto-Canaanite Relations’. Other research topics and publications include analyses of varieties of pottery from the Neolithic period to medieval times. She seeks to identify trade networks, group migrations, and other social contacts and their mechanisms.

His main topics of research are the cultural interactions between Ancient Near Eastern peoples, particularly West Semitics and early Transcaucasians, and the problematic of otherness in the Early Bronze Age and later. Haim Goldfus teaches classical archaeology in the Department of Bible, Archaeology and Near Eastern Studies at Ben-Gurion University of the Negev. His main fields of interest and research are related to early Roman and Late Antique/ early Byzantine archaeology.

Lior Enmar is a geologist residing at Kaslon, Israel. He earned his MA in geology at the Hebrew University of Jerusalem. He is a tour guide in Israel and around the world. His interest in cultic sites in the desert began when he met Uzi Avner while working as a tour guide at the Eilat Field School.

Nimrod Getzov is a researcher in the archaeological research department of the Israel Antiquities Authority. He specialized in archeology of the Galilee and in the proto-history of Israel. He participated in the Upper Galilee survey and the Western Galilee survey. He also participated in, managed and published the findings of many excavations, including Ḥorbat ‘Uẓa, Tel Bet Yeraḥ, Ḥorbat Beth Zeneta, Ha-Gosherim, Kaukab Springs, Midrakh ‘Oz, Yiftaḥ’el, and ‘En Ẓippori.

Peter Fabian completed his studies at Ben-Gurion University of the Negev in 2005 with a PhD dissertation concerning the Roman army camp in Oboda-Avedat. After working for many years in the Israel Antiquities Authority as a senior archaeologist and an advisor, he was given a tenure track position as a classical archaeologist in the Department of Bible, Archaeology and Ancient Near Eastern Studies at Ben-Gurion University of the Negev, where he is currently a Senior Lecturer. Dr. Fabian combines his interests and expert knowledge and long track-record of fieldwork in Hellenistic, Roman, Byzantine and Early Islamic archaeology, together with excavations and studies concerning the proto-history of the Southern Levant, particularly the Chalcolithic period of the Negev and the Shephela lowlands of Israel.

Avi Gold is an Independent Scholar. Avi Gopher is Professor of Archaeology at the University of Tel Aviv. His publications include The Flint Assemblages of Munhata – Final Report, Les Cahiers Du Centre de Recherche Francais de Jerusalem, Vol. 4, Paris: Association Paléorient, Paris, 1989; Arrowheads of the Neolithic Levant – A Seriation Analysis, American Schools for Oriental Research Dissertation Series Vol. 10, Winona Lake, Indiana: Eisenbrauns, 1994; and Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel: Archaeological Investigations at the Sites of Nahal Zehora. (forthcoming).

Elad Filler teaches Jewish Hellenistic Philosophy in the Department of Jewish Thought at Bar-Ilan University. His research interests include Philo of Alexandria, philosophical exegesis of the Bible, and early Christian theology and philosophy. Publications include the book Philo’s Quod Deterius Potiori insidiari solet: A Complete Sequential Philosophical and Literary Analysis (in press).

A. Nigel Goring-Morris is Professor at the Institute of Archaeology at the Hebrew University of Jerusalem. His publications include the book At the Edge: Terminal Pleistocene Hunter-Gatherers in the Negev and Sinai, BAR International Series 361, (Oxford: British Archaeological Reports, 1987); the book, co-authored with Anna Belfer-Cohen More Than Meets the Eye: Studies on Upper Palaeolithic Diversity in the Near East (Oxford: Oxbow Press, 2003); and the book, co-authored with Ofer Bar-Yosef and Avi Gopher, Gilgal: Excavations at Early Neolithic Sites in the Lower Jordan Valley: The Excavations of Tamar Noy (American School of Prehistoric Research Monographs, Oxford: Oxbow/Oakville, CT: David Brown, 2010).

Rafael Furman is a doctoral student of Shamir Yona in the Department of Bible, Archaeology and Ancient Near Eastern Studies at Ben-Gurion University of the Negev in Beer-Sheva. Furman specializes in biblical prophecy. Bernardo Gandulla is the director of the ‘André Finet’ Free Chair of Canaanean Studies at the University of Buenos Aires and Professor Emeritus of Ancient Near Eastern History at the University of Buenos Aires and the National University of Luján (Argentina). He was director of the Research Project ‘Economy and Society in the Chalcolithic Period of the Southern Levant (4500-3700 BC): Production and Exchange [2010-2014]’ At present he is Senior Researcher in the Project ‘Egypt and Palestine Relations at the End of the 4th millennium BC’. Both projects are under the auspices of the Ministry of Science and Technology (Argentina).

Milena Gošić earned her PhD at Ben-Gurion University of the Negev as a student of Prof. Isaac Gilead. She teaches Near Eastern archaeology at the University of Belgrade in Serbia. Udi Grinburg completed his MA studies in prehistoric archaeology at the Department of Archaeology, the University of Haifa, in 2005. ix

Mayer I. Gruber is Professor Emeritus and Past Chair of the Department of Bible Archaeology and Ancient Near Eastern Studies at Ben-Gurion University of the Negev in Beer Sheva. His publications include Aspects of Nonverbal Communication in the Ancient Near East (2 vols.; Rome: Biblical Institute Press, 1980); The Motherhood of God and Other Studies (Atlanta: Scholars Press, 1992); Women in the Biblical World: A Study Guide (American Theological Library Association, Bibliography Series, no. 38; Lanham, MD, and London: Scarecrow Press, 1995); Rashi’s Commentary on Psalms (Leiden & Boston: Brill, 2004); Women in Ancient Israel and Studies in Early Jewish Civilization, translated from English into Chinese by Shuqing Zhang (Beijing: China Social Sciences Press, 2009); The Women of Israel by Grace Aguilar with a New Introduction and Commentary (Piscataway, NJ: Gorgias Press 2011, 2013); and Hosea: A Textual Commentary (London: Bloomsbury T & T Clark, 2017).

Southern Africa. She is associate editor of the Journal of Arid Environments. Janet Levy earned her PhD in 2019 in the Department of Bible, Archaeology and Ancient Near Eastern Studies at Ben-Gurion University of the Negev under the supervision of Professor Isaac Gilead. Atar Livneh is Lecturer in the Department of Bible, Archaeology, and Ancient Near Eastern Studies, BenGurion University of the Negev. As part of her doctoral dissertation (Haifa University), Livneh edited and annotated several Qumran scrolls. She published a number of articles devoted to the interpretation of legal and narrative material in the Book of Jubilees. One of her primary fields of interest is family structure and roles during the Second Temple period, as reflected in such texts as Philo’s De Abrahamo. Over the past two years, she has been engaged in a comparative study of some forty Jewish and Christian Second Temple texts, which summarize Israelite history. The goal is to arrive at new insights into the form, content, and context of the historiographical genre during this period.

Daniel Kaufman completed his dissertation research at Southern Methodist University (Dallas). He is Professor Emeritus in the Department of Archaeology at the University of Haifa, and he is an active researcher at the Zinman Institute of Archaeology at the University of Haifa. His two major periods of interest are the Middle Paleolithic and Epipaleolithic of the Levant. He is currently involved in the excavation of the Natufian occupation of el-Wad Terrace. His areas of specialization are the studies of lithic assemblages through the use of statistical analyses with an emphasis on the application of diversity indices.

Dan Malkinson received his PhD from the Hebrew University of Jerusalem in Ecology. Currently he is a researcher at the Department of Geography and Environmental Studies, and at the Golan Research Institute at the University of Haifa. His fields of research include human – wildlife interactions, the effects of humans on ecological processes and patterns in natural and human dominated environments. He applies spatial statistics and analysis tools to address these research questions.

Hamoudi Khalaily is currently responsible for field technology and archaeological documentation at the Israel Antiquities Authority (IAA). He served as deputy director of Excavations and Surveys for the IAA from 2011 to 2016. He joined the IAA in 1993 as a prehistoric archaeologist. Between 2000 and 2010, he was twice the head of the Prehistory Branch of the IAA. Khalaily received his PhD in Archaeology from Ben-Gurion University in 2006. Prior to joining the IAA in 1993 he was a high school teacher in Sachnin (Galilee) and a researcher at the Centre de Recherche Français à Jérusalem. Khalaily’s research interests are stone tools and South Levantine prehistory. Khalaily co-directed several major excavations including Abu Ghosh, Ain Mallaha, Motza, Yiftaḥ’el, Beisamoun, and Hagoshrim He directed more than 22 salvage excavations in prehistoric and proto- historic sites all over Israel. He is author and co-author of four monographs and over 130 professional papers.

Ofer Marder is Associate Professor and Chair of the Department of Bible, Archaeology and Ancient Near Eastern Studies at Ben-Gurion University of the Negev in Beer Sheva. He has directed a large number of prehistoric excavations and research projects ranging from Early Paleolithic through Epipaleolithic and Neolithic. He is co-director of the Manot Cave excavations, one of the most significant sites in the Levant for understanding the emergence of modern humans outside Africa and their dispersal to Europe. In addition to work done in Israel, Marder was a member of research expeditions working in East Africa and in China. His main topics of research are the rise of sedentary societies and the emergence of agriculture both in the Levant and China; lithic technology, including the refitting and technotypological analysis of cultural traditions during the Upper Paleolithic and Epipaleolithic (45,000 to 10,000 BP); and the study of human adaptations to ancient environments during different periods in diverse ecological and geographical settings.

Liora Kolska Horwitz is Curator of the Life Sciences Collection in the Faculty of Science at the Hebrew University of Jerusalem. Her research interests include Paleoecology of vertebrates, Archeozoology of the Levant, Taphonomy of Mammals, and Archeozoology of

Ianir Milevski, PhD, is the Head of the Prehistoric Branch at the Israel Antiquities Authority. He also x

teaches archaeology in universities in Argentina in the framework of a special program of the Ministry of Science and Technology of Argentina. He has published several books and dozens of articles on socio-economic aspects of the late Prehistory of the Southern Levant.

worked extensively on chipped stone tool assemblages, especially those from the early historical periods, pioneering the study of stone tools in the Metal Ages. He has published six books and over 200 professional papers. He is a member of the Archaeological Council of Israel, the board of directors of the Israel Prehistoric Society, the editorial board of Paléorient, and past editor of the Journal of the Israel Prehistoric Society. He currently serves as Vice President for External Affairs at BenGurion University of the Negev.

Dani Nadel received his PhD from the Hebrew University of Jerusalem. Currently he is Professor at the Zinman Institute of Archaeology at the University of Haifa in Israel. His areas of research and teaching are prehistoric archaeology, lithic technology, and ancient game traps. He has excavated at several sites, including Ohalo II, a 23,000 year-old fisher-hunter-gatherers’ camp on the shore of the Sea of Galilee, and Raqefet Cave, a Natufian burial site. He is now excavating ‘desert kites’, which are ancient large-scale game traps in the Negev and Armenia.

Danny Rosenberg is Professor and Head of the Department of Archeology at the University of Haifa, Head of the Laboratory for Ground Stone Tools Research at the Zinman Institute of Archeology, and Chair of the International Association for Ground Stone Tools Research. He leads an interdisciplinary project for tracing the emergence of the Mediterranean diet in the Eastern Mediterranean – culinary traditions and their environmental, economic and cultural background. He heads projects that study food remains and complementary food processing tools from prehistoric and later sites in Israel. Currently, Rosenberg‘s prime focus is the archaeological site of Tel Tsaf in the Jordan Valley, where tens of thousands of organic remains are carefully preserved for over 7,000 years. Professor Rosenberg also leads another project that makes use of extensive geochemical analyses to locate sources of raw materials used for producing food processing tools and other stone tools.

Sarit Paz, PhD, is an archaeologist specializing in the Early Bronze Age of the Southern Levant and the South Caucasus. She is co-director of excavations at Tel Bareqet and Tel Bet Yerah, Israel, and director of the Kvatskhelebi research project in Georgia. Yitzhak Paz is a Senior Researcher at the Israel Antiquities Authority, specializing in proto-historic periods, mainly the Early Bronze Age, and also in landscape archaeology, megalithic phenomena and theoretical aspects of material culture. Gideon Ragolski is Coordinator of the Dead Sea-Arava Science Center in Eilat, Israel. A resident of Paran, Israel, he is a desert tour guide, and he analyses archaeological remains in the central Arava region of Israel.

Dina Shalem is a researching archaeologist and editor at the Israel Antiquities Authority, and an adjunct lecturer at the Kinneret Academic College. She has directed a number of excavations in various sites in northern Israel dating from the Chalcolithic to the early Islamic periods. She has co-directed the excavation at the Late Chalcolithic secondary burial cave at Peqi‘in, Upper Galilee and has recently joined the team at ‘Ein Asawir to lead the excavations of the Chalcolithic and Neolithic strata. She has also co-directed the survey of the Shomera map, as part of the Israel Survey, and is co-organizer of the yearly meeting of the Chalcolithic Forum. Her main area of research is northern Israel, especially the Galilee, and her main topics of interest are the Late Chalcolithic period, its emergence, the characteristics of its extraordinary material culture and the social viewpoint; the sequence of the earlier Chalcolithic periods up to the Late Chalcolithic as well as the end of the period and the transition to the Early Bronze Age; prehistoric art, ritual and burial customs; and the Galilee as a geographic entity. She is the cofounder and co-publisher with Zvi Gal of Ostracon Press, which publishes books on archaeological topics.

Avraham Ronen is Professor Emeritus of Prehistory in the Department of Archaeology at the University of Haifa. He is the Founder of the Zinman Institute of Archaeology at the University of Haifa. His major projects and excavations are the archaeological survey of Mount Carmel, Tabun Cave, Evron, Ruhama, Sefunim, Hatula, and Nahal Reuel. His major research topics are stone tool technology and typology; human adaptation; and social behavior. Avraham Ronen was born July 4, 1935 in Jerusalem. He passed away December 15, 2018 in Haifa. Steven A. Rosen is Canada Professor of Ancient Near Eastern Archaeology in the Department of Bible, Archaeology, and Ancient Near Eastern Studies at Ben-Gurion University. He received his doctorate from the University of Chicago and worked as a field archaeologist for the Negev Emergency Survey before moving to Ben-Gurion University full-time in 1988. He has excavated numerous sites in the Negev, ranging from the Paleolithic through recent Bedouin occupations, but focusing especially on the mobile pastoral societies of the region. His research has also

Anastasia Shapiro, PhD, is a researcher at the Israel Antiquities Authority [IAA] (Northern district). Her research interests include petrographic and petrologic xi

researches of pottery; field geological and mineralogical consulting; archaeological surveying including Geographic Information System (GIS) and mapping. She received her MA in engineering with honors in Research Geology of Rare Elements from the Moscow Geological Prospecting Institute in 1986. Prior to repatriation to Israel in 1991, she worked as a laboratory assistant at the department of rare elements of the Institute of Geology and Mineralogy of the Academy of Science of the USSR (1988-1989), and as engineer-geologist at the Moscow Institute of Rare Elements Geology (19861988). During her work at the IAA (since 1992) Shapiro participated as registrar and area supervisor in a number of archaeological excavations. Since 2004 she is responsible for GIS and mapping for the northern district. She took part in international projects such as The Levantine Ceramics Project and the European project POMEDOR (people, pottery and food in the Medieval Eastern Mediterranean). She is a partner in the publication of the results of the excavation of the 12th century monastery of St. Mary of Carmel at Wadi esSiah (2011) on behalf of the Carmelite Church. Shapiro is author and co-author of more than 30 professional papers including international projects of excavations at Shikhin, Tel Ashkelon, and Tel Kedesh.

addition to Epipaleolithic and Neolithic sites in the arid region of the Negev. Recently, Vardi has focused on the research of craft specialization, mass production contexts of lithic tools, and the research of quarries and fabrication sites of stone tools. He also studies the mobility patterns of foraging groups in the Negev. Eran Viezel is Senior Lecturer in the Department of Bible, Archaeology and Ancient Near Eastern Studies at Ben-Gurion University of the Negev and Head of the Bible and Ancient Near East Division in that department. Viezel is the author of The Commentary on Chronicles Attributed to Rashi (Jerusalem: Magnes Press, 2010) and numerous journal articles in biblical exegesis, biblical realia, higher criticism, and the history of Jewish biblical exegesis. For his novel, In Praise of Loneliness, he was awarded the 2015 Aharon Ashman Prize for Creative Writing Submitted Anonymously by The Society of Authors, Composers and Music Publishers in Israel. Allan Witztum completed his doctorate at Cornell University in 1966 and had a research fellowship in the department of Botany at the Hebrew University for the next two years. In 1968 he joined what is now called the Ben-Gurion University of the Negev where he taught the first graduating class in Biology. He retired as a Professor Emeritus in 2007. His interests include plant anatomy, seed dispersal, plant biomechanics and economic botany (useful plants). He remains interested in the many plant genera he studied over the years including Lemna, Spirodela, Cucumis, Blepharis, Ruellia, Centaurea and Typha.

Moti Shem-Tov is a tour guide in Eilat, Israel. He discovered most of the Rodedian sites in the Negev of Israel. Rahamim Shem-Tov is a technician at the Geological Survey of Israel and a research assistant at the Dead Sea-Arava Science Center in Eilat, Israel. He works with Uzi Avner and other archaeologists on various projects.

Yuval Yekutieli is Senior Lecturer in the Department of Bible, Archaeology and Ancient Near Eastern Studies at Ben-Gurion University of the Negev. His research interests include international relations during the Southern Levantine Bronze Ages; Ancient colonialism; Archaeology of slavery and bonded labor; Proto-urbanism; Arid-zones archaeology; Ancient pastoralism; Landscape archaeology; Operation of power in antiquity.

John D. Speth is Professor Emeritus of Anthropology at the University of Michigan at Ann Arbor (USA). He studies hunter-gatherers, focusing on diet, subsistence strategies, food processing technologies, and how foragers cope with unpredictable resources. Oded Tammuz received his PhD in Assyriology from Yale University in 1994. He is a Senior Lecturer in the department of Bible, Archaeology, and Ancient Near Eastern Studies at Ben-Gurion University of the Negev. His main field of interest and research is Ancient History of the Levant and the Eastern Mediterranean Basin.

Shamir Yona is Associate Professor in the Department of Bible, Archaeology and Ancient Near Eastern Studies at Ben-Gurion University of the Negev. He served for two terms as Chair of that department and two terms as Head of the Bible and Ancient Near Eastern Studies Division in that department. He is Co-chair with Elad Filler of the Judaica Section for Society of Biblical Literature International Meetings. He is a member of the board of directors of the New Israel Society for Biblical Research. Yona’s publications focus on the poetics of Ancient Near Eastern literature; the wisdom literature of the Bible, the ancient Near East, and rabbinic literature; biblical prophecy and narrative; Ugaritic language and literature; and Semitic lexicography.

François R. Valla is a researcher at the Centre National de la Recherché. He has devoted most of his career to the Natufian Culture in Israel. He conducted excavations at Ain Mallaha (Eynan) and at Hayonim Terrace. Jacob Vardi is a researcher in the prehistoric branch of the Israel Antiquities Authority Research Division. He has directed prehistoric excavations in late prehistoric sites of the Neolithic period both in the Mediterranean regions of the Galilee and in the Judean mountains in xii

Preface and Acknowledgments

As we present this Festschrift to our colleague, Isaac (Itzik) Gilead on the occasion of his seventieth birthday, we are delighted to thank all of the people who enable us to bring this project to fruition.

We thank the honoree, Professor Isaac Gilead, for providing both the photograph of the honoree, which appears within the volume, and the cover illustration, which depicts the staff and the students during the 2006 season of the excavations at Mitham C in Beer Sheva, which revealed levels beginning with the Chalcolithic Period and extending to the Byzantine Era.

Four editors—Professor Haim Goldfus, Professor Mayer I. Gruber, Professor Shamir Yona, and Dr. Peter Fabian— friends and colleagues of Isaac Gilead in the Department of Bible, Archaeology, and Ancient Near Eastern Studies worked tirelessly and unstintingly to complete this project.

Finally, we thank the forty authors, who include the honoree’s teachers, colleagues, and students. These authors, who are associated with research institutions in Israel and around the world, kindly responded to the editors’ invitation to contribute to this collection of articles. The honoree and the project have been blessed by their fascinating and original contributions to scholarship.

We are especially grateful to Dr. David Davison and Dr. Rajka Makjanić, the Founders and Directors of Archaeopress and Dr. Dan Stott, of the marketing division of Archaeopress and the outstanding staff of Archaeopress for their dedicated work in publishing Isaac Went Out to ... the Field. We are delighted by the beautiful format of this Festschrift. We extend special thanks to Darko Jerko for his meticulous work in the formatting and the final editing of this beautiful book.

The reader will note that 24 of the articles are in English, and four of the articles are in Hebrew. The volume also includes English abstracts of the Hebrew articles. The articles, which reflect the variety of subjects of interest to the honoree and his colleagues and friends, have not been arranged according to subjects. Articles in both the English section and the Hebrew section are arranged in the alphabetical order of the first-named author of each article.

We express our thanks to the graphic artist, Ms. Sefi Sinay for her tireless efforts in formatting the Hebrew section of this book. We are most grateful to Dr. Eli Cohen Sasson, graphic artist in the Department of Bible, Archaeology and Ancient Near Eastern Studies at Ben-Gurion University for editing the illustrations and performing many other tasks, both great and small, wholeheartedly and pleasantly.

The Editorial Board Haim Goldfus, Mayer I. Gruber, Shamir Yona, and Peter Fabian

xiii

xiv

Isaac Gilead: An Appreciation

Isaac (Itzik) Gilead was born in 1948 in Bamberg, Germany. He was the first born son of parents who came from the town of Milawa in Poland.

Itzik wrote his MA thesis Layer I-15 in the Lower Palaeolithic site of ‘Ubeidiya and his PhD dissertation on The Upper Palaeolithic in the Negev and Sinai: sites in Gebel Maghara, Kadesh Barnea and Nahal Zin under the supervision of Prof. O. Bar-Yosef.

Itzik and his family came to Israel in 1949 and settled in Jerusalem. During Itzik’s childhood the family lived in the Kerem Avraham neighborhood.

Isaac Gilead began his work in archaeology in 1972 as an assistant at the Institute of Archaeology at the Hebrew University. He gained field experience in the projects of ‘Ubeidiya and surveys in Sinai. From 1977 he began teaching in the Department of Bible, Archaeology and Ancient Near Eastern Studies. In fact, from 1972 until his retirement 2016, Itzik taught only in that department. In 1982 he was appointed Lecturer, and he climbed the ladder to the rank of full professor, which he was granted in the year 2000.

Itzik’s parents enrolled him in the Tachkemoni School, which was an all-boys school, which belong to the national religious educational stream. The choice of this school did not reflect the family’s ideology but only the proximity of this particular school to the family’s home. When his schoolmates found out that Itzik was born in Germany, the children referred to this child of holocaust survivors as ‘Nazi’.

Gilead’s archaeological work concentrated in the Negev and Sinai, and includes excavations of sites belonging to many different periods, including Upper Palaeolithic sites at Kadesh Barnea and the Late Neolithic site of Tel Qatif in the Sinai, and a series of sites in the Negev, such as the Middle Palaeolithic site of Fara’h II, the Chalcolithic sites of Grar, the Nahal Sekher area, Nahal Besor, Abu-Matar, Beer Sheva, Nevatim, and Tel Sheva and the Early Bronze Age site of Tell ’Erani. He is the author of Grar – a Chalcolithic site in the Northern Negev (Beer Sheva: Ben-Gurion University Press, 1995), as well as dozens of articles, most of which deal with the Upper Palaeolithic and Chalcolithic periods. Since 1975, he is a teacher and a Ph.D. and MA advisor to a cohort of archaeologists, many of them now teaching and leading archaeological research, in Israel and around the world.

Itzik loves all kinds of music. He is especially familiar with popular music both Israeli and international. However, he is especially fond of classical music. In fact, Itzik was gifted from an early age with an unusual Itzik is an avid reader, and his knowledge is both wide and deep and spans many disciplines. When he was a high school student he worked as a salesperson in a book store on Geulah Street in Jerusalem. With the money he earned he bought books and tickets to see motion pictures. Because of his great administrative skills the owners of the book store left him in charge of the store while they went on a two-week summer vacation. In August 1966 Itzik was drafted into the Israel Defense Forces for a period of two years and two months. However, in the course of his initial period of service he was informed that his period of service had been extended, first for two and a half years and later to three years. He served for three years in the Engineering Corps. He spent a significant part of his military service in various courses including an officers training course. Itzik was honorably discharged from regular military service at the rank of lieutenant.

During his years at BGU, Isaac Gilead served the university in many administrative functions, including as the head of the Archaeological Division (1980/81; 1994/95; 1999/2000; 2002-2009), Chairman of the Department of Bible, Archaeology and Ancient Near Eastern Studies (1987-1991); Chairman of the University Library Committee (1994-2000); Member of the University’s Centralizing Committee (19951998); Vice Dean of College Affairs (1997-2000); and Director of the Kreitman Foundation (2007-2009). As an archeologist, he held various offices in the Israel Prehistoric Society, the Israel Association of Archaeologists, the International Union of Prehistoric and Protohistoric Sciences (UISPP), the Archaeological Council of Israel, the Council for the Preservation of Built Heritage, and in the Israel Antiquities Authority, as a member of the Directory Council (1990-2001).

During the Yom Kippur War of 1973 Itzik was called up for reserve duty in the army engineers’ unit that set up the pontoon bridge across the Suez Canal. It was across this bridge that most of the Israeli forces crossed the Suez Canal into Egypt. In October 1970, a short while after Itzik was discharged from his active military service, he began to study archaeology, world history, and Jewish history at the Hebrew University of Jerusalem. He completed his B.A. in 1974, his M.A. in 1977, and his Ph.D. in 1982. xv

Isaac Gilead has mentored numerous graduate students, and he continues to mentor both M.A. and Ph.D. students. Many of his students teach in institutions of higher learning, and they are at the forefront of archaeological research in Israel and around the world. Two of his former graduate students are faculty members in the Department of Bible, Archaeology and

Ancient Near Eastern Studies at Ben-Gurion University of the Negev. The editors and contributors to Isaac Went out to the Field wish Itzik many many more years of health and happiness and more fruitful research. The Editors

xvi

The Religious Dimension of Copper Metallurgy in the Southern Levant Nissim Amzallag Introduction

that the technology [= metallurgy] was understood not only in practical terms, but also conceived in the realm of ideas, symbols and beliefs ... This is why we argue that the technology itself – the production process, from preparing the smelting to the finished artifact – was ritualized’.

The emergence of metallurgy was an important factor in the development of Ancient Near Eastern societies. Metal became the raw material for the production of tools, utilitarian implements, jewels, items of prestige, objects of art, and ritual artifacts. Its imperishability opened up new possibilities with respect to the concentration of wealth and power. The control of metal production and trade contributed to the emergence of the first important colonial movements in the Ancient World, the Uruk expansion (Algaze 1989: 581, 584; Butterlin 2003: 353-357; Avilova 2008: 88) and the Egyptian colonization of Nubia (Adams 1984), Sinai and Canaan (de Miroschedji 1998; Butterlin 2003: 156158). The diffusion of metallurgy from the Ancient Near East created networks of metal production and trade in Europe (Brodie 1997: 305-309; Kristiansen and Larsson 2005: 43-51; Brück 2011: 387-388) and in Asia (Chernykh, 1992: 140-171; Kohl 2007: 29-48, 146-150), which propelled the circulation of knowledge, ideas and religious concepts all over Bronze Age societies. Despite these considerations, metallurgy is rarely approached as an important factor in the development of the Bronze Age religions, in comparison with the sun, atmospheric elements, fertility, crop production, water and the subterranean universe.

This idea is not as surprising as it may appear at first sight. Native copper is not found in the Southern Levant. For this reason, local copper necessarily originates from the reduction of ore. This process was long and difficult at the end of the fifth millennium BC: it required mining, transporting the ores from desert regions (the Arabah valley), smelting, crushing the slag, and separating the copper from the mineral fraction by remelting the crushed slag (Shugar 2003). The development of such a process may hardly be justified on the basis of any practical application. After all, copper is rare relative to flint, and flint tools are easy to produce. Furthermore, flint is harder than the copper produced from the Arabah ore. Thus, no improvement of tools may emerge from the earliest stages of metal production. For these reasons, the simplest justification for the development of metallurgy in the Southern Levant is that metal production was a process that held significance in and of itself. Indeed, this is exactly what Mircea Eliade (1968: 76) already argued:

The situation differs when looking at the Ghassulian culture (ca. 4500-3900 BC), in which are attested the earliest stages of metallurgy in the Southern Levant. This culture is characterized by extensive production of prestige metal artifacts in regard to utilitarian items. Apparently, this bias was not accompanied by any improvement in social hierarchization, so that Rosen (2002: 16) concluded: ‘there does not seem to be any obvious connection between elite control and copper production in this early period’. Prestige artifacts were not produced in the Southern Levant to be traded in distant countries. Given this, the simplest justification for their production is to assume a ritual function of prestige artifacts among the Ghassulians (Gilead 2002). This view is supported by the transformations in rituals and burial practices concomitant to the emergence of metallurgy in the Ghassulian culture (ca. 4500-3900 BC) (Gošić 2013 [esp. 281-284] and 2015). Metallurgy itself, beyond any utilitarian perspective of application, may have held religious importance. ‘The symbolic role of artifacts, argued Gošić and Gilead (2015a: 169), suggests

The discovery of metals and the progress of metallurgy radically modified the human mode of being in the universe. Not only did the manipulation of metals contribute considerably to man’s conquest of the material world; it also changed his world of meaning. The metals opened for him a new mythological and religious universe. Metallurgy is not an isolated case. Ceramic technology also seems to have emerged because of non-practical motivations (Rosen 2002). Even agriculture and herding apparently developed because of their resonance in the universe of beliefs rather than because of their advantage over gathering and hunting (Cauvin 2002: 245-247). In these latter instances, the ritual dimension that stimulated the earliest stages of development has practically disappeared in the course of time. It remains a crucial factor in the emergence of these techniques, at a stage in which the material perspectives and advantages may hardly justify their development. However, this archaic stage becomes progressively 1

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead (Amzallag, in press) subordinated to the extensive use of the technique and its standardization in other spheres of life.

including the gods (Finnestad 1976: 102; Gordon and Gordon 1996). Furthermore, the Egyptians attributed a metallic nature to the flesh (gold) and bones (silver) of the deities (Aufrère 1991: 310-313). It is likely, therefore, that the deity producing the metals (= the smelting god) was approached by the Egyptians as the genuine creator and father of all the gods. His/her silenced identity suggests that this greatest god was not worshipped in the official cult, but rather through an esoteric worship.

In the Southern Levant, copper production substantially increased during the Early Bronze Age (Adams 2002). The metal produced was mainly exported to the Nile valley (together with metallurgy itself), a process which promoted social transformations and urbanization (Kempinski 1989: 165-166; Gophna and Milevski 2003; de Miroschedji 1998: 22-26). At the same time, the production of prestige artifacts using the complex technique of lost-wax casting ceased. Apparently, as with other techniques, the ritualized stage of metallurgy disappeared almost completely after this technique reached maturity, and after the production and trade of copper became intensified. Alternately, it may be proposed that the ritual dimension of metallurgy did not disappear. Rather, it simply became an esoteric and hidden fundament of the religions in the Southern Levant, and more generally, in the Ancient Near East. The aim of this paper is to investigate this premise.

Mesopotamia The esoteric nature of metallurgy is revealed, in Mesopotamia, through the secret rituals inherent in the construction of furnaces and in metallurgical activity (Eliade 1977: 60-64). The impact of this cultural dimension of metallurgy on the official religion is reflected in the smith god’s multiplicity of functions. Ea/Enki was considered the guardian of the secret knowledge, a feature justifying his status as god of witchcraft, exorcism, divination and magic powers (Lenzi 2008: 104-105; Schwemer 2015: 41-42; Galter 2015: 66-69). His homolog, Nusku, was similarly considered the guardian of the secret knowledge that he revealed to his devotees, the metalworkers and diviners (Lenzi 2008: 53, 350). Ningizzida, the patron deity of Gudea and of metalworking, was even ‘... a mystic underworld divinity, who protected the living by his magic spells, and could ward off death and heal disease for the benefit of those who worshipped him devoutly’ (van Buren 1934: 89). The healing/vitalizing function, attributed to Ea/Enki and the parallel between smith-god deities (Jayne 1962: 118-121), confirms the essential, though silenced, importance of metallurgy in Mesopotamian religion.

The esoteric dimension of metallurgy in the Bronze Age The religious importance of metallurgy is difficult to observe in the Southern Levant in the Bronze Age, mainly because the smelting god and companion deities involved in metallurgy are not identified. Consequently, we must examine the religious importance of metallurgy in neighboring cultures, such as Egypt, Mesopotamia and the Aegean. Egypt From the Old Kingdom period, the economy of precious metals was officially justified by the need to enrich the temples with gold and other precious metals (Aufrère 1991: 317-320). Apparently, this motivation might be interpreted as the wish to honor the gods with highly valued items and materials. However, this explanation does not account for the frequent addition of precious minerals, metals and metallic ores in the foundations of the Egyptian temples. As suggested by Aufrère (1991: 193), this practice promoted a symbolic homology between the sanctuaries and the mines from where metallic ores were extracted. By extension, this practice shows that the Egyptians approached metallurgy as the hidden foundation of the sanctuaries. The addition of precious metals to the foundations of temples represents, therefore, an esoteric layer underlying the official religion.

Minoan Crete In addition to urban temples, the Minoan religion is characterized by mountain peak sanctuaries with caves where the great god of Crete was worshipped. The hidden nature of this cult is reflected in the uncertainty surrounding the identity, name and attributes of this great deity. The restricted access to his peak sanctuaries, and the function of these in initiation rituals (Faure 1997: 297-300) are further indications of the esoteric dimension of this cult. The great mysterious god of Crete has been identified with Welkhanos (Bloedow 1991: 166; Capdeville 1995: 166-167). His metallurgical affinities are deduced from his homology with Vulcan, the Roman patron of metalworkers, and with Hephaestus, and the patron of the Cyprus smelters (see Capdeville 1995). This is confirmed by the central involvement of metalworkers in the rituals performed in these peak sanctuaries (Capdeville 1995: 187-191; Faure 1997: 170173; Blakely 2006: 13), and by the bronze altar and the hoard of copper ritual artifacts (double axes) found in the peak sanctuary at Mount Iouktas (Bloedow 1991:

Further indications confirm the religious importance of metallurgy. Ptah, the patron of the Egyptian metalworkers, was approached as the master of the ka, the principle that vitalized the whole universe, 2

N. Amzallag: The Religious Dimension of Copper Metallurgy in the Southern Levant 160-163). These data, together, support the assumption of an esoteric cult of metallurgical nature underlying the official Minoan religion.

acknowledged for a long time (Lucas 1948; Meyerowitz 1960; Rowlands 1971). Recent investigations have identified parallels between the social status of African blacksmiths and of Iron Age metalworkers from the Southern Levant (Qenites) and from the Aegean (Dactyls, Kuretes, Cyclops, Telchines, Kabeiroi, Korybants) (McNutt 1990, 1999; Blakely 2006; Pfoh 2014). These parallels are especially interesting when we consider the central importance devoted to metallurgy in religions from traditional Africa. This reality is subsumed by Dominique Zahan (1979: 30) as follows:

Bronze Age Europe As in the Southern Levant, the rise of metallurgy was in Europe closely related to deep transformations. ‘The rise of chiefdoms, argued Kristansen and Larsson (2005: 52), often corresponds to an increased development of metallurgical skill and a whole new set of myths and gods linked to the sacred role of mining, smithing and ritual transformation’. The religious dimension of metallurgy is confirmed by the fact that many Bronze Age cultures from Europe are characterized by abundance of metallic prestige artifacts with cosmologic significance. These artifacts include axes, sun chariots, scepters, solar discs, crossed circles, and metallic representations of the firmament (e.g. Davidson 1969: 174; Brück 2011: 389-392; Ionescu 2012: 159; Scarano and Maggiulli 2014). The metallic nature of these ritual artifacts is interpreted as being involved in the representation of the Universe (Kristiansen and Larsson 2005: 294-303).

The forge is often a place of worship, too. Its ground is sacred, and one enters barefoot in order not to communicate to the ‘temple’ the impurity of the shoe. No dispute is tolerated there, not because of the ‘spirits’ which live there but because it represents a celestial space [...] The artisan-priest’s mediation with the invisible powers takes place in the enclosure of this workshop-temple. In addition, the forge offers a place of refuge for unfortunates seeking asylum. Connected to the notions of fecundity, life and liberty, it is the most typical sanctuary in African religion [...] If the comparison were not so extreme, we could almost say that the forge is the church of the African village.

The metallurgical affinities of many burial practices, such as the furnace/burials homology, or cremation in areas of metallurgical activities (Dieterle 1987: 5; Goldhahn and Oestigaard 2007: 217-219), support the assumption that metal production per se, and not simply the metal as raw material of precious value, was of ritual importance in Bronze Age Europe. A comparative analysis of ancient mythologies confirms that, in the Bronze Age, metalworkers were regarded as masters of occult sciences, instructors and counselors to the political and religious elite (Kristiansen and Larsson 2005: 52-53). Here too, cultural metallurgy appears as the esoteric foundation underlying the official power.

The prominent status of the forge is confirmed by the multiplicity of the blacksmith’s functions in Africa. Beyond his craft, this artisan was also the poet, musician, healer, judge, diviner, initiator, grave-digger, rain maker and well sinker of the community (de Maret 1980: 273; Boyer 1983: 49; Reid and McLean 1995: 153). Two elements suggest that this religious dimension is anchored in ancient traditions. The first is the frequent identification of the primordial smith as the civilizing hero, especially in Western Africa (Tegnaeus 1950: 16109). The second is the higher symbolic dimension attached to the process of metal production (smelting) vis-à-vis the production of metal implements (metalworking) (de Maret 1980: 269; Boyer 1983: 46; Blakely 2006: 68). The parallels existing between African blacksmiths and metalworkers from the Southern Levant and the Aegean suggest, by extension, that a part of the ritual dimension of metallurgy was maintained at the Iron Age, both in the Southern Levant and in the Aegean.

These observations hint that the ritualized dimension of metallurgy constituted a fundamental feature of many Bronze Age religions (Budd and Taylor 1995; Kristiansen and Larsson 2005: 49-56; Avilova 2012). This means that it survived the first phase of intensification of production and trade of copper (mid-fourth millennium BC), and even the second phase inherent to the exploitation of copper sulphide ores (mid third millennium BC). The latter development considerably extended the use of copper for production of tools and other usual implements. For these reasons, it is likely that also in the Southern Levant, this ritualized dimension of copper metallurgy had also survived the collapse of the Ghassulian culture.

Iron is by far the most common metal in Africa. Consequently, most of the rituals described by anthropologists belong to iron smelting and working. One of the most common features of African iron metallurgy is the female/womb symbolism of the furnace (the blowing apparatus being therefore identified with male genitals). Smelting, therefore, is likened to procreation (Eliade 1977: 48-53; de Maret 1980: 275; Reid and McLean 1995: 149; Blakely 2006: 99-

The African perspective Similarities between metallurgical traditions from the ancient Near East and central Africa have been 3

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead 104; Gosic and Gilead 2015b: 32-34). In the Ancient Near East, however, this kind of symbolism is not attested either in iron metallurgy or in copper metallurgy that preceded it. Apparently, something fundamental differs in the way iron metallurgy was approached in the Ancient Near East and in traditional Africa. If this is so, it may be that the common features concerning the cultural dimension of metallurgy in Africa and in the Ancient Near East are rooted in copper rather than iron metallurgy traditions.

of beliefs of the Dogons, their way of thinking and even their social organization. Parallels identified between this esoteric metallurgical tradition and rituals from Antiquity (Lambert 1980) suggest that in the Iron Age a similar distinction existed between the ritual value of copper and of iron metallurgy. Figures of the smelting god in the Iron Age The apparent extension to the Iron Age of the esoteric dimension of copper metallurgy, together with the renewed production of copper in the southern Levant (Levy et al. 2012; Ben Yosef et al. 2012), invites us to look for expressions of the (hidden) smelting god in the first millennium BC.

Copper ore is extremely scarce in Africa. As a result, the smelting and working of copper is very limited there. Nevertheless, copper, and especially pure copper, displays a special and far more prestigious status than iron in western Africa. It is globally approached as the ‘holy material’, and it is well researched (more so than gold) for its apotropaic, vitalizing and magical properties (Herbert 1973). This preferential status given to copper has a parallel in the Ancient Near East. In the Iron Age, copper was abundantly found in sanctuaries while iron was frequently absent. This bias is even attested in Assyria, the first empire whose military power was built on iron weapons (Pleiner and Bjorkman 1974). In Egypt, Israel and Greece, iron is explicitly excluded from the holy sphere (McNutt 1990: 147-148, 209-218; Blakely 2006: 200).

Dionysus and the metallurgical mysteries The mythology of Hephaestus, one of the figures of the Greek smith god during the first millennium BCE, encloses an esoteric dimension reflected by his nature as magician god, his infirmity (limping), and his nine-year subterranean initiation which lead to his integration into the Greek pantheon (Martin 2005: 1720). The affinities between Hephaestus and Welkhanos (Capdeville 1995: 275-282) confirm the preservation in the Aegean of the esoteric metallurgical traditions identified at the Bronze Age. The importance of Lemnos and Naxos in the initiatory apprenticeship of Hephaestus suggests that these ancient esoteric traditions were especially well preserved in these islands.

The preference for copper over iron is difficult to justify if metals were simply introduced in sanctuaries for their utilitarian properties, preciousness, or esthetic considerations. Rather, it suggests that in the Ancient Near East in general, and in the southern Levant in particular, the ritualized dimension of copper was preserved in the Iron Age. The reason why iron, the newly-smelt metal, was excluded from the holy sphere is never explicitly justified. We may deduce, here again, that the exclusion is based upon esoteric considerations relative to the ritual dimension of metallurgy.

The island of Naxos is also named Dionysia in reference to its patron-god, Dionysus. Furthermore, Dionysus is praised in the first Homeric hymn for being the god who brought Hephaestus into the Pantheon at the end of his initiation. These features designate Dionysus as the master of Hephaestus’s skill and apprenticeship to the esoteric dimension of metallurgy. Dionysus is subsumed by Euripides (The Cretans, frag. 475) into the great god of Minoan Crete, a feature confirmed by further details of his mythology (Kerenyi 1976: 113). His identification with Welkhanos is strengthened by the parallel between the union of the Cretan princess Ariane (one of the appellations of the great Cretan goddess) with Dionysus and her close relation with Welkhanos (Capdeville 1995: 180). The association of Dionysus with metallurgical traditions is confirmed by his appellation as the ‘father’ of the Kabeiroi guild of metalworkers (Schachter 1986 vol. 1: 189-190, vol. 2: 93-95). At Thebes, members of this guild (also identified as the sons of Hephaestus) devoted to Dionysus an esoteric cult, which was distinct from the public worship of the deity (Schachter 1986 vol. 2: 96; Freyburger-Galland 2006: 104-109). Furthermore, the Kabeiroi were the guardians and promoters of religious mysteries practiced in Samothrace, Lemnos and Thebes from the early first millennium BCE (Blakely

This conclusion is supported, again, by examination of some of the African metallurgical traditions. For example, among the Kapsiki, a tribe living in the Chad basin, iron metallurgy is an ‘open’ activity performed in the village while copper working remains a secret activity occurring far from the village. In fact, it is closely related to initiation rituals and esoteric knowledge (van Beek, 1991: 293-298). A glimpse at Dogon traditions confirms this view. Here too, the smelting and working of iron, the utilitarian metal, are ritualized activities. However, copper (especially in a molten state = the divine ‘water’) and not iron, is positioned at the center of the cosmogony, mythic history and primeval events. This centrality of copper is especially reflected in the esoteric tradition of the Dogons, in which iron is only of minor importance (Griaule 1948). Copper metallurgy represents therefore the esoteric fundament, which organizes the universe 4

N. Amzallag: The Religious Dimension of Copper Metallurgy in the Southern Levant 2006: 13-40; Bremmer 2014: 46-47). Considered the most ancient mysteries in Greece (Pausinias 1.4.6), it is likely that these cults prolong, at least partly, the Bronze Age esoteric metallurgical traditions. Though the nature of these Kabeiroi mysteries remained almost totally hidden, their content differed from the official religion in their emphasis on moral virtues and the systematic absence of figuration of deities. Kerinyi (1955: 35) characterizes these mysteries as

••

... a ritual action that is not bound up with a cult image of the godhead, as in the case of Kallynteria and Plynteria, but with the people who through action become in some special way the object and subject of the festival. The mystes (μύστης) suffers the mysteries, he becomes their object, but he also takes an active part in them. According to the popularity of the Kabeiroi mysteries, especially celebrated at Samothrace, it seems that the esoteric metallurgical knowledge inherited from the Bronze Age metallurgical traditions became diffused by the Kabeiroi far beyond its initial context and audience.

••

These findings support the identification of Dionysus with the Bronze Age smelting god. Accordingly, the extensive popularity of Dionysus in the Mediterranean area, in the first millennium BCE, reflects the metamorphosis of this mysterious Bronze Age deity into a god openly worshipped by everyone, and at the same time disconnected from his metallurgical context.

••

The parallels between YHWH and Dionysus Michael Astour was one of the first scholars to identify substantial parallels between the cult of YHWH in ancient Israel and that of Dionysus in Greece. He emphasized, for example, their similar ecstatic prophecy and its akin contagious nature, the affinities between the Dionysian sparagmos and the ritual related in 1 Sam 15:33, and the central importance of musical processions in their cult (Astour 1967: 176-194). Additional parallels are mentioned as follows:

••

•• The extent to which the cult of Dionysus was associated with wine is seen in that the expansion of viticulture was considered to spread his cult (Stanislawsky 1975). Similarly, YHWH displays a privileged relationship with wine (Ex 29:40; Lev 23:13; Numb 15:5-10; Isa 5:7; Jer 6:9, 12:10). •• In Greece, exudation of honey and milk from the Maenads’ staff (thyrsos) was considered a theophany of Dionysus. This exudation was even interpreted as the call of Dionysus admonishing the Maenads to put aside their domestic activities in order to worship him (Euripides, Bacchae, vv. 141-144, 160-169 and 708-711). Also

in the Bible, Canaan, YHWH’s dominion, is evoked as a land ‘flowing with milk and honey.’ This food is typically identified in Isa 7:14-15 as devoted to those initiated in the knowledge of YHWH. Choral singing was so intimately associated with the cult of Dionysus that it was called Dionysia (Jeanmaire 1991: 234). Participation in a choir was even considered to be the first stage in the revelation of the mysteries of the deity (Pailler 1995: 115). During performance, the choirmaster (choregos) was promoted to the rank of prophet of Dionysus and even regarded as his incarnation. A similar central importance of choral song characterizes the cult of YHWH. Exactly as in Greece, the choirmaster had the status of prophet of YHWH (1 Chr 25:2-5). Choral ritual performances revealed the secrets of the deity (Amzallag, 2015a), and even stimulated his theophany (1 Sam 10:5-10). Dionysus displayed a privileged relationship with serpents, and handling them apparently remained an essential component of his cult. In the Bible, YHWH’s essential relationship with serpents is also seen by their presence around his throne (Isa 6:2-3) and their function as YHWH’s guardians/emissaries (Gen 49:17; Num 21:6-9; Amos 9:3) (Amzallag 2014, 2016). The cult of Dionysus was subversive with respect to the official pantheon. It threatened the social order by inviting everyone to worship the deity, irrespective of age, gender, social status, and ethnic origin (Kraemer 1979; Dabdab-Trabulsi 1990: 86-110). Exactly the same subversive dimension is observed among the Israelites (Ps 148:11-12), whose theology is founded upon the liberation of a group of slaves from servitude, and their rejection of any divine authority other than YHWH. In Greece, the public worship of Dionysus was superimposed on a traditional esoteric cult (mysteries) anchored in Bronze Age traditions. Also, YHWH was formerly a hidden deity. This is revealed by evidence that YHWH’s genuine name was hidden before the Exodus (Ex 3:1314), and was even unknown to the patriarchs (Ex 6:3). In Isaiah 45, the superimposition of a public cult on the esoteric knowledge is revealed by the approach of YHWH as being both in time a hidden god (ēl mīsttattēr) (v. 15) and the deity who openly instructs Israel (v. 19).

These parallels, among others (see Amzallag 2011 for further details), reveal a homology between YHWH and Dionysus in their essential attributes, their requests from worshipers, their mode of action, and the similar superposition of an esoteric and a public cult. This invites us to examine to what extent the 5

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead esoteric knowledge of YHWH is anchored in Bronze Age metallurgical traditions from the Southern Levant.

tradition of Thessalonica, the large city on the coast opposite to Samothrace, tells of two Kabeiroi who killed a third and hid his head in a blood-red cloth. On Imbros, an island in the same region, the names of the Titans, the original criminals of Greek mythology, were listed in an invocation of the Kabeiroi.

YHWH and the Canaanite metalworkers Egyptian documents from the 13th Century BC refer to a nomadic population living in the metallurgical area of the Arabah designated as Shosu-YHW or Shosu-Penan (Blenkinsopp 2008: 139-140; Levy 2009; Römer 2015: 313-314). This suggests that, at the end of the Bronze Age, YHWH was identified with populations living in the metallurgical area of the Arabah, who were probably involved in copper production. These people are designated as Qenites in the Bible. Their metallurgical activity is deduced from etymological considerations (qny = to forge), from the mention of their involvement in metalworking (Gen 4:22; 1 Chr 4:13-14), and from their affiliation with the Seirites mentioned in Genesis 36 (Abramsky 1953; Weinfeld 1988; McNutt 1990: 239-243; 1999; Blenkinsopp 2008; Mondriaan 2011). Members of these tribes also dwelled among the Israelites (Jer 35:11; 2 Kgs 10:15), Canaanites (Judg 4:17-18), Midianites (Ex 2:6, 3:1; 18:1) and Amalekites (1 Sam 15:6).

Accordingly, the ‘crime’ of Cain is fully compatible with his closeness to YHWH. It should even be regarded as an indication of the esoteric nature of the metallurgical worship of YHWH performed by the Qenites, and of its ancientness, in regard to the Israelite theology. Like the cult of Dionysus in Greece, the cult of YHWH in Israel seems to be anchored in metallurgical esoteric traditions in the Southern Levant. In both cases, this metallurgical layer remains unapparent in the public worship of the deity. However, unlike the rare Greek texts mentioning Dionysus, the Bible is an outstanding source of knowledge concerning YHWH. Consequently, we should examine to what extent the esoteric dimension of the cult of YHWH is reflected, at least partly, in the Israelite theology.

On the basis of the biblical elements informing us about their way of life, social position and ritual activity, Paula McNutt (1990: 39-42; 1999) observed that the Qenites display similarities with metalworkers from traditional Africa. Moreover, exactly as in traditional Africa, Cain, the eponymous ancestor of the tribe, is evoked in Genesis 4 as the civilizing hero par excellence (Sawyer 1986; McNutt 1999; Day 2009: 342).

The metallurgical component of Yahwism in the Bible YHWH is not officially acknowledged as the smelting god in the Bible, either in the Israelite theology or in the rare allusions to his pre-Israelite cult. Nevertheless, many elements belonging to the esoteric metallurgical background are visible in biblical sources.

A Qenite privileged relationship to YHWH is suggested by the deity’s specific involvement in the birth of Cain (Gen 4:1), by the designation of Cain as being the first worshipper of YHWH (Gen 4:3), and by specific protection of Cain (and the Qenites in general) by YHWH (Gen 4:15) (Sawyer 1986). This prestigious status is acknowledged by the Israelites in 2 Kgs 10:15-16, Jer 35:14-19 and especially in Ex 18:12, where Jethro, the Qenite father-in-law of Moses, conducted the Yahwistic ceremony in the presence of Moses, Aaron and the elders of Israel (Weinfeld 1988; Blenkinsopp 2008: 134-135). Even in the mid-first millennium BCE, metalworkers were still considered heroes expected to liberate Israel in the name of YHWH (Zech 2:3-4).

The god of mining areas The vision of YHWH dwelling within mountains of copper (Zech 6:1-5) confirms his origin in metallurgical areas. Furthermore, it reveals that the Israelites did not forget the metallurgical origins of their deity, even in the early post-exilic period. This origin is also reflected in the mention of YHWH coming from the south (Hab 3:3), and more specifically from the mountains of Seir (Judg 5:4; Deut 33:2), of Paran (Deut 33:2; Hab 3:3) and of Sinai (Deut 3:3; Judg 5:5). An area of copper mining and production can be associated to each of these locations: Punon/Feinan (near Seir), Wadi Abu Kusheiba (located near the outfall of Nahal Paran), and Serabit el Khadim (in the mountains of Sinai). It could be, furthermore, that the origin of YHWH from the south (Teman) evokes the southeast mining area of the Arabah, the Timna Mountains.

Today, the prestigious status of the Qenites is generally denied because Cain is explicitly cursed for murdering Abel (Gen 4:11-13). However, we should keep in mind that exactly the same primeval crime characterizes the Kabeiroi. Rather than discrediting them, Kerenyi (1955: 45) stressed that this sin was an integrative part of their mysteries:

The importance of this origin of YHWH in metallurgical areas is reflected in the description of the country given to the Israelites as ‘... a land whose stones are iron, and out of whose hills you may dig copper’ (Deut 8:9). Considering the absence of iron and copper ores in the

The Kabeiroi themselves, the prototypes of all subsequent initiates, had been criminals. A 6

N. Amzallag: The Religious Dimension of Copper Metallurgy in the Southern Levant land of Israel, it seems that this claim primarily reflects the theological request to transform the ‘promised Land’ into a giant metallurgical area in order to justify YHWH’s presence among the Israelites (Amzallag 2013:163-164).

of this celestial fire is confirmed by the radiant material among the coals designated as hašmal (vv. 4, 27). This hapax is known in cognate languages as designating both amber and electrum, an alloy of silver and gold of a pale yellow color. In Ezekiel 1, hašmal is evoked within glowing coals/consuming fire, as something intensely radiant. This prevents its identification both as amber (which burns in fire) and as solid metal (low radiance). Rather, it seems that here hašmal designates metal in a molten state through the yellow pale color of its radiance (Driver 1951). These elements support the identification of the celestial throne of YHWH as a giant furnace (Amzallag 2013:172-175).

The serpent attack against the Israelites, related in Number 21 occurred between Mount Hor (v. 4) and Oboth (v. 10), so that its precise location (although concealed by the biblical author) was probably Punon, the area positioned between these two stations (Num 33: 41-43). An examination of the text of Numbers 21 reveals that this attack followed upon the Israelite penetration into the forbidden area of copper mining and production guarded by serpents in YHWH’s name (this was, in Antiquity, the traditional mythic function of the serpents, see Grottanelli 1987: 433-434). This means that the Israelites kept the memory not only of the link between YHWH and copper production, but also of the esoteric nature of this activity, whose access was denied even to them (Amzallag 2015b).

Ezekiel is not the only Israelite envisioning this fiery celestial reality. Before him, the elders of Israel had already contemplated the ‘celestial throne’ (Ex 24:10). However, nothing about their experience is detailed in the Bible, except the intense radiance emanating from it (a detail probably introduced to confirm that they truly contemplated the celestial domain). This silence is not fortuitous. The specification that these elders should have died because of their contemplation of the celestial domain (Ex 24:11) indicates that access to this knowledge was denied to people who were unauthorized or improperly prepared. We may assume, by extension, that divulging this celestial reality was strictly forbidden. This deduction is corroborated by an Assyrian text from the early first millennium BCE (KAR 307, 30-38). This source discloses few enigmatic indications concerning the celestial universe (once again, of fiery metallurgical nature), and it closes with the following formula: ‘Secret of the great gods: let the initiate reveal it to the initiate, but do not let the uninitiated see it’ (Livingstone 1986: 82-83; Horowitz 1998: 3-15).

Demiurgic metallurgy The term rāqîʿa designates the firmament as a hammered (√rqʿ) piece of metal (Brown 1968: 37-42; van Wolde 2009: 9). This metallic nature is supported by the mention of its brightness (Dan 12:3) and its likeness with a bronze mirror (Job 37:18). Even more, YHWH is explicitly praised for having stretched (√nṭy) the heavens (Isa 42:5; 44:24; 45:12; 51:13; Job 9:8; Ps 104:2), so that the psalmist sings that ‘the firmament claims his handiwork’ (Ps 19:2). In Egypt, too, the firmament is identified as a giant plate of copper hammered by Ptah, the smith-god (Budge 1904: 502, 511). Similarly, in Ancient Greece, the term chalkeon uranon (= copper sky) (Iliad 5:503-504, 17:424-425; Odyssey 3:1-2) suggests that the dome of the heavens was represented as a giant piece of copper.

The nature of Ezekiel’s first vision reveals that people among the Israelites were well-informed regarding the mysteries of metallurgy and their link to YHWH. Furthermore, the detailed description of the celestial throne in Ezekiel 1 indicates that a part of this esoteric metallurgical knowledge, which was jealously kept secret in other cultures, became unveiled by some biblical authors.

In the Bible, the earth is also plated by hammering (Isa 42:5; 44:24; Ps 136:6), a feature indicating that, as a whole, YHWH’s demiurgic activity was envisioned as a metallurgical process. This representation of creation is not trivial. It is justified only if at least in the past, an essential link existed between YHWH and metallurgy.

The volcanic theophany

The celestial furnace

In Ex 19:16-19, the covenant at Sinai is accompanied by intense fire, smoke and violent quakes shaking the entire mountain. This description stimulated geologists and biblical scholars to identify the Sinai theophany with a volcanic eruption (e.g. Bentor 1990: 336; Humphreys 2004:84-87; Dunn 2014: 388-397). This event is not unique in the Bible, volcanism being one of the privileged markers of divine presence and mode of action (e.g. Pss 46:7; 97:5; 104:32; 114:8; 144:5). This is why volcanism should be regarded not as a

A furnace is mentioned in close relation to YHWH in Isa 65:5 (‘these are a smoke in my nose, a fire that burn all the day’) and in Ps 18:9 (‘Smoke arose up in His nostrils, and fire out of His mouth did devour; coals flamed forth from Him’). Its detailed description is reported in Ezekiel’s opening vision. This text (Ezekiel 1) relates the existence, upon the firmament, of a celestial throne positioned in the midst of an intense bright fire (v. 4) with burning coals (vv. 13-14). The metallurgical nature 7

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead metaphor for divine powers, but rather as an essential feature characterizing YHWH’s theophany (Koenig 1966; Dunn 2014; Amzallag 2014). Within a geological context (Egypt, Canaan, Sinai peninsula) devoid of recent volcanic activity, the association between YHWH and volcanism probably held theological rather than historical significance.

melting, by which the copper of corroded metallic artifacts becomes recycled without any loss of matter (Amzallag 2015d). It is noteworthy that furnace re-melting remained an essential attribute of YHWH in the Bible. Furthermore, the visions, in late prophecies (e.g. Isa 8:19-9:6; 37:22-32; Zeph 3:8-9; Zech 7:11-8:6) of a giant destructive event of volcanic nature anticipating a renewed improved universe reveals that at the post-exilic period, furnace re-melting still constituted the fundamental element of the Israelite eschatology.

The meaning of YHWH’s essential relation with volcanism becomes clear once it is noted that in Antiquity volcanic eruptions evoked the theophany of gods patronizing metalworking (such as Vulcan and Hephaestus). This feature reflects the evidence that metallurgy was the only human activity producing flowing molten silicates (slag) very similar to lava emanating from a volcano. Here, the association of the most important event of the Israelite theological history with volcanism reveals that the Israelites acknowledged the former identity of YHWH as the smelting god, and even integrated this metallurgical background into their own religious experience.

YHWH’s emissary YHWH’s emissary, the divine being speaking and intervening in his name, is figured in Mal 3:1-3 as a metalworker. This metallurgical dimension is confirmed by his affinities with Koshar, the Ugaritic smith god, and with two Iron Age divine beings closely related to the latter: Melqart and Herakles. •• Koshar and YHWH’s emissary are mentioned in practically the same way: both are messengers of the supreme deity sent to visit the childless hero (Danel, Abraham). A similar meal is prepared by the hero for the divine visitor. After the meal, both Koshar and YHWH’s emissary foretell in a very similar fashion the birth of a new and blessed lineage, Aqhat (KTU 1.17 v 14-20) and Isaac (Genesis 18). This parallel suggests that the biblical story is inspired by the Ugaritic tale (Xella 1978) or its equivalent from south Canaan. If so, YHWH’s emissary should be considered the Israelite substitute for the Ugaritic smith god. •• The god of Tyre, Melqart, patronized metal working and trade, the main source of wealth of the Tyrians (Amzallag 2012: 132-135). He should be considered the Tyrian version of Koshar. According to Nonnus (Dionysiaca XL, 471-492), Melqart’s theophany is symbolized by a fire burning at the base of a sacred tree in the presence of a serpent. This is exactly the same theophany which is related in Ex 3:2 when the emissary of YHWH spoke for the first time to Moses. •• Herakles, a deity who displays many parallels with Melqart (Brundage 1958), ultimately saved Phrixos from being sacrificed by his father Athamas who intended to fulfill a divine instruction. Exactly the same story is related by YHWH’s emissary who intervenes to save Isaac from being sacrificed by Abraham, who in turn was fulfilling a divine request (Genesis 22). In both cases, the story is concluded by substituting, as a sacrifice, an animal incidentally present, and by the perpetual blessing of the lineage of the hero. These elements suggest that the two

The divine radiance The term kābȏd, in the Bible is generally understood as evoking YHWH’s splendor, glory and majesty. Scholars have also noticed that, beyond this meaning, kābȏdYHWH is apparently a technical term designating a heavy liquid with fiery and radiant properties, such as in Ex 24:17; Deut 4:36; 5:19-20; Isa 60:1-2 (Collins 1997: 580-584; Kutsko 2000: 80). The metallurgical dimension of this divine radiance is hinted in the text of Ex 20:2223; Isa 42:8; 48:11; Ps 106:19-20. It is confirmed in Ezek 1:27-28, where kābȏd-YHWH is closely associated to the radiations emanating from something identified as hašmal. These observations, together with others linking metallurgy and other expressions of kābȏdYHWH (e.g. solar radiance and volcanism) suggest that kābȏd-YHWH basically designates the yellow-pale radiance of metal in its molten state (Amzallag 2015c). Furnace re-melting In the Bible, qnʾ is an essential attribute of YHWH (Ex 34:14) which is even equated to his whole holiness (Jos 24:19). An examination of the mention of this divine attribute reveals that, instead of designating jealousy, it refers to a fiery mode of action closely related to volcanism (e.g. Deut 32:21-22; Zeph 1:18; 3:8; Nah 1:2-5). Though this divine intervention is strongly destructive, it is frequently mentioned as a process leading to the emergence of an improved, renewed reality (e.g. Zeph 3:8-9; Isa 37:22-32; Zech 7:11-8:6). This means that qnʾ refers to the regenerating/rejuvenating dimension of a fiery destructive mode of divine action. These considerations, together with the designation of rust as qnʾ in ancient Hebrew (see Driver 1934: 276), suggest that the divine qnʾ refers to the process of furnace re8

N. Amzallag: The Religious Dimension of Copper Metallurgy in the Southern Levant stories have the same mythological background (Wajdenbaum 2010).

with the metal produced in the furnace, either in color, density or mechanical properties. This means that the production of copper from ore, in the Southern Levant, was probably approached from its origin as a demiurgic activity. This may easily transform the god patronizing smelting into the master of demiurgic powers, that is, the supreme deity. The reiteration of this demiurgic operation, at each smelting process, may even justify why the ritualized dimension of smelting did not disappear after the collapse of the Ghassulian culture. Furnace metallurgy, once developed in the Southern Levant, propagated rapidly in the Ancient Near East (Amzallag 2009). Accordingly, it is likely that the ritual dimension of furnace metallurgy diffused together with this practical knowledge, to become also a fundament for other Ancient Near Eastern religions.

The affinities between Koshar, Melqart and Herakles, and their common interrelations with YHWH’s emissary suggests that the latter is the Israelite version of the smith god acknowledged in neighboring cultures, whose origin is anchored in Bronze Age traditions (Amzallag 2012). These tales also reveals that in the Bronze Age the smith god represented the emissary, in the official religion, of the smelting god of esoteric nature. All these considerations, when gathered, leave little doubt concerning the essential dimension of metallurgy in ancient Yahwism. Furthermore, the absence of iron in YHWH’s sanctuary (where gold, silver and copper are abundant), together with the attributes and modes of action inspired by copper metallurgy, strongly suggest that the metallurgical background of the god of Israel is anchored in Bronze Age traditions. This enables us to identify YHWH, in his pre-Israelite worship, as the South Levant deity patronizing copper smelting, whose origin is rooted in Bronze Age traditions, and probably even before. Furthermore, the central importance devoted to the metallurgical theophany, mode of action and attributes of YHWH reveals that the Israelites did not consider these features as vestigial, but rather as a central, though hidden, background founding their theology.

Though this esoteric knowledge remains enigmatic today, the present study suggests that it is possible to characterize its nature and general significance through a cross-cultural investigation that combines the following approaches: (i) examination of the Chalcolithic and Bronze Age rituals and the prestige metallic artifacts from the Southern Levant; (ii) analysis of the literary sources concerning rituals and mythologies relative to the smith gods from Bronze Age societies; (iii) investigations of the religious mysteries practiced during the Iron Age in the Ancient Near East; (iv) identification of the official cults of the smelting god during the Iron Age, both in Canaan and in neighbor cultures; (v) integration of testimonies regarding the esoteric layer of religions from traditional Africa.

Conclusion

The present paper especially stresses the importance of re-examining the Bronze Age literary sources relative to the smith god and his hidden patron-deity. It also indicates that both the transformation of esoteric Bronze Age traditions into a public cult and the spread of some mystery cults had an outstanding influence on the development and evolution of Iron Age societies, and even, on the emergence of monotheism. It also reveals that the Bible may be an exceptional source of information concerning the nature of the esoteric metallurgical knowledge.

The cultural dimension of metallurgy, which was probably of outstanding importance in the Southern Levant from the beginning, apparently did not disappear following the increase in production and trade of metals and utilitarian artifacts. Rather, it became the esoteric foundation on which religions emerged and further developed. The present study reveals similarities between the esoteric dimension of metallurgy in the Bronze Age, its further developments in the Iron Age and even its affinities with metallurgical traditions from traditional Africa. These observations confirm the antiquity of the metallurgical esoteric traditions, and their probable diffusion from a specific homeland.

Historians generally express reservations in approaching the Bible as a reliable source of information. The reason for this is the apologetic dimension that characterizes many of the biblical sources, which were conceived to defend one specific theology and cult over others. Some considerations minimize these reservations in the present case. The metallurgical background of ancient Yahwism, though omnipresent in the Bible, is never emphasized or even openly mentioned. This means that this reality does not belong to the apologetic layer of the biblical discourse. This is not difficult to understand why. The Israelite theology is elaborated on the basis of the transfer to a new ‘people of YHWH’ (= the Israelites)

The Southern Levant is the only area where metallurgy emerged independently from the exploitation of sources of native copper. It is also the only region where metallurgy was apparently practiced in a furnace rather than a crucible from its very beginning (Amzallag 2009). Accordingly, metallurgy, in the Southern Levant, was not simply an extension of the work and exploitation of an already existing material. It was a process that enabled the production of a new material. Furthermore, copper ore from the Arabah does not show affinities 9

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead of the religious authority previously attached to the Canaanite metalworkers (= the Qenites / Seirites), the traditional guardians of the esoteric knowledge. This transfer of authority is evoked in Genesis 27 in an unfavorable way (through the deceiving stratagem used by Jacob/Israel to obtain the primogeniture rights initially granted to Esau/Edom). This probably betrays deep Israelite self-critical problems of legitimacy. The demonization of Edom in late biblical sources reveals that this struggle for Yahwistic authority was still acute at the end of the First Temple period (Amzallag 2015e: 53-65). In such a context, the mention of the metallurgical background of YHWH may have hardly been introduced by the Israelites for apologetic purposes, because it strengthened the legacy of the Edomites/Qenites (= the Canaanite metalworkers) religious authority at the expense of the Israelite one. This is why the metallurgical dimension expressed in the Bible should be globally approached not only as a key element for understanding the emergence of monotheism, but also as a reliable source informing us about the esoteric metallurgical beliefs founding the South Levant religions in the Bronze Age.

Amzallag, N. 2015b. The Origin and Evolution of the Saraph Symbol. Antiguo Oriente 13: 99-126. Amzallag, N. 2015c. The Material Nature of the Divine Radiance and its Theological Implications. Scandinavian Journal for the Old Testament 29: 80-96. Amzallag, N. 2015d. Furnace Re-melting as Expression of YHWH’s Holiness: Evidence from the Meaning of qanna (‫ )אנק‬in Divine Context. Journal of Biblical Literature 134: 233-252. Amzallag, N. 2015e. Esau in Jerusalem – The Rise of a Seirite Religious Elite in Jerusalem at the Persian Period. Pendé (France), Gabalda. Amzallag, N. 2016. The Serpent as a Symbol of Primeval Yahwism. Semitica 58: 208-239. Amzallag, N. Technopoiesis (in press) – The Forgotten Dimension of Techniques in Ancient Societies. Time and Mind. Astour, M.C. 1967. Hellenosemitica. An Ethnic and Cultural Study in West Semitic Impact on Mycenaean Greece. Leiden, Brill. Aufrère, S. 1991. L’univers minéral dans la pensée Égyptienne. Le Caire, Institut Français d’Archéologie Orientale. Avilova, L. 2008. Regional Models of Metal Production in Western Asia in the Chalcolithic, Early and Middle Bronze Ages. Trabajos de Prehistoria 65: 73-91. Avilova, L. 2012. On the Characteristic of Temple Complexes in the Near East in the 4th – 3th Millennia BC. Asian Culture and History 4: 3-15. Ben-Yosef, E., Shaar, R., Tauxe, L., Ron, H. 2012. A New Chronological Framework for Iron Age Copper Production at Timna (Israel). Bulletin of the American Schools for Oriental Research 367: 31-72. Bentor, Y. K. 1990. Geological events in the Bible. Terra Nova 1: 326-338. Blakely, S. 2006. Myth, Ritual and Metallurgy in Ancient Greece and Recent Africa. Cambridge, Cambridge University Press. Blenkinsopp, J. 2008. The Kenite-Midianite Hypothesis Revisited and the Origin of Judah. Journal for the Study of the Old Testament 33: 131-153. Bloedow, E. F. 1991. Evidence for an Early Date for the Cult of Cretan Zeus. Kernos 4: 139-177. Boyer, P. 1983. Le statut des forgerons et ses justifications symboliques: une hypothèse cognitive. Africa 53: 4463. Bremmer, J. N. 2014. Initiation into the Mysteries of the Ancient World. Berlin, De Gruyter. Brodie, N. 1997. New Perspectives on the Bell-Beaker Culture. Oxford Journal of Archaeology 16: 297-314. Brown, J. P. 1968. Cosmological Myth and the Tuna of Gibraltar. Transactions and Proceedings of the American Philological Association 99: 37-62. Brück, J. 2011. Fire, Earth, Water. An Elemental Cosmography of the European Bronze Age. In T. Insoll (ed.), Oxford Handbook of the Archaeology of Ritual and Religion: 387-404. Oxford, Oxford University Press.

References Abramsky, S. 1953. The Qenites. Eretz-Israel 3: 116-124 (heb). Adams, R. B. 2002. From Farms to Factories: The Development of Copper Production at Faynan, Southern Jordan, During the Early Bronze Age. In B. S. Ottaway and E. C. Wager (eds), Metals and Society: 21-32. Oxford, Archaeopress. Adams, W. Y. 1984. The First Colonial Empire: Egypt in Nubia, 3200-1200 B.C. Comparative Studies in Society and History 26: 36-71. Algaze, G. 1989. The Uruk Expansion: Cross-Cultural Exchange in Early Mesopotamian Civilization. Current Anthropology 30: 571-591. Amzallag, N. 2009. From metallurgy to Bronze Age civilizations: The synthetic theory. American Journal of Archaeology 113: 497-519. Amzallag, N. 2011. Was YHWH worshipped in the Aegean? Journal for the Study of the Old Testament 34: 387-415 Amzallag, N. 2012. The identity of the emissary of YHWH. Scandinavian Journal for the Old Testament 26: 123-144. Amzallag, N. 2013. Copper Metallurgy: A Hidden Fundament of the Theology of Ancient Israel? Scandinavian Journal for the Old Testament 27: 155180. Amzallag, N. 2014. Some Implication of the Volcanic Theophany of YHWH on his Primeval Identity. Antiguo Oriente 12: 11-38. Amzallag, N. 2015a. Psalm 67 and the Cosmopolite Musical Worship of YHWH. Bulletin for Biblical Research 25: 31-48. 10

N. Amzallag: The Religious Dimension of Copper Metallurgy in the Southern Levant Brundage, B. C. 1958. Heracles the Levantine: a comprehensive view. Journal of Near Eastern Studies 17: 225-236. Budd, P. and Taylor, T. 1995. The Faerie Smith meets the Bronze Industry: Magic Versus Science in the Interpretation of Prehistoric Metal-Making. World Archaeology 27: 133-143. Budge, E. A. W. 1904. The Gods of the Egyptians (vol. 1). London, Methuen. Butterlin, P. 2003. Les temps proto-urbains de Mésopotamie. Contact et acculturation à l’époque d’Uruk au Moyen Orient. Paris, CNRS Éditions. Capdeville, G. 1995. Volcanus – Recherches comparatistes sur les origines du culte de Vulcain. Rome, École Française de Rome. Cauvin, J. 2002. The Symbolic Foundations of the Neolithic Revolution in the Near East. In I. Kuijt (ed.), Life in Neolithic Farming Communities – Social Organization, Identity and Differentiation: 235-252. New York, Kluwer. Chernykh, E. N. 1992. Ancient Metallurgy in the USSR. The Early Metal Age. Cambridge, Cambridge University Press. Collins, C. J. 1997. Kabod. In W. A. van Gemeren (ed.), The New International Dictionary of Old Testament Theology and Exegesis: 577-587. Grand Rapids: Zondervan. Dabdab-Trabulsi, J. A. 1990. Dionysisme- pouvoir et société en Grèce jusqu’à la fin de l’époque classique. Paris, Les Belles Lettres. Davidson, H. R. E. 1969. The Chariot and the Sun. Folklore 80: 174-180. Day, J. 2009. Cain and the Kenites. In G. Galil, M. Geller and A. Millard (eds), Homeland and Exile – Biblical and Ancient Near Eastern Studies in Honour of Bustenay Oded: 335-346. Leiden, Brill. de Maret, P. 1980. Ceux qui jouent avec le feu: la place du forgeron en Afrique central. Africa 50: 263279. de Miroschedji, P. 1998. Les égyptiens au Sinaï du nord et en Palestine au Bronze ancien. In D. Valbelle and C. Bonnet (eds), Le Sinaï durant l’Antiquité et le Moyen Âge: 20-32. Paris, Errance. Dieterle, R. L. 1987. The Metallurgical Code of the Volundarkvida and its Theoretical Import. History of Religions 27: 1-31. Driver, G. R. 1934. Hebrew Notes on the ‘Wisdom of Jesus Ben Sirach’. Journal of Biblical Literature 53: 273-290. Driver, G. R. 1951. Ezekiel’s Inaugural Vision. Vetus Testamentum 1: 60-62. Dunn, J. E. 2014. A God of Volcanoes: Did Yahwism Take Root in Volcanic Ashes? Journal for the Study of the Old Testament 38: 387-424. Eliade, M. 1968. The forge and the crucible: a postscript. History of Religions 8:74-88. Eliade, M. 1977. Forgerons et alchimistes. Paris: Flammarion. Faure, P. 1997. La Crète au temps de Minos. Paris: Hachette.

Finnestad, R. B. 1976. Ptah, Creator of the Gods: Reconsideration of the Ptah Section of the Denkmal. Numen 23: 81-113. Freyburger-Galland, M. L. 2006, Sectes religieuses en Grèce. In G. Freyburger, M. L. Freyburger-Galland and J. C. Tautil (eds), Sectes religieuses en Grèce et à Rome dans l’Antiquité païenne: 19-161. Paris, Les Belles Lettres. Galter, H. D. 2015. The Mesopotamian God Enki/Ea. Religion Compass 9: 66-76. Gilead, I. 2002. Religio-Magic Behavior in the Chalcolithic Period of Palestine. In E. Oren and S. Ahituv (eds), Aharon Kempinski Memorial Volume – Studies in Archaeology and Related Disciplines: 103-128. Beer Sheva: Ben Gurion University Press. Goldhahn, J. and Oestigaard, T. 2007. Smith and Death – Cremations in Furnace in Bronze and Iron Age Scandinavia. In K. Chilidis (ed.), Facets in Archaeology. Essays in honour of Lotte Hedeager on her 60th birthday: 215-241. Oslo, Oslo Academic Press. Gophna, R. and Milevski, I. 2003. Feinan and the Mediterranean during the Early Bronze Age. Tel Aviv 30: 221-231. Gošić, M. 2013. Metallurgy, Magic and Social Identities int he Ghassulian Culture of the Southern Levant (ca. 4500-4000 BC). Unpublished PhD Thesis, Ben-Gurion University of the Negev, Beer Sheba, Israel. Gošić, M. 2015. Skeumorphism, Boundary Objects and Socialization of the Chalcolithic Metallurgy in the Southern Levant. Issues in Ethnology and Anthropology 10: 717-740. Gošić, M. and Gilead, I. 2015a. Casting the sacred: Chalcolithic metallurgy and ritual in southern Levant. In N. Laneri (ed.), Defining the Sacred – Approaches to the Archaeology of Religion in the Near East: 161-175. Oxford, Oxbow. Gošić, M. and Gilead, I. 2015b. Unveiling Hidden Rituals: Ghassulian Metallurgy of the Southern Levant in light of the Ethnographical Record. In K. RosinskaBalik, A. Ochal-Czarnowicz, M. Czarnowicz and J. Debowska-Ludwin (eds), Copper and Trade in the South-Eastern Mediterranean. Trade Routes of the Near East in Antiquity: 25-37. Oxford, Archaeopress. Griaule, M. 1948. Dieu d’eau – Entretien avec Otogommêli. Paris, Fayard. Grottanelli, C. 1987. Dragons. In M. Eliade (ed.), The Encyclopedia of Religion (vol. 4): 431-436. New York, McMillan. Herbert, E. W. 1973. Aspects of the Use of Copper in PreColonial West Africa. Journal of African History 14: 179-194. Horowitz, W. 1998. Mesopotamian Cosmic Geography. Winona Lake, Eisenbrauns. Humphreys, C. 2003. The Miracles of Exodus. London, Continuum. Ionescu, D. and Dumitrache, C. 2012. The Sun Worship with the Romanians. Romanian Astronomical Journal 22: 155-166. 11

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Jayne, W. A. 1962. The Healing Gods of Ancient Civilizations. New Haven, Yale University Press. Jeanmaire, H. 1991. Dionysos – Histoire du culte de Bacchus. Paris, Payot. Kempinski, A. 1989. Urbanization and Metallurgy in Southern Canaan. In P. de Miroschedji (ed.), L’urbanisation de la Palestine à l’âge du Bronze Ancien: 163-168. Oxford, Archaeopress. Kerenyi, C. 1955. The Mysteries of the Kabeiroi. In The Mysteries – Papers from the Eranos Yearbooks (translated by R. Manheim): 32-59. Princeton, Princeton University Press. Kerenyi, C. K. 1976. Dionysos – Archetypal Image of Indestructible Life. Princeton, Princeton University Press. Koenig, J. 1966. Aux Origines des Théophanies Iahvistes. Revue d’histoire des religions 169: 1-36. Kohl, P. L. 2007. The Making of the Bronze Age Eurasia. Cambridge, Cambridge University Press. Kraemer, R. S. 1979. Ecstasy and Possession: the Attraction of Women to the Cult of Dionysus. Harvard Theological Review 72: 55-80. Kristiansen, K. and Larsson, T. 2005. The Rise of the Bronze Age Society. Travels, Transmissions and Transformations. Cambridge, Cambridge University Press. Kutsko, J. F. 2000. Between Heaven and Earth. Divine Presence and Absence in the Book of Ezekiel. Winona Lake, Eisenbrauns. Lambert, J. N. 1980. La divinité du Mont Bégo (Alpes Maritimes). Première approche d’après l’histoire comparée des religions. Revue de l’histoire des religions 197: 355-407. Lenzi, A. 2008. Secrecy and the Gods. Secret Knowledge in Ancient Mesopotamia and Biblical Israel. Helsinki, NeoAssyrian Text Corpus Project. Levy, T. E. 2009. Pastoral Nomads and Iron Age Metal Production in Ancient Edom. In J. Szuchman (ed.), Nomads, Tribes and the State in the Ancient Near East – Cross-Disciplinary Perspectives: 147-177. Chicago, The Oriental Institute of the University of Chicago. Levy, T. E., Ben-Yosef, E. and Najjar, M. 2012. New Perspectives on Iron Age Copper Production and Society in the Faynan Region, Jordan. In V. Kissianidou and G. Papasavvas (eds), Eastern Mediterranean Metallurgy and Metalwork in the Second Millennium BC: 197-214. Oxford, Oxbow. Livingstone, A. Mystical and Mythological Explanatory Works of Assyrian and Babylonian Scholars. Oxford, Oxford University Press. Lucas, J. O. 1948. The Religion of the Yoruba in Relation to the Religion of Ancient Egypt. Lagos, C.M.S. Martin, M. 2005. Magie et magiciens dans le monde grécoromain. Paris, Errance. McNutt, P. M. 1990. The Forging of Israel. Iron Technology, Symbolism and Tradition in Ancient Society. Sheffield, Almond Press. McNutt, P. M. 1999. In the Shadow of Cain. Semeia 87: 45-64.

Meyerowitz, E. L. R. 1960. The Divine Kingship in Ghana. London, Faber and Faber. Mondriaan, M. E. 2011. Who were the Kenites? Old Testament Essays 24: 414-430. Pailler, J. M. 1995. Bacchus. Figures et pouvoirs. Paris, Les Belles Lettres. Pfoh, E. 2014. Metalworkers in the Old Testament: An Anthropological View. In D. J. Chalcraft; F. Uhlenbruch and R. S. Watson (eds), Methods, Theories, Imagination – Social Scientific Approaches in Biblical Studies: 201-217. Sheffield: Sheffield Phoenix Press. Pleiner, R. and Bjorkman, J. K. 1974. The Assyrian Iron Age: The History of Iron in the Assyrian Civilization. Proceedings of the American Philosophical Society 118: 283-313. Rowlands, M. J. 1971. The Archaeological Interpretation of Prehistoric Metalworking. World Archaeology 3: 210-224. Reid, A. and MacLean, R. 1995. Symbolism and the Social Contexts of Iron Production in Karagwe. World Archaeology 27: 144-161. Römer, T. 2015. The Revelation of the Divine Name to Moses and the Construction of a Memory about the Origins of the Encounter between YHWH and Israel. In T. E. Levy, T. Schneider and W. H. C. Propp (eds), Israel’s Exodus in Transdisciplinary Perspective. Text, Archaeology, Culture and Geoscience: 305-315. Berlin: Springer. Rosen, S. A. 2002. Invention as the Mother of Necessity: An Archaeological Examination of the Origins and Development of Pottery and Metallurgy in the Levant. In R. Harrison, M. Gillespie and M. PeuramkiBrown (eds), Eureka: The Archaeology of Innovation and Science: 11-21. Calgary, University of Calgary Press. Sawyer, J. F. A. 1986. Cain and Hephaestus. Possible Relics of Metalworking Traditions in Genesis 4. AbrNahrain 24: 155-166. Scarano, T. and Maggiulli, G. 2014. The Golden Sun Discs from Roca Vecchia, Lecce, Italy: Archaeological and Cultural Context. Tagungen des Landesmuseums fur Vorgeschichte Halle 11: 1-21. Schachter, A. 1986. Cults of Boiotia. London, University of London. Schwemer, D. 2015. The Ancient Near East. In D. J. Collins (ed.), The Cambridge History of Magic and Witchcraft in the West: 17-51. Cambridge, Cambridge University Press. Shugar, A. N. 2003. Reconstructing the Chalcolithic Metallurgical Process at Abu Matar, Israel. In Proceedings of the Archaeometallurgy in Europe Conference (vol. 1): 449-458. Milano, Associazione italiana di metallurgia. Stanislawsky, D. 1975. Dionysus Westward: Early Religion and the Economic Geography of Wine. Geographical Review 65: 427-444. Tegnaeus, H. 1950. Le héros civilisateur. Contribution à l’étude ethnologique de la religion et de la sociologie 12

N. Amzallag: The Religious Dimension of Copper Metallurgy in the Southern Levant africaines. Studia Ethnographica Upsaliensia. Uppsala, Almqvist and Wiksells. van Beek, W. E. 1991. Iron, Brass and Burial: The Kapsiki Blacksmith and his Many Crafts. In Y. Monino (ed.), Forge et Forgerons: 281-310. Paris, Orstom. van Buren, D. 1934. The God Ningizzida. Iraq 1: 60-89. van Wolde, E. 2009. Why the Verb bārā’ Does not Mean Create in Genesis 1.1-2.4. Journal for the Study of the Old Testament 34: 3-23.

Wajdenbaum, P. 2010. Is the Bible a Platonic book? Scandinavian Journal for the Old Testament 24: 129-142. Weinfeld, M. 1988. The Tradition of Moses and Jethro on the Mountain of God. Tarbiz 56: 449-460 (Heb). Xella, P. 1978. L’épisode de Dnil et Kothar et Gen 18:1-16. Vetus Testamentum 28: 483-488. Zahan, D. 1979. The Religion, Spirituality and Thought of Traditional Africa (transl. K. E. Martin and L. E. Martin). Chicago, University of Chicago Press.

13

Neolithic Cult Sites in the Southern Negev, Israel Uzi Avner, Moti Shem-Tov, Lior Enmar, Gideon Ragolski, Rachamim Shem-Tov, and Omry Barzilai Abstract An archaeological survey in the southern Negev, Israel, uncovered hundreds of Neolithic cult sites built on the mountains, with specific characteristics: ‘regular’ and perforated standing stones, anthropomorphic stone images and other intriguing stone features. Although no such sites reported in other surveys in the Negev, Sinai and southern Jordan, these cult sites are not unique to the southern Negev. The sites and their content add new dimensions to our understanding of the Early and Late Neolithic periods in the Levantine deserts. The paper describes the sites and offers some interpretation of their features.

Introduction

The first mountain cult site was discovered in 1979, during an archaeological survey east of ‘Uvda Valley (Avner 1979, 1982a, 2002:69). It contained a barely visible elongated stone cell, 3.8 x 1.6 m, with an in situ triad of perforated small standing stones, up to 32 cm high (Fig. 1). Later on, 27 similar installations were discovered in 11 sites (surveyed by L. Enmar and U. Avner).

Until quite recently, little was known of both the Early and Late Neolithic periods in the southern Negev. The earliest Neolithic occurrence in this region dates to the Pre-Pottery Neolithic B (PPNB, 8th-7th millennia cal. BC), of which four stone-built habitations and a few smaller sites found in the vicinity of the ‘Uvda Valley, southern Negev (Avner 1979, 1982a, 1982b, 1998, 2002 Ch.2). Two of the habitations were excavated: Naḥal ‘Issaron (Goring Morris and Gopher 1983, 1987; Carmi et al. 1994) and Naḥal Re’uel (Ronen et al. 2 001). PPNB flint items were also collected from several ancient roads in the region, indicating the existence of an intensive road network in the desert during this period (Avner 2002, Ch. 6). The Late Neolithic period was known from even fewer sites: an open-air sanctuary in the ‘Uvda Valley (Avner 1979, 2002, Ch. 5; Yogev 1983), a cemetery in Eilat (Avner 1990b, 2002 App. 1; Eshed and Avner 2018), some occupation levels and 14C dates beneath three habitation site of later periods in the ‘Uvda Valley (Avner 2006, Table 1:21, 25) and a possible activity related to early copper production at Yotvata (Rothenberg et al. 2004).

Following the discovery of several more such sites on the mountains surrounding Naḥal Roded, near Eilat, a

In 1996, the Neolithic settlement scenario has somewhat changed with the discovery of a large habitation site near Be‘er ‘Ada (Naḥal Paran), with an intensive bead manufacturing industry, probably dated to both PPNB and Late Neolithic (Avner 1997). Since 2009, five small PPNB habitations were discovered (by M. Shem-Tov, L. Schwimmer and U. Avner), one near Naḥal Paran and four in the Eilat region. The major change, however, was the discovery of dozens of cult installations mostly on mountains, dated to both the PPNB and Late Neolithic. The sites are characterised by low stone cells of several forms, with unique stone objects and features (see below). In this article, we focus on the distinctive phenomenon of these mountain cult installations.

Figure 1. Eastern ‘Uvda Valley, remains of an elongated cell with an in-situ triad of perforated maṣṣeboth, up to 32 cm high.

14

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel systematic survey in the area was initiated in 2004, on behalf of The Hebrew University, Jerusalem, sponsored by the CARE Foundation. The first survey stage focused on an area of 12 sq km, where 103 cult sites were documented (Fig. 2, Avner et al. 2014). Later, many more sites were recorded in other parts of the greater Eilat Region (mainly by M. Shemtov) and a renewed survey east of ‘Uvda Valley discovered 71 of these sites.

Currently, 324 such sites are recorded in the greater Eilat Region (including 21 on the ridges between the Qetura Junction and Kibbutz Yahel). Twenty eight cult sites were found by G. Ragolsky in the central ‘Araba area, 11 sites were encountered on the southern ridges of the Negev-Highlands (by M. Shemtov, L. Enmar and U. Avner) and two installations were also observed (by U. Avner) on a mountain north of Kh. Naḥas in

Figure 2. Map of Naḥal Roded area with the recorded Rodedian sites, in relation to lithology.

15

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Jordan (not counted with the Negev sites). Presently (December 2016), the total number of these cult sites in the southern Negev is 363 (Fig. 3). Since the densest cluster of sites was on the mountains surrounding Naḥal Roded, near Eilat (44 in an area of only 0.8 sq km, Figure. 2), we termed them ‘Rodedian’. The surveys of the Eilat

Mountains and ‘Uvda Valley also contributed nearly 200 other cult installations with ‘regular’ maṣṣeboth, mostly dated to the 6th-3rd millennia BC, as well as of later periods (not addressed in this paper). Since large areas in the Negev are still un-surveyed, the true numbers of these Rodedian sites are expected to increase.

Figure 3. Map of currently recorded Rodedian sites in the Negev, with regions’ names mentioned in the text.

16

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel One preliminary conclusion as to the presently known distribution of Rodedian sites is that they are not ‘endemic’ to the Eilat Region. They are scattered over large mountainous areas of the Negev desert but their limits are still unknown. Interestingly enough, no such sites have been reported to date from other surveys carried out in the Negev,1 Sinai or the deserts of Jordan.

sites, was found on the igneous mountains of the Eilat area, six in the limited sandstone area and 231 sites were recorded in the much larger limestone terrain. Narrow trails leading to the sites were often observed. The locations of these sites contrast to those of desert habitations, which are almost always built on foothills, on wadi terraces. Only four small habitations were found in the mountains, within clusters of Rodedian sites. Two of these were found on the igneous mountains above Naḥal Roded and two on a limestone plateau above Naḥal ‘Eteq and one on Har Orah. The unusual location of these habitations, their proximity to the cult sites and PPNB blades collected in all, clearly associate them to the cult complex.

Methodology The survey of these sites faces several obstacles: They cannot be detected by aerial photos, while the rugged topography makes the survey strenuous and time consuming. In addition, all gear has to be carried on back. The sites are barely visible, although on the igneous mountains, they are sometimes detected from some distance due to the use of imported light coloured limestones. Considering the difficulties, it is certain that not all existing sites were detected in the surveyed areas.

Most Rodedian sites were found disturbed to various degrees or even vandalised in antiquity. Standing stones were usually tumbled or broken, and so were the shaped stone objects, sometimes discarded up to 30 m off-site. The original location and position of stone items could be learned in only a limited number of sites.

Documentation of the Rodedian sites included mapping by GPS and GIS, measurement of the stones installations and features, a written description, collection of surface artefacts and photography. Stone-by-stone plans were made for only 40 sites, but vertical photography was sufficient in others. An Excel chart was prepared and updated continuously (by R. Shemtov) displaying the sites’ spectrum of characteristics and allowing quantitative analyses of their properties. Since stone objects were rarely found in situ, they were documented first as found. Some were set or laid on the ground for additional photography, then placed back in their original position. Some sites had suffered theft of stone objects. Therefore, special items were removed from easily accessible sites. To date, none of the Rodedian sites has been excavated. Two elongated cells were cleaned of 1-3 cm of dust with arkose or gravel, exposing a pavement; four probes, up to 40 x 40 x 30 cm, were excavated next to specific features and one shallow ash spot was sampled for 14C dating (see all below). Flint items from selected sites were analysed by O. Barzilai (in Avner et al. 2014).

The sites’ pattern The installations are very low, built in three different ways: small flagstones set vertically into the ground, horizontally laid flagstones, or one course and one row of small field-stones. In the latter, the stones are usually partially imbedded in the ground, an indication of their antiquity. Four main forms of installations recur. One is circular 1.5-2 m across (currently, 178 circles in 131 sites), second is oval with similar dimensions (105 ovals in 98 sites), third is an elongated cell, usually ca. 4 m long and 0.7-1.5 m wide (133 in 112 sites), fourth is a stone alignment (11 in 11 sites); 139 installations were termed ‘others’. Two installations were often built as pairs of repeated pattern, in which the elongated cell points to the circle (128 pairs in 104 sites, Figs. 4, 5). Most sites contain one or two installations or one regular pair, but some contain more, up to 11 cells (in total 566 cells in 363 sites). In 21 sites, the installations are built on top of a low artificial terrace, supported by a semi-circular retaining wall. In three sites, the installations are larger than usual. One is Site 239 in the ‘Uvda Valley, where the elongated cell is 9.7 x 4 m, and shaped as a shoe sole (Fig. 6). The elongated cells are paved sometimes by small limestone flagstones (Fig. 7), by large limestone slabs or by other flat stones (Fig. 4). The circular installation in the pairs is commonly built of regular field stones, on a somewhat lower ground than that of the elongated cell. No dominant orientation was found in the installations, they usually relate to the local topography.

The Sites’ Location and Condition The vast majority of the sites are situated at various elevations on mountain ridges, on topographic ‘shoulders’ below summits, or on small relatively flat areas on the slopes. On limestone terrain, many are built on plateaus and next to cliff edges, while only a few were found at the bottom of the hill. The highest density, 126 One site was reported by Anati from the survey of Har Karkom, but described as an ‘anthropomorphic geoglyph’. In several publications the site is indicated under different numbers and in different locations (HK 96b, HK 190c and HK 130c: Anati 1993:31, 2001, Fig 61, 2005:41, 2009:168). The true location of the site is G.R. 175768-468890. Two additional Rodedian sites on Har Karkom were found by M. Shemtov and U. Avner (G.R. 176450467465; 175760467376). 1 

Stone objects Several types of engraved or naturally shaped stones are found in and around the installations. Some sites 17

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead 1. Regular maṣṣeboth (standing stones) are stone slabs originally set vertically into the ground. Most were set individually but they also occur in groups of repeating numbers. They are termed here ‘regular’ for two reasons. One is that they are very common in the desert, whether in independent shrines or incorporated in open-air sanctuaries and in tumuli tombs (Avner 1984, 1993, 2001, 2002 Chs. 4, 5, App. 1, 2018), second- they are distinguished from perforated maṣṣeboth (see below) which are unique to the Rodedian sites. Maṣṣeboth in Rodedian sites are quite small, up to 60 cm high, some are found in situ, vertically set or tilted, many others are fallen or even broken (Fig. 8). To date, 255 maṣṣeboth are recorded: 151 individual ones, 25 pairs, eight triads (Fig. 9), two groups of five and three groups of seven. One group of seven small stones was found arranged in an arch, still upright as set some 8000 years ago, with a copper nodule laid at the foot of the central stone (Fig. 10). Another possible group of seven small elongated maṣṣeboth, 12-21 cm long, was hidden in a closed natural cavity next to another site above Naḥal Roded. 2. Small Perforated Maṣṣeboth- mostly of limestone slabs, usually from 10 x 8 to 25 x 15 cm, with a rounded, biconical perforation below the top. The perforation was made first by chiselling, and then carefully smoothed; otherwise, the stones remained unshaped. Altogether, 182 perforated maṣṣeboth were recorded in 79 sites. When found in situ, they were set both inside and outside the elongated cell; only one site contained an in-situ triad (Fig. 1) while others were individual. Most perforated maṣṣeboth are found fallen, sometimes broken or even discarded off-site in antiquity (see below).

Figure 4. Har Assa, a typical pair of installations, the elongated cell points to the circle (pavement was covered by 1-3 cm of dust and arkose).

contained many items of several types, in others- only a few are found, while in some, no stone objects or features were found. Almost all engraved stone were made of limestone; a few are of sandstone, granite, diorite or porphyry. Altogether, these objects form an intriguing repertoire:

3. Natural Holed Stones- almost all are limestone, with natural holes created by chemical or biological weathering (Fig. 11). On limestone terrain, such stones are quite common and easily ignored. However, since many of them occur in sites built on igneous ridges and

Figure 5. Har Argaman, plan of a typical pair of installations.

18

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel

Figure 6. ‘Uvda Valley, a large elongated cell with maṣṣebah, pointing to a barely preserved circle.

Figure 7. Naḥal Roded, an elongated cell paved with small limestone slabs.

some are found vertically set into the ground, it is clear that they were deliberately selected and brought to the sites from some distance, due to an unknown symbolic value related to them. A total of 825 of these stones were found, in 134 sites. 4. Anthropomorphic Images are elongated stones, mostly limestone, 21-46 cm long with a human appearance, either natural or schematically shaped with various degrees of workmanship. Altogether, 329 stone images were found in 149 sites, with the largest cluster of 17 stone images in one site. One hundred ninety-nine were natural, unshaped (Fig. 12:1) selected due to their human-like silhouette, 122 had a delineated, pecked neck (Fig. 12:2) or lightly flaked ‘shoulders’, while eight were carefully shaped and entirely pecked (Fig. 12:3). Four stone images are broad, having a female look, three of them are natural cobbles, and one was shaped by flaking the sides (Figs. 13:1, 2). In one site, three small figurine-like images were found together, 10-12 cm high; another is a finely shaped basalt figurine, 9.5

Figure 8. Naḥal Roded, a pair of limestone maṣṣeboth, found tumbled and broken.

19

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 9. Naḥal Roded, a triad of maṣṣeboth in a circular cell, as found (left) and after re-setting the central one.

Figure 10. Naḥal Shehoret, a group of seven small maṣṣeboth as found, up to 34 cm high, with a small copper nodule at the foot of the central maṣṣebah.

Figure 11. Naḥal Roded, perforated and naturally holed limestones, found scattered in and around the installations (recorded first as found).

20

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel

Figure 12. Naḥal Roded, three examples of anthropomorphic stones: 1. Unshaped 2. With a hammered neck 3. Entirely hammered.

Figure 13. Unique stone images: 1. Eastern ‘Uvda Valley, a natural broad ‘female’-look cobble. 2. Naḥal Paran, a broad ‘female’-look cobble partially shaped. 3. Eastern ‘Uvda Valley, basalt figurine entirely hammered.

cm high from a site east of ‘Uvda Valley (Fig. 13:3). Most images were found tumbled, some were deliberately broken (in antiquity), and some were found discarded up to 30 m off-site. In only 16 sites, stone images were found in situ; in 12 of them pairs of stone images were set on the narrow end of the elongated cell (Fig. 14). In one site, a stone image was attached to a row of four, small regular maṣṣeboth, also at the end of the elongated cell. In two sites, a broad stone image was set in the centre of the circular cell while in another circle the stone image was buried with the head up. Three images were set inside ‘vase-shaped’ stone installations, two with the head up, one with the head down (see below).

5. Stones With Elongated Perforation are all of limestone, ca. 25 x 15 x 4 cm, with an elongated or oval perforation, made first by chiselling and then carefully smoothed (Fig. 15). Thirty such stones were found, in 17 sites, most were around Naḥal Roded. Eight were complete while 22 were fragments. Only one was found in-situ, vertically set in the centre of an elongated cell, others were found discarded off-site. One half stone bears an engraving of a snake in addition to the elongated perforation (Fig. 15:2). 6. Stone Bowls- Forty two of which were found in 29 sites, almost all were made of limestone, a few are of 21

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 14. Eastern ‘Uvda Valley, a pair of stone images at the end of an elongated cell (16 and 14 cm above surface, scale in the arrow- 5 cm).

Figure 15. Naḥal Roded, three examples of stones with elongated perforation, the middle one bears an engraving of meander or a snake.

Figure 16. Examples of stone bowls: 1. Naḥal Roded, natural bowl. 2. Naḥal Roded, a complete small bowl and fragments of large one. 3. Naḥal Botem, a well shaped bowl.

22

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel

Figure 17. Naḥal Roded, a ‘vase-shaped’ installation, as found and after cleaning.

Figure 18. Naḥal Roded, a ‘vase-shaped’ installation with a head-down stone image: left- As found, middle- after cleaning, right- with minor restoration.

sandstone, and most were found broken. The bowls are 15 to 40 cm in diameter and they differ greatly in workmanship. Sixteen bowls were well shaped, one was uniquely shaped, and 25 were totally naturally created by chemical or biological weathering (Fig. 16).

Second is a ‘miniature house’ ca. 30 x 20 x 20 cm, also built of flagstones. Three individual houses were found in one site, one was long and divided into three rooms. Two pairs were found in two other sites. In one, they were built next to a regular pair of cells (Fig. 19), in the other they were built on the narrow end of the elongated cell (Fig. 20), similar to the location of the pairs of stone images (Fig. 14), but spaced. Two other pairs of ‘houses’ and several single ones were also recorded, that were unrelated to Rodedian sites. Often, small stones or flint items are vertically set into the ‘houses’ floor.

Stone features Two types of features were found in the sites, though in a lower frequency than stone objects. One is a small, square, ‘vase-shaped’ installation, ca. 15 x 15 cm, usually built of four flagstones set vertically into the ground. To date, 68 such installations were recorded in 47 sites. Their original content or use is currently unknown (only three were cleaned, Fig. 17). However, as mentioned above, an anthropomorphic stone image was set in two of them, and in another, a stone image was incorporated, with the head down (Fig. 18).

‘Vase-shaped’ installations and ‘miniature houses’ are also found next to ancient roads and to open-air sanctuaries (6th-5th millennia BC).2 A ‘village’ of 11 ‘miniature houses’ was found next to an ancient road west of Be‘er Menucha, in the central ‘Araba Valley, quite well preserved although most roofs had fallen. 2 

23

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 19. Naḥal ‘Eteq, a pair of miniature houses as found, with a natural stone image lying on the flagstone in front of the left one.

Figure 20. Naḥal Demama, two miniature houses built on the end of an elongated cell.

Buried items Small probes excavated adjacent to specific features revealed the phenomenon of ‘buried’ stone items. Two perforated maṣṣeboth were found in two different sites, set in the ground with the perforation facing down (Fig. 21), recalling the anthropomorphic stone image set with the head down in a ‘vase-shaped’ installation (Fig. 18). As mentioned above, one stone image was found buried in the circular cell of a regular pair, with the head up. In three sites, we encountered the very top of a vertically set stone, just 1 cm above surface. Excavation of small probes around two of them exposed regular maṣṣeboth deliberately buried vertically into the ground to a depth of 25 and 34 cm, with attached small offering benches. In one site, a perforated maṣṣebah was set on the side next to the buried regular one (Fig. 22). In the third case, a perforated maṣṣebah was buried with the perforation up (Fig. 23).

Figure 21. Naḥal Roded, perforated maṣṣeboth set with the perforation down, originally set into the ground to the level of the light/dark line.

24

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel

Figure 22. Naḥal Roded, regular and perforated buried maṣṣeboth: 1. As found. 2. As originally set before being covered.

Figure 23. Naḥal Roded, a perforated buried maṣṣebah: 1. As found. 2. As originally set, before being covered.

Artefacts and dating the sites

appearance and brought to the site as offerings. Copper nodules were also found in a number of sites, one was naturally anthropomorphic, 37 mm long (Fig. 24). In two sites animal bone fragments were observed, most probably indicating sacrifices.

Currently, artefacts from the sites were collected from the surface only, but they do provide some information as to the sites’ nature and general date. In several sites, natural rocks of unusual colours or shapes were found, as well as seashells, similar to the finds in maṣṣeboth sites and in open-air sanctuaries (Avner 2002 Chs. 4, 5, App. 1; 2018; Avner and Horwitz 2018). This implies that these items were intentionally selected due to their

Flint items were found in 192 sites, divided into two main groups. One (in 110 sites) consists of bidirectional blades, cores, and a few arrowheads, all are typical to the PPNB (8th-7th millennia BC, Fig. 25). The second (in 167 25

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead In addition to the flint, two 14C dates are currently available, retrieved from a shallow ash spot in a site above Naḥal Roded, dated ca. 7000 BC (Fig. 27). An OSL date measured in a nearby site was 4200±600 years BP (ca. 2200 BC, Sohabti et al. 2015), probably indicating a later intervention in the site. Continuation of the sites to the 6th millennium BC is presently supported by circumstantial, but significant evidence. In the cluster of open-air sanctuaries of Jebel Ḥashm al-Taref, eastern Sinai, a perforated maṣṣebah was found in situ, within a group of nine animals made of small stones set into the ground (Fig. 28). The site is dated by three 14C dates to the middle and late 6th millennium BC (Segal and Carmi 1996, 102; Eddy and Wendorf 1999:192, 280; Avner 2002 Table 1:47). Another open air sanctuary with stone-drawings of animals, in ‘Uvda Valley, is also dated to the mid-6th millennium BC by four 14C dates (Avner 2002, Table 1:20). The same technique of building the stone drawings, of small flagstones set into the ground, was shared by many of the Rodedian elongated cells and by courtyards of

Figure 24. Naḥal Roded, a natural, anthropomorphic copper nodule (two views).

Figure 25. Naḥal Roded, flint blades typical for the PPNB, an ‘Amuq arrow-head (center), on the and a Byblos arrow-head from a site on Har ‘Eshet (central ‘Araba, on the right).

Figure 26. Naḥal Roded, ad-hoc tools, typical for desert sites of the 6th-3rd millennia BC.

sites) comprises ad-hoc tools and flakes, characteristic to the Late Neolithic, Chalcolithic and Early Bronze Age (6th-3rd millennia BC, Fig. 26). In 85 sites, both groups occurred together.3 3 

some open sanctuaries (Avner 1984:22, Pl. 21: 3; 2002 Figs. 5:109-14). The many schematic anthropomorphic stones found in the Rodedian sites find parallels in several Late Neolithic sites (Stekelis 1972, Pl. 50:4; Yzraeli-Noy

See details of the flint analysis, by O. Barzilai, in Avner et al. 2014.

26

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel

Figure 27. Naḥal Roded Site 110, radiocarbon dates from an ash spot. By L. Cummings, the PaleoResearch Institute, Golden Colorado. Calibrated with OxCal 4.2 (Ramsey 2013). Mean dates (red lines) are only approximate, based on the dominant peak in the curve.

Figure 28. Jebel Ḥashem alTaref, eastern Sinai, a perforated maṣṣebah among a group of animal ‘stonedrawings’.

1999:111; Gebel 2010, Abb. 2, 5), but also in the PPNB site of Shaqrat Mazyad, southern Jordan.4 Similarities are also found in the Late Neolithic cemetery in Eilat, dated by many artefacts and by ten 14C dates from 5490 BC to 4350 BC, situated only 3 km away from the densest cluster of the Rodedian sites. At this cemetery, a natural, unshaped anthropomorphic maṣṣebah, 64 cm high, was set in Tomb V, with a ‘nest’ of six skulls next to it (Avner 1991b, 2002, figure. 10:19; Eshed and Avner 2018). Also, one of two open sanctuaries in the site is anthropomorphic, built of small stones (Avner 2002, Figs. 5:94, 95).

Altogether, the flint assemblages, the two 14C dates currently available and the circumstantial evidence, indicate the sites first appeared around 7000 BC and continued through the 6th millennium. Pottery sherds, however, were not found in any of these sites, so it seems that they did not continue into the 5th millennium BC. Discussion The Rodedian sites are obviously different from habitations, campsites or other utilitarian installations. Their symbolic, cultic nature is evident from their location, manner of construction (small stones outlines) and content. Following is a discussion of several characteristics of these sites:

The stone image from Shaqrat Mazyad has not been presented yet in academic publication but see the website- http://shkaratmsaied. tors.ku.dk/images/DSC_0434_bone_figurine.jpg (August 2016). 4 

27

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead 1. Location

3. Local context

As described above, almost all Rodedian sites are located on mountains, which are unsuitable and impractical for habitations, offering no subsistence opportunities (besides ibex hunting). In fact, as noted above, habitations or campsites of other periods are not found on the mountains, while the four small habitations discovered on the Eilat Mountains are all associated with the Rodedian sites. It seems, therefore, that the only reason for their unique location was their cultic role. In later periods, throughout history, mountains were often perceived as the seat of the gods, mountain gods occupied a high position in various pantheons, temples were built on mountain tops and served as a focus for pilgrimage.5 Possibly, the Neolithic Rodedian sites represent the first real evidence for this concept.

The numbers of Rodedian sites becomes more striking when compared to other types of sites known to date in the Negev, dating to the Early Neolithic and more so of the Late Neolithic. In the entire Negev ca. 30 PPNB habitation sites are known (Barzilai 2010: 84-86, 106-109), ten are in the southern Negev, a few more are flint industry or campsites. From the desert Late Neolithic, the information is even poorer. Only three stone-built habitations sites are currently safely dated as Late Neolithic in the Negev, northeastern Sinai and southern Jordan: Qadesh Barne’a 3 (Bar Yosef 1987), Tell Wadi Feinan (Najjar et al. 1990) and Naḥal ‘Issaron Stratum B (Goring-Morris and Gopher 1983, 1987; Carmi et al. 1994). Additional Late Neolithic living levels were found in probes beneath three habitations of the 5th3rd millennia BC in ‘Uvda Valley (Avner 2002: 138-140, Table 1:21,25), including Site 160 recently excavated (by Nadel and Avner).6

The sites’ position on topographic ‘shoulders’, on relatively flat areas, would have enabled several tens of people to stand around in a ritual, an extended family for example. Commonly, a broad view is seen from the sites and possibly, the scenery was one element in selecting their location. The trails leading to the sites indicate that they were repeatedly visited, over a significant period of time and by many people, enough to create trails, which are still visible after several millennia. The fact that some sites contain several installations, many stone objects and both types of flint items, support the view that they were not just temporary.

While the large numbers of Rodedian sites stand in sharp contrast to the scarce habitations, they accord well with the larger numbers of other types of Late Neolithic cult and burial sites. Maṣṣeboth shrines became very common in the desert from ca. 6000 BC, over 450 are presently recorded in the Negev and eastern Sinai;7 open-air sanctuaries first appeared during the PPNB,8 of which 220 are currently recorded, and so are hundreds of burial sites with ample innovations in mortuary customs. Hundreds of sites of these types are already recorded, as well as other types of cult installations (Avner 2002, Ch. 4, 5 and App. 1, 2018). Now, the bulk of desert cult sites is greatly enlarged by the Rodedian sites.

2. Distribution The area currently surveyed in detail is very limited compared to the general mountainous zones of the Negev, Sinai and southern Jordan. Even so, the presently known distribution of Rodedian sites indicates that many more still await discovery. This is also true of areas previously surveyed, such as the Negev Highlands. If 363 Rodedian sites are presently recorded in the southern Negev alone, one may think now of a vast phenomenon, perhaps even thousands of mountain cult site in the desert. However, mapping the real extent of these sites requires a great investment of time and energy, by many survey teams.

4. State of preservation Disruption of sites and the presence of broken stone objects recall the phenomenon of breaking clay figurines in Neolithic sites, interpreted as the result of a ritual act (e.g. Rollefson 2000: 168; Kuijt 2000: 150; Chapman 2001; Verhoeven 2007; Orrelle 2014: 104-5). This may also be the explanation for the situation in Rodedian sites, but there may have been another cause. Disturbed sites and broken stone objects were recorded mainly on the igneous mountains around Naḥal Roded. In this area, there are two gold production sites, dated

A few examples: In the bible, Yahweh is the God of of Mount Sinai, M. Se‘ir, M. Paran (Deut. 33:2; Hab. 3:3) or a mountain God in general (1 kgs 20:28). In the Ugaritic texts, Mount Zaphon is the seat of El, of Ba‘al and of the assembly of gods (Nieher 1995). For the sacredness of mountains in both the Bible and Ugarit see Clifford 1972, Chs. 2, 3. In ancient Greece, Mount Olympus was the seat of the gods (Mussies 1995). In the Nabataean religion, Dushara, the head of pantheon, was a mountain god (‘the one of Jebel Ash-Shara’, Healey 2001: 87; Zayadin 2003: 59). Cult places were built for him on several summits around Petra (Brünnow and Domaszewski 1904: 239-249; Musil 1907, 80-97; Dalman 1908: 157-83, 332-44) and on the summit of Jebel Serbal in southern Sinai (Avner 2015). Christian pilgrimage to Jebel Musa, identified by various Churches as Mount Sinai, began with Egeria in the 4th century A.D. (Caner 2010: 211-31) and continues still today with ca. 300,000 pilgrims annually (pers. com. Father Justine in the St. Katrine Monastery, July 2009). 5 

A trench excavated across the central courtyard to a depth of 50 cm yielded two 14C dates. The lower (PRI-12-082-2B) was 6116±20 BP, calibrated around 5100 BC We thank Linda Cummings of the PaleoResearch Institute, Golden Colorado for analyzing the samples. 7  This number of maṣṣeboth refers to independent shrines only and excludes those incorporated in the Rodedian sites, in tombs or in open-air sanctuaries (Avner 1984, 2002, Chs. 4, 5, App. 1 and Tables 1, 11, 14; Avner and Horwitz 2018; Avner 2018). 8  Presently, the earliest 14C dates from open sanctuaries are in mid 6th millennium BC (Avner 2018, Tables 1, 2), however, in some of them typical, bidirectional PPNB flint blades were found on surface, suggesting their earlier appearance, in the 7th millennium BC. 6 

28

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel to the Early Islamic periods (Avner 1993a; Avner and Magness 1989: 44-5; Gilat et al. 1993). Gold industry here must have been preceded by a ‘geological’ survey of that region. Possibly, the Islamic surveyors viewed the maṣṣeboth (regular and perforated), the anthropomorphic stones and the stones with elongated perforation as a violation of Islamic law, and therefore broke them as an act of iconoclasm.9 Nevertheless, since no Early Islamic pottery sherds or other evidence were found in Rodedian sites, this possibility remains purely theoretical.

as representing the ancestors.10 Effigies are also well known in many recent and present-day traditional societies as representing the ancestors.11 We propose, therefore, to interpret also the stone images in the Rodedian sites as representing the ancestors. Notably, all types of stone images in the Rodedian sites are schematic, so they still conform with the basic abstract, desert notion (or theology) reflected by the maṣṣeboth. The miniature houses recall an Early Neolithic house model from Çayönü (Biçakçi 1995) and a Late Neolithic one from Jericho (Garstang 1936: 71, Pl. 40b). Later, Chalcolithic ossuaries were also often shaped like houses, interpreted as houses for the dead (Callaway 1963:80; Perrot and Ladiray 1980, passim).12 In ancient Egypt, house models were found mainly in tombs. They were termed ‘soul houses’ (Petrie 1907: 14-20, Pls. 1522) or ‘Ka houses’ (Kaplony 1980) i.e. houses for spirits of the dead. Hundreds of house models of later periods from the Near East and beyond have garnered various interpretations (Katz 2016, Ch. 7, with references) including houses for spirits of the dead. Miniature houses for the ancestors are even known in recent and present-day traditional societies.13 Based on the examples from different times, we suggest that also in the Rodedian and other desert sites, the miniature houses were built to provide dwellings for the ancestral spirits.

5. Interpretation of elements Naturally, the unique characteristics of the Rodedian sites and their content provoke attempts at interpretation, even if preliminary: Regular maṣṣeboth are probably better understood than the other elements, based on their frequent occurrence in the Near East, especially in the deserts, based on later written sources and on previous studies (see references in Avner 1984, 1990, 2000, 2001, 2002a, 2002 Ch. 4, 2018). It is quite safe to say that desert maṣṣeboth mainly represented deities, perceived as containing in them the deities’ power and spirit. In this, they resembled statues of gods, but were actually the opposite of statues since they were unshaped in principle (cf. Exodus 20:3,22 etc.). Already in prehistory, desert people developed a specific theology, refraining from figurative representation of their deities, a theology that persisted later in the Israelite religion, the Nabataeans and Islam, all religions with desert foundations. Groups of maṣṣeboth of repeating numbers represented ‘organic’ groups of deities of the same numbers, as are later known from dedication inscriptions, from mythologies and from ample artistic presentations (Avner op. cit.).

Several additional items seem to symbolise death: The upside-down setting of perforated maṣṣeboth (with the ‘eye’ down), the burial of regular and perforated maṣṣeboth (with the ‘eye’ up but below surface) and in the setting of anthropomorphic stone images with their head facing down (Figs. 21-23). Stones with elongated perforation generally recall the Yarmukian grooved pebbles, interpreted as symbolising the vulva, i.e. a female fertility symbol (e.g. Stekelis 1972:25-27, 33; Gopher and Orrelle 1996). Currently, this

As to the perforated maṣṣeboth, presently we cannot offer any interpretation, but only a guess that the perforation may represents an ‘eye’. The naturally holed stones most probably bore the same meaning as the perforated ones. The great number of the latter may indicate that they were preferred over the perforated ones, probably since they were created by nature (i.e. by gods).

See e.g. Tubb and Grissom 1995; Schmandt-Besserat 1998, 2013; Rollefson 2000, esp. 185; Mabri 2003. From the site of Rizqeh, southern Jordan, Kirkbride (1969) published three types of stone monuments: unshaped standing stones, anthropomorphic silhouetted stones and figurative human images. Based on ‘Arabian sources she interpreted them as ancestral monuments. Kirkbride dated the site close to the 1st century BC, but she also presented Late Neolithic artifacts and a 14 C date- ca. 4900 cal. BC. In a meeting with Avner in March 1983, she accepted that the older artifact and radiometric date do belong to the site and to the stone monuments. 11  See e.g. Frazer 1913a VIII: 53, XI: 155 etc.; Frazer 1913b: 307-317, 321; Mack 1986: 86-92; Colson 2005 and a list of references in Mabri 2003: 102. In 1999, one of the writers (Avner) visited five islands in Indonesia, in all of them he saw and photographed maṣṣeboth, menhirs and ancestral effigies. Most famous are the ‘Tau-Tau’ in the Tana Toraja, Sulawesi– wooden statues representing the ancestors, standing on verandahs cut into cliffs next to rock-cut tombs (Sandarupa 1996: 70-71; Louis-Vincent 2005: 3237; Waterson 2009: 381-4). 12  For interpretation of the ossuaries as representing granaries, not houses, see Bar-Yosef and Ayalon 2001. 13  See e.g. Frazer 1913b: 315-6; Colson 2005: 1508; Ray 2005: 2577; Rabuzzi 2005: 4106; Middleton 2005: 5527. 10 

Anthropomorphic images, whether schematic or figurative, made of stone, clay or plaster, are known from several Neolithic sites and mostly interpreted For prohibition of images in the Islam see Sell 1914; Wensinck 1997. For islamic iconoclasm see Vasiliev 1956; Kirkbride 1969. Still recently (2001-2016), the world media frequently reports massive destruction of ancient sites, art works and Christian symbols by Taliban, ISIS and others in Afghanistan, Iraq and Syria, and see a special issue on the cultural heritage loss in the Near East- Near Eastern Archaeology 78 (Sep. 2015). 9 

29

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 29. Naḥal Roded, a vertical view of installation with a stone alignment, 6.5 m long, connecting to a circle, and a limestone image on the right (enlarged on the lower right corner).

is the only interpretation we can offer for these stones in the Rodedian sites.

Unlike a number of Neolithic habitations in the southern Levant, the Rodedian sites show continuity from PPNB to LN.16 They demonstrate intensive ritual activity, a rich spiritual culture and a fully established religion of the desert inhabitants in these periods, not just a ‘belief system’, ‘ideology’ or ‘symbolic behaviour’. Without entering into a discussion on theories and definitions of religion, it is enough to mention here the consistency of elements in the sites and the repeating numbers of maṣṣeboth in groups, which indicate the existence of a complex pantheon and mythologies. In addition, the four small habitations built on the mountains within clusters of Rodedian sites may hint at a clergy serving and guarding the sites, i.e. an establishment.

The frequent recurrence of an elongated cell pointing to a circle is also intriguing. By itself, the repetition of this pattern indicates some idea or symbolism. In addition, the design resembles ample examples of stone structures in the desert, combining alignments with circles or with circular tombs; one of the Rodedian sites was built just as a stone alignment connecting to a circle (Fig. 29). Dimensions of these alignments vary greatly, from a few meters to several kilometres long (e.g. Haiman 2000; Kennedy 2011). The same combination of forms is also common in world rock-art as a circle and a line, or ‘balloon’. Based on the connection of stone alignments to burials, and the association of ‘balloons’ with other symbols in rock-art, it is suggested that in both cases they actually symbolise the combined major powers in nature, or of deities, the male and female, i.e. fertility. Fertility for the dead through stone structures, fertility for the living through rock art (Avner and Avner 1999).14

Several mysteries remain in this stage of research. The primary question is- what caused the ‘eruption’ of so many and such a variety of cult sites in the desert during the 7th-6th millennia BC? Until recently, artifacts and radiometric dates pointed to the abrupt proliferation of cult sites and major innovations in burial customs during the 6th millennium BC. The suggested explanation was that the religious development followed the major economiccultural shift, from hunting and gathering to farming and herding, which took place in the desert around 6000 BC (Avner 1998, 2002, Chs. 2, 9, 10, 2006; Rosen et al. 2005; Rosen 2015). Now, with the discovery of hundreds of Rodedian sites, and knowing that they appeared a full millennium before the adoption of farming and herding by the desert societies, another explanation is required.17

If the above interpretations are accepted, two principal symbolic aspects unfold: death and fertility. They seem contradictory, but the combination of both is actually characteristic of ancestral cult, a highly important element in the life of traditional societies worldwide.15 Summary The many Rodedian sites presently recorded add a new dimension to our knowledge of Neolithic desert cultures. They show that many more people lived in the desert than can be inferred from the small numbers of PPNB and LN habitations currently known in the Negev.

For a crisis at the end of PPNB see Rollefson et al. 1992; Simmons 1997, 2007:43-5; Bar-Yosef 2001:26-8. 17  A similar question rises about the sequence of developments in the Levant further north during the 10th-7th millennia BC, whether the ‘religious revolution’ followed the agricultural one (e.g. Childe 1962, Bar-Yosef 2001), preceded cultivation (Cauvin 2000a,b; Schmidt 2000:48, 2012:226-242, 256) or whether both revolutions were simultaneous (Watkins 2010, 2011). No consensus has been reached on this question, which also depends on another debated issue- was 16 

For the dominance of the male and female symbols in the Neolithic Near East see Orrelle 2014. 15  E.g. Frazer 1913b; Mack 1986; Radin 1991; Matclif and Huntington 1991; Kiong 1993; Barley 1997; Murray-Parker et al. 1997; Hardacre 2005). 14 

30

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel Presently, we have no clear answer to this question. However, it should be noticed that contemporarily with the appearance of Rodedian sites, cult installations, buildings and artifacts are found in the broader southern Levant. For example, sanctuaries with maṣṣeboth in Jericho (Kenyon 1957, Pl. 17) and ‘Ain Ghazal (Rollefson 1999:181-3; Kafafi 2010, 2013), plaster statues in the same sites,18 cult installations with maṣṣeboth at esSifiya (Maḥasneh 2000), Shaqrat al-Mazyad (Jensen et al. 2005:119, 124), al-Baseet (‘Amr 2004) and Beidha (Kirkbride 1968:92-6),19 burials rich with symbolism at Kfar HaḤoresh (Goring-Morris 2000, 2005; GoringMorris & Horwitz 2004; Simmons et al. 2007) and many anthropomorphic and zoomorphic figurines in PPNB sites.20 All these, and more, are only part of a broader and profound cognitive, cultic and artistic development that actually began in the Natufian culture (e.g. BarYosef and Belfer-Cohen 1998; Shaham and Belfer-Cohen 2013) and greatly intensified through the subsequent Neolithic phases.21 Does this means that desert societies were influenced by developments in the north? Not sure. Presently, the Rodedian sites seem to be unique to the desert.

Smelensky, 14C analysis- by L. Cummings, photos and plan of Fig. 5 were made by U. Avner. Postscript In December 2017, excavation was carried out in one cult site above Naḥal Roded (Site 110), by M. Birkenfeld, L. K. Horwitz and U. Avner. A circular structure began to be exposed, the ‘small’ ash spot was found to be at least 5 x 5 m and at least 20 cm deep. The site is highly rich with flint debitage, several ‘Amuq Points were found, typical to the PPNB, as well as hundreds of raptor bones and some Red Sea shells. Two radiocarbon dates from the present bottom of the dig in the ash spot (by L. Scott Cummings, the PaleoResearch Institute, Golden, Colorado) were ca. 200 years earlier than those from the top of the ash (see above). First report on the dig is forthcoming in Antiquity. References Abbo, S., Lev-Yadun, S. and Gopher, A. 2010. Agriculture Origins: Centers and Concenters; A Near Eastern Reappraisal. Critical Reviews in Plant Sciences 29: 317328. ‘Amr, H. 2004. Note on al-Baseet, a New central Settlement in Wadi Musa. In, H. B. Beignet, H. G. Gebel and R. Neef (eds), Central Settlements in Neolithic Jordan: Proceeding of a Symposium Held in Wadi Musa, Jordan, 21st-25th of July 1997. Berlin. 6569. Anati, E. 2001. The Riddle of Mount Sinai. Capo di Ponte, Edizioni del Centro. Anati, E. 2005. Har Karkom, A Guide to Major Sites. Capo di Ponte, Centro camuno di studi peristorici. Anati, E. and Mailland, F. 2009. Map of Har Karkom (229). Capo di Ponte, Centro internazionale di studi preistorici ed etnologici. Anderson, P. C. 1998. History of Harvesting and Threshing Techniques for Cereals in the Prehistoric Near East. In A. B. Domain, J. Valkoun, G. Willcox and C. O. Qualset (eds), The Origins of Agriculture and Crop Domestication. The Harlan Symposium. Pp. 145159. Aleppo, International Center for Agricultural Research m the Dry Areas; International Plant Generic Resources Institute; Food and Agriculture Organization of the United Nations; Genetic Resources Conservation Program, Division of Agriculture and Natural Resources, University of California. Avner, U. 1979. A Survey in the ‘Uvda Valley. Hadashot Arkheologiyot 1-3: 69-71 (Hebrew). Avner, U. 1982a. Biq’at ‘Uvda Survey. Excavations and Surveys in Israel 1: 79-80. Avner, U. 1982b. Survey of Biq’at ‘Uvda-Eilat Road. Excavations and Surveys in Israel 1: 81-2. Avner, U. 1984. Ancient Cult Sites in the Negev and Sinai Deserts. Tel Aviv 11: 115-31.

Once the surrounding mountainous desert areas are meticulously surveyed, the numbers of Rodedian sites will certainly increase, their geographic distribution will come clearer, the assemblages of associated objects and features will expand. Excavations will supply additional details and radiometric dates, to help determine their time span. Surveys and excavations may shed new light on the role of these sites in the cultures of the desert Neolithic societies and on their place in the general development of Near Eastern religions. Acknowledgments We are thankful to the Irene Levi-Sala CARE Archaeological Foundation for enabling the initiation of the cult sites survey, to O. Bar-Yosef and to L. Kolska Horwitz for their, fruitful comments on the manuscript and for adding references. Maps (Figs. 2, 3) were prepared by R. Shemtov, the flint (Fig 25) drown by M. the advent of agriculture a slow process (Hillman 1984; Willcox et al. 2008; Bar-Yosef 2001, 2014; Groman-Yaroslavski et al. 2016) or a fast one (Anderson 1998; Abbo et al. 2010; Gopher and Abbo 2016). 18  Garstang 1935:166-7, Pls. 25-6, 52-3; Kenyon 1960:91-2. Pl. 12a; Rollefson 1983, Walker and Grissom 1995; Schmandt-Besserat 1998, 2013, Ch. 7.2; Grissom 2013. 19  A shaped, narrow stone set in a room (Kirkbride 1968, Pl. 24 A) was not identified by Kirkbride as maṣṣebah, but in a conversation with her in March 1983 she accepted this interpretation. Originally she dated this room to the Natufian stage of the site, but it was not included in the final report on the Natufian Beidha by Byrd (1989). In a letter from March 2001, Byrd attributed the room to the earliest Neolithic stage of the site. 20  E.g. Kuijt and Chesson 1988; McAdams 1997; Maḥasneh 1999; Cauvin 2000:105-8; Marbi 2003; Bailey 2005; Makarewicz 2006:21-2; Verhoeven 2007; Shcmandt-Besserat 2013, Chs, 3-5. 21  Mithen 1996, 2003; Kuijt (ed.) 1999; Cauvin 2000a,b; Lewis-Williams and Pearce 2005; Hodder 2006; Watkins 2010; Schmidt 2012.

31

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Avner, U. 1990a. Ancient Agricultural Settlement and Religion in the ‘Uvda Valley in Southern Israel. Biblical Archaeology 53: 125-41. Avner, U. 1990b. Excavation of Tumuli Tombs in Eilat. Excavations and Surveys in Israel 9: 76-8. Avner, U. 1991. Late Neolithic-Early Chalcolithic Burial Site in Eilat. American Journal of Archaeology 95: 496-7. Avner, U. 1993a. Wadi Tawahin, an Early Islamic Gold Exploitation Site. American Journal of Archaeology 97: 160-2. Avner, U. 1993b. maṣṣeboth Sites in the Negev and Sinai and their Significance. In, A. Biran and J. Aviram (eds), Biblical Archaeology Today, Proceedings of the Second International Congress on Biblical Archaeology. Jerusalem, Israel Exploration Society: 166-181. Avner, U. 1998. Settlement, Agriculture, and Paleoclimate in ‘Uvda Valley, Southern Negev Desert, 6th-3rd Millennia BC In. A. Issar and N. Brown (eds), Water, Environment and Society in Times of Climate Change. Pp. 147-202. Dordrecht, Springer Netherlands. Avner, U. 2000. Nabatean Standing Stones and Their Interpretation. Aram 11-12: 97-122. Avner, U. 2002. Studies in the Material and Spiritual culture of the Negev and Sinai Population, During the 6th-3rd millennia BC. PhD dissertation, The Hebrew University of Jerusalem. http://www.adssc.org/ sites/default/files/PhD-Uzi-RS.pdf. Avner, U. 2006. Settlement Pattern in the ‘Araba Valley and the Adjacent Desert Areas, a View from the Eilat Region. In P. Bienkowski and K. Galor. (eds), Crossing the Rift. Oxford, Oxbow: 51-74. Avner, U. 2015. The Nabataeans in Sinai. Aram 27: 1-33. Avner, U. 2018. Protohistoric Developments of Religion and Cult in the Negev Desert. Tel Aviv 45:23-62. Avner, U. and Avner, R. 1999. Circles, Triangles and Lines in Desert Archaeological Remains and Rock Engravings, and Their Interpretations. In. P. Bahn and A. Fossati (eds), Rock Art Studies: NWES of the World 1. Proceeding of the International Rock Art Congress, Turin 1995. Torino and Pinerolo, Oxbow. (CD). Avner, U. and Magness, J. 1998. Early Islamic Settlement in the Southern Negev. Bulletin of the American Schools of Oriental Research 310: 39-57. Avner, U., Shem-Tov, M., Enmar, L., Ragolski, G, ShemTov, R. and Barzilai, O. 2014. Survey of Neolithic Cult sites in the Eilat Mountains, Israel. Journal of the Israel Prehistoric Society 44: 101-116. Avner, U. and L. K Horwitz. 2017. Animal Sacrifice and Offering from Cult and Mortuary Sites of the Negev and Sinai, 6th-3rd Millennia BC. Aram 29:35-70. Bailey, D. W. 2005. Prehistoric Figurines: Representation and Corporeality in the Neolithic. London and New York, Routledge. Barley, N. 1997. Grave Matter: A Lively History of Death Around the World. New York, Henry Holt and Company.

Bar-Yosef, O. 1987. The Prehistory of the Sinai Peninsula. In G. Gvirtzman, A. Shmueli, Y. Gradus, I. Beit-Arieh and M. Har-El. (eds), Sinai: Part Two: Sinai―Human Geography. Tel Aviv, Ministry of Defence: 559-578 (Hebrew) Bar-Yosef, O., 1998. On the Nature of Transitions: The Middle to Upper Palaeolithic and the Neolithic Revolution. Cambridge Archaeological Journal 8 (2), 141-63. Bar Yosef, O. 2001. From Sedentary Foragers to Village Hierarchies: The Emergence of Social Institution. In G. Runciman (ed.), The Origin of Human Social Institutions. Proceedings of the British Academy, 110. Oxford, Oxford University Press: 1-38. Bar-Yosef, O. 2011. PPNB Interaction Sphere. Cambridge Archaeological Journal 11: 114-120. Bar-Yosef, O. 2014. Southern Turkish Neolithic: A View from the Southern Levant. In M. Özdogan, N. Baflgelen and P. Kuniholm (eds), The Neolithic in Turkey 6: 293-320. Istanbul, Archaeology and Art Publications. Bar-Yosef, O. and Belfer-Cohen, A. 1998. Natufian Imagery in Perspective. Revista di Scienze Preistoriche 49: 247-263. Bar-Yosef, O. and Ayalon, E. 2001. Chalcolithic Ossuaries – What do they imitate and why? Qadmoniot 34: 3443 (Hebrew). Barzilai, O. 2010. Social Complexity in the Southern Levantine PPNB as Reflected Through Lithic Studies: The Bidirectional Blades Industries. British Archaeological Reports International Series 2180. Oxford, Archaeopress. Biçakçi, E. 1955. Çayönü House Models and a Reconstruction Attempt for the Cell-Plan Buildings. In Reading in Prehistory: Studies Presented to Halet Çambel: 101-127. Istanbul, Graphis. Brünnow, R. E. and Domaszewski, A. 1904-1909. Die Provincia Arabia 1-3. Strassburg, Truebner. Byrd, B. F. 1989. The Natufian Encampment at Beidha. Moesgard, Aarhus. Callaway, J. 1963. Burial in Ancient Palestine from the Stone Age to Abraham. Biblical Archaeology 26: 74-91. Caner, D. 2010. History and Hagiography from the Late Antique Sinai. Liverpool, Liverpool University Press. Carmi, I., Segal, D., Goring-Morris, A. N. and Gopher, A. 1994. Dating the Prehistoric Site Naḥal ‘Issaron in the Southern Negev, Israel. Radiocarbon 36: 391-98. Cauvin, J. 2000a. The Birth of the Gods and the Origins of Agriculture. Cambridge, Cambridge University Press. Cauvin, J. 2000b. The Symbolic Foundation of the Neolithic Revolution in the Near East. In, I. Kuijt (ed.), Life in Neolithic Farming Communities: Social Organization. Identity and Differentiation: 235-251. New York, Springer Science and Business Media: 235-251. Chapman, J. 2001. Object Fragmentation in the Neolithic and Copper Age of Southern Europe. In, P. F. Biehl. The Archaeology of Cult and Religion: 8-106. Budapest: Archaeolingua Alapitvany. 89-106. 32

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel Childe, V. Gordon 1958. The Prehistory of European Society. Hammondsworth, Penguin Books. Clifford, J. 1972. The Cosmic Mountain in Canaan and in the Old Testament. Cambridge, Harvard University Press. Colson, E. 2005. Central Bantu Religion. In L. Johns (ed.), Encyclopaedia of Religion, second edition. Detroit, Macmillan Reference: 1505-1512. http:// www.e-reading.club/bookreader.php/133766/ Encyclopedia_of_religion. Dalman, G. 1908. Petra und Seine Felsheiligtümer. Leipzig, J. C. Hinrich. Eddy, F. W. and Wendorf, F. 1999. An Archaeological Investigation of the Central Sinai, Egypt. Boulder, University Press of Colorado. Journal 68:5-29. Eshed, V. and Avner, U. In press. Late Neolithic Burial Site in Eilat, by the Red Sea, Israel. Israel Exploration Journal. Frazer, J. 1906-1915. The Golden Bough, a Study of Magic and Religion. 12 vols. 3d ed. London, Macmillan. Frazer, J. 1913. The Belief in Immortality and the Worship of the Dead. 3 vols. London, Macmillan. Garfinkel, J. 2004. The Goddess of Sha’ar HaGolan: Excavations at a Neolithic Site in Israel. Jerusalem, Israel Exploration Society. Garstang, J. 1936. Jericho: City and Necropolis, Report for sixth and Concluding Season, 1936. Annals of Archaeology and Anthropology 23: 67-100. Garstang, J. and Garstang, J. B. E. 1948. The Story of Jericho. revised edition. London and Edinburgh, Marshall, Morgan and Scott, Ltd. Gebel, H. G. K. 2010. Untergegengen im Klimawandel: Die paläo-beduinische Kultur von Qulban Beni Murra, Jordanien. Antike Welt 10: 40-4. Gilat, A., Shirav, M., Bogoch, R., Halicz, L., Avner, U. and Nahlieli, D. 1993. Significance of Gold Exploitation in the Early Islamic Period, Israel. Journal of Archaeological Science 20:429-437. Gopher, A. 1994. Arrowheads of the Neolithic Levant. American Schools of Oriental Research Dissertation Series 10. Winona Lake, Eisenbrauns. Gopher, A. and Orrelle, E. 1996. An Alternative Interpretation for the Material Imagery of the Yarmukian, a Neolithic Culture of the Sixth Millennium BC in the Southern Levant. Cambridge Archaeological Journal 6:255-279. Gopher, A. and Abbo, S. 2016. Plant Domestication and the Origin of Agriculture in the Near East. Tel-Aviv, Resling (Hebrew). Goring-Morris, A. N. 1993. From Foraging to Herding in the Negev and Sinai: the Early to Late Neolithic Transition. Paléorient 19:65-89. Goring-Morris, A. N. 2000. The Quick and the Dead: The Social Context of Aceramic Neolithic Mortuary Practices as seen from Kfar HaḤoresh. In: I. Kuijt, (ed.), Life in Neolithic Farming Communities: Social Organization, Identity and Differentiation. New York, Springer: 103-136.

Goring-Morris, A. N. 2005. Life, Death and the Emergence of Differential Status in the Near Eastern Neolithic: Evidence from Kfar HaḤoresh, Lower Galilee, Israel. In, J. Clark (ed.), Archaeological Perspectives on the Transmission and Transformation of Culture in the Eastern Mediterranean. Oxford, Oxbow: 89-105. Goring-Morris, A. N. and Belfer-Cohen, A. 2002. Symbolic Behaviour from the Epipalaeolithic and Early Neolithic of the Near East. In C. Jensen and H. G. Gebel (eds), Proceedings of the 2nd International Congress on the Archaeology of the Ancient Near East: Magic Practices in the Near East, Copenhagen, Denmark 2000. Berlin, Ex Oriente: 67-80. Goring-Morris, A. N. and Gopher, A. 1983. Naḥal ‘Issaron: A Neolithic Settlement in the Southern Negev. Israel Exploration Journal 33: 149-62. Goring-Morris, A. N. and Gopher, A. 1987. Naḥal ‘Issaron: A Neolithic site in the Southern Negev. Qadmoniot 20: 18-21 (Hebrew). Goring-Morris, A. N. and Horwitz L. K. 2004. Animals and Ritual during the Levantine Pre-Pottery Neolithic B: A Case Study from the Site of Kfar HaḤoresh, Israel. Anthropozoologica 39:165-178. Grissom, C. 2013. The Statuary. In, D. SchmandtBesserat, (ed.), Symbols at ‘Ain Ghazal. Berlin, Ex Oriente: Ch. 7.1. Haiman, M. 2000. The ‘K-Line’ at Har Romem in Light of the Survey of the Map of Har Ramon (203). Atiqot 39:21-29. Hardacre, H. 2005. Ancestor Worship. In L. Johns (ed.), Encyclopaedia of Religion, second edition. Detroit, Macmillan Reference: 320-32. http:// www.e-reading.club/bookreader.php/133766/ Encyclopedia_of_religion. Healey, J. F. 2001. The Religion of the Nabataeans. Leiden, Brill. Hillman, G. 1984. Interpretation of Archaeological Plant Remains: The Application of Ethnographic Models from Turkey. In W. van Zeist and W. A. Casparie (dir), Plants and Ancient Man: Studies in Palaeoethnootany,: 1-57. Rotterdam, Balkema. Hodder, I. 2006. The Leopard’s Tale: Revealing the Mystery of Çatal Höyük. London, Thames and Hudson. Hodder, I. 2011. Symbolism and the Origin of Agriculture in the Near East. Cambridge Archaeological Journal 11:107-112. Jensen, C. H., Hermansen, B. D., Peterson, M. B, Kinzel, M., Hald, M. H., Bangsgaard, P., Lynnerup, N. and Thuesen, I. 2005. Preliminary Report on the Excavations at Shaqarat al-Muzay’id, 1999-2004. Annual of the Department of Antiquities of Jordan 49:115-134. Kafafi, Z. 2010. Clans, Gods and Temples at the LPPNB ‘Ain Ghazal. In, M. Banz (ed.), The Principle of Sharing: Segregation and Construction of Social Identities at the Transition from Foraging to Farming: Proceedings of a Symposium Held on 29th-31st January 2009 at the Albert33

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Ludwigs-University of Freiburg: 301-312. Berlin, Ex Oriente. Kafafi, Z. 2013. Standing Stones of the Neolithic Village of ‘Ain Ghazal. In. D. Schmandt-Besserat. (ed.), Symbols at ‘Ain Ghazal: 353-358. Berlin, Ex Oriente. Kaliszan, L. R., Hermansen; B. D., Jensen, C. H., Skuldbol, T. B. B., Bille, M., Bangsgaard, P., Ihr, I., Sorensen, M. L. and Markussen, B. 2002. Shaqrat Mazyad- The Village on the Edge. Neo-Lithics 1 (02): 16-19. Kaplony, P. 1980. Ka House, in, W. Helck and E. Otto (eds), Lexikon der Ägyptologie III: 284-7. Wiesbaden, Harrassowitz. Katz, H. 2016. Portable Shrine Models: Ancient Architectural Clay Models from the Levant. British Archaeological Reports International Series 2791. Oxford, Archaeopress. Kennedy, D. 2011. Works of the Old Men in Arabia: Remote Sensing in Interior Arabia. Journal of Archaeological Science. 38: 3185-203. Kiong, T. C. 1993. The Inheritance of the Dead: Mortuary Rites among the Chinese in Singapore. Southeast Asian journal of social science. 21: 131-58. Kirkbride, D. 1968. Beidha 1967, An Interim Report. Palestine Exploration Quarterly 100: 90-96. Kirkbride, D. 1969. Ancient Arabian Ancestor Idols: The Discovery of the Sanctuary at Risqeh. Archaeology 22: 116-121, 188-195. Kuijt, I. (ed.) 1999. Life in Neolithic Farming Communities: Social Organization, Identity and Differentiation. New York, Springer Science and Business Media. Kuijt, I. 2000. Keeping the Peace, Ritual, Skull Caching and Community Integration in the Levantine Neolithic. In I. Kuijt (ed.). Life in Neolithic Farming Community: Social Organization, Identity, and Differentiation. New York, Springer Science and Business Media: 137-164. Kuijt, I. and M. S. Chesson 1988. Lumps of Clay and Pieces of Stone: Ambiguity, Bodies and Identity as Portrayed in Neolithic Figurines. In S. Pollok and R. Bernbeck (eds), Archaeologies of the Middle East, Critical Perspectives: 153-183. Oxford, Blackwell Publishing. Kuijt, I. and Goring-Morris. N. 2002. Foraging, Farming, and Social Complexity in the Pre-Pottery Neolithic of the Southern Levant: A Review and Synthesis. Journal of World Prehistory 16: 361-440. Louis-Vincent, T. 2005. Funeral Rites. In, L. Johns (ed.). Encyclopaedia of Religion, second edition. Detroit, Macmillan: 3233-3241. http://www.e-reading.club/ bookreader.php/133766/Encyclopedia_of_religion. Lewis-Williams, D. and Pearce, D. 2005. Inside the Neolithic Mind: Consciousness, Cosmos snd the Realm of the Gods. London, Thames and Hudson. Mack, J. 1986. Madagascar, Island of the Ancestors. London, British Museum Publications. Maḥasneh, Ḥ. M. 1999. Anthropomorphic Figurines from the Early Neolithic Site of Es-Sifiya (Jordan) Zeitschrift des deutschen Pal‫ה‬stina-Vereins 115:109-127. Maḥasneh, Ḥ. M. 2000. The Early Neolithic Mortuary Customs in es-Sifiya, Jordan. The Second Faynan

Conference, Amman April 2000, Abstracts of lectures: 16. Amman. Makarewicz, C. A. and Austin, A. E. 2006. Late PrePottery Neolithic B Occupation at El-Hemmeh: Results from the Third Excavation Season 2006. NeoLithics 2006:19-22. Marbi, J. 2003. The Birth of the Ancestors: The Meaning of Human Figurines in the Near Eastern Neolithic Villages. In B. A. Nakhai (ed.), The Near East in the Southwest: Essays in Honour of William G. DeverL 85-116. Boston, American Schools of Oriental Research. Matcalf, P. and Huntington, R. 1991. Celebration of Death. 2d ed. Cambridge: Cambridge University Press. McAdams, E. 1997. The Figurines from the 1982-5 Seasons at Ain Ghazal. Levant 29:115-145. Mellaart, J. 1975. The Neolithic of the Near East. London, Thames and Hudson. Middleton, J. 2005. Lugbara Religion. In L. Johns (ed.). Encyclopaedia of Religion, 12 vols. 2d ed. Detroit, Macmillan Reference: 5526-5528. http:// www.e-reading.club/bookreader.php/133766/ Encyclopedia_of_religion. Miroschedji de, P. 1993. Cult and Religion in the Chalcolithic and Early Bronze Age. In: A. Biran and J. Aviram (eds), Biblical Archaeology Today, 1990. PreCongress Symposium: Population, Production and Power: 208-220. Jerusalem: Israel Exploration Society. Miroschedji, P. 2011. At the Beginning of Canaanite Cult and Religion: The Early Bronze Age Fertility Ritual in Palestine. In, H. Geva and A. Paris (eds), Eretz Israel 30, Dedicated to Amnon Ben-Tor. 74-103. Mithen, S. 1996. The Prehistory of Mind: A Search for the Origin of Art, Religion and Science. London, Thames and Hudson. Mithen, S. 2003. After the Ice: A Global Human History, 20,000-5000 BC. Cambridge, Harvard University Press. Moore, A. M. T. 1973. The Late Neolithic in Palestine. Levant 5: 36-68. Moore, A. M. T. 1982. A Four Stage Sequence for the Levantine Neolithic, ca. 8500-3750 BC. Bulletin of the American Schools of Oriental Research 246: 1-34. Mortensen, P. 1970. A Preliminary Study of the Chipped Stone Industry from Beidha: An Early Neolithic Village in Southern Jordan. Acta Archaeologica 41: 1-54. Mussies, G. 1995. Olympus. In K. van der Toorn, B. Becking and P. van der Horst (eds), Dictionary of Deities and Demons in the Bible: 1745-1758. Leiden, Brill. Musil, A. 1907. Arabia Petraea, Vol. II. Vienna, Holder. Najjar, M., Abu Dayya, A., Suleiman, E., Weisgerber, G. and Hauptmann, A. 1990. Tell Wadi Feinan: The first pottery Neolithic tell in the South of Jordan. Annual of the Department of Antiquities of Jordan 34: 27-56. Nieher, H. 1995. Zaphon. In, K. van der Toorn, B. Becking and P. van der Horst (eds), Dictionary of Deities and Demons in the Bible: 1218-1220. Leiden: Brill. 34

U. Avner et al.: Neolithic Cult Sites in the Southern Negev, Israel Orrelle, E. 2014. Material Images of Humans from the Natufian to the Pottery Neolithic Periods in the Levant. BAR International Series S2595. Oxford, Archaeopress. Parkes, C. M., Launganim, P., and Young, B. 1997. Death and Bereavement Across Cultures. London; Routledge. Perrot, J. and Ladiray, D. (eds) 1980. Tombes a Ossuaires de la Region Cotiere Palestinienne au IVe Millenaire Avant l’ere Chretienne. Memoires et Travaux du Centre de recherches préhistoriques français de Jérusalem. Vol 1. Paris, Association Paléorient. Petrie, W. M. F. 1907. Giza and Rifeh. London, School of Archaelogy in Egypt andBernard Quaritch. Rabuzzi, K. A. 2005. Home. In L. Johns (ed.), Encyclopaedia of Religion, second edition. Detroit, Macmillan Reference: 4104-4107. http://www.e-reading.club/ bookreader.php/133766/Encyclopedia_of_religion. Radin, P. 1945. The Road of Life and Death. Princeton, Pantheon Books. Ray, B. C. 2005. East African Religion in Northeast Bantu Religion. In L. Johns (ed.). Encyclopaedia of Religion, second edition. Detroit, Macmillan Reference: 2574-79. http://www.e-reading.club/bookreader. php/133766/Encyclopedia_of_religion. Rollefson, G. 2000. Ritual and Social Structure at Neolithic ‘Ain Ghazal. In I. Kuijt (ed.). Life in Neolithic Farming Community: Social Organisation, Identity, and Differentiation. New York: Springer Science and Business Media: 163-190. Ronen, A., Milstein, S., Lamdan, M., Vogel, J. C., Mienis, H. K. and Ilani, S. 2001. Naḥal Re’uel, A Middle PrePottery Neolithic B Site in the Negev, Israel. Quartar 51/52: 115-156. Rosen, S. 2005. Dung in the Desert: Preliminary results of the Negev Holocene Ecology Project. Current Anthropology 46: 317-327. Rothenberg, B., Segal, I. and Khalaily, H. 2004. Late Neolithic and Chalcolithic Copper Smelting at the Yotvata Oasis (south-west Arabah). Institute for Archaeo-Metallurgic Studies 24: 17-28. Sandarupa, S. 1996. Life and Death in the Toraja. Ujung Pandang, Indonesia, Tiga Tauru. Schmandt-Besserat, D. 1998. ‘Ain Ghazal ‘Monumental’ Figures. Bulletin of the American Schools of Oriental Research 310: 1-17. Schmandt-Besserat, D. (ed.) 2013. Symbols at ‘Ain Ghazal. Berlin, Ex Oriente. Schmidt, K. 2000. Göbekli Tepe, Southeastern Turkey. A Preliminary Report on the 1995-1999 Excavations. Paléorient 26:45-54. Schmidt, K. 2010. Göbekli Tepe – the Stone Age Sanctuaries: New results of Ongoing Excavations with a Special Focus on Sculptures and High Reliefs Documenta Praehistorica 37:239-256. Schmidt, K. 2012. Göbekli Tepe: A Stone Age Sanctuary in South-Eastern Anatolia. Translated by M. Wittwar. Berlin, Ex Oriente. Sell, E. 1914. Images and Idols. In J. Hastings (ed.). Encyclopaedia of Religion and Ethics, 7:150-151. Edinburgh, T. & T. Clark.

Shaham, D., and Belfer-Cohen, A. 2003. Incised Slabs from Hayonim Cave: A Methodological Case Study for Reading Natufian Art. In F. Borrell, J. J. Ibanez and M. Molist (eds), Stone Tools in Transition: From Hunter-Gatherers to Farming Socities in the Near East. 7th Conference on PPN Chipprd and Ground Stones Industries of the Fertile Crescent: 407-420. Barcelona, Universitat Autonomia de Barcelona, Servei de Publicacions. Simmons, A. H. 1997. Ecological Changes During the Late Neolithic in Jordan: A Case Study. In H. G. K. Gebel, Z. Kafafi and G. O. Rollefson (eds), The Prehistory of Jordan II. Perspectives from 1997. Studies in Early Near Eastern Production, Subsistence and Environment 4: 309318. Berlin, Ex Oriente Lux. Simmons, A. H. 2007. The Neolithic Revolution in the Near East, Transforming the Human Landscape. Tucson, University of Arizona Press. Simmons, T., Goring-Morris, A. N. and Horwitz, L. K. 2007. What Ceremony Else? Taphonomy and the Treatment of the Dead in the Pre-Pottery Neolithic B Mortuary Complex at Kfar HaḤoresh, Israel. In: M. Faerman, L. K. Horwitz, T. Kahana and U. Zilberman (eds), Faces From the Past: Diachronic Patterns in the Biology and Health Status of Human Populations from the Eastern Mediterranean: 1-27. British Archaeological Reports International Series, 1603  Oxford, Archaeopress Stekelis, M. 1972. The Yarmukian Culture of the Neolithic Period. Jerusalem, Magnes (Hebrew). Vasiliev, A. 1952. The Iconoclastic edict of the Caliph Yazid II, A.D. 721. Dumbarton Oaks Papers 9-10: 25-47. Verhoeven, M. 2007. Losing One’s Head in the Neolithic: On the Interpretation of Headless Figurines. Levant 39:175-183. Walker-Tubb, K. and Grissom, C. 1995. ‘Ain Ghazal: A comparative study of the 1983 and 1985 statuary caches. Studies in History and Archaeology of Jordan 5: 437-47. Waterson, R. 2009. Paths and Rivers: Sa’dan Toraja in Transformation. Leiden, Koninklij Instituut voor Taal, Land en Volkenkunde Press. Watkins, T. 2010. New Light on the Neolithic Revolution in South-West Asia. Antiquity 84:621-34. Wensinck, A. J. 1997. Sura. In C. E. Bosworth (ed,), The Encyclopaedia of Islam. Vol. IX: 889-92. Leiden, E. J. Brill. Willcox, G., Fornite, S. and Herveux, L. 2008. Early Holocene Cultivation Before Domestication in Northern Syria. Vegetation History and Archaeobotany 17: 313-25. Yizraeli-Noy, T. 1999. The Human Figure In Prehistoric Art in the Land of Israel. Jerusalem, Israel Museum and the Israel Exploration Society (Hebrew). Yogev, O. 1983. A Fifth Millennium BC Sanctuary in the ‘Uvda Valley. Qadmoniot 16: 118-22 (Hebrew). Zayadine, F. 2003. The Nabataean Gods and Their Sanctuaries. in, G. Markoe (ed.), Petra Rediscovered: Lost City of the Nabataeans: 57-64. London, Harry N. Abrams. 35

A Basket-Handled Teapot from Fazael 2 Shay Bar

Introduction

holemouth jars and kraters (Garfinkel 1999: Figs. 137, 145, Types D2, F2). Another spouted vessel—the teapot— is considered a common Early Bronze Age type (Amiran 1971: 63) and is found in many sites in the region, but until now it was missing from Chalcolithic assemblages (except for the unprovenanced vessel published by Amiran).

Basket-handled vessels and teapots rarely occur in the Chalcolithic pottery assemblages of the southern Levant. Three basket-handled vessels from a burial chamber in the cemetery at Kissufim Road (Goren 2002: Fig. 4.2: 2, 4-5) are characterized by non-fenestrated pedestals and rich painted decoration including plant motifs and cross-hatched designs. Another baskethandled vessel with a rich painted decoration was found as part of a foundation offering at Shiqmim (Levi and Alon 1987: Figs. 6.11, 12.14 and Goren 2002: 25): its resemblance to the Kissufim Road objects suggests that they all may have been the work of the same artisan.

Fazael 2 The site of Fazael 2 (Israel Grid 2412/6618) was excavated between 2007 and 2018 (Bar 2013, 2014; Bar et al. 2013). Stratum 2 of this three-stratum site was dated to a late stage in the Chalcolithic continuum, with radiometric dates falling within the 1st century of the 4th millennium BC (Bar 2014: 319-320). The main feature discovered in the excavation of Stratum 2 is a very large courtyard house, covering an area of approximately 620 sqm, comprising the following main structures:

Another vessel from a private collection, of unknown provenance (possibly from the Negev), was published by Amiran (Amiran 1986). This vessel was described as a deep globular bowl with a basket handle, a spout and two bird figurines attached to the rim. The morphology of this vessel resembles that of three copper baskethandled bowls in the Nahal Mishmar hoard (Bar-Adon 1980). This resemblance and the notable similarity between Amiran’s bird figurines and copper figurines from the hoard led her to suggest that there was a new type of ritual vessel introduced during the Chalcolithic.

1.

2.

Basket-handled vessels in the northern Levant include five vessels found in Byblos (Dunand 1973: Fig. 160), characterized by their flat bases and incised herringbone decoration. Although there are some similarities between the examples from Byblos and those from the southern Levant, cultural diversity between these distant regions hinders creating a clearcut connection between them. In the Early Bronze Age I the basket-handled vessels became more common (e.g., Azor: Ben-Tor 1975: Figs. 8: 14, 10: 5; et-Tell: Marquet-Krause 1949: Pl. 67; Giv’atayim: Sussman and Ben-Arieh 1966). Most of the basket-handled vessels dating to this period are spouted and labeled as teapots. Spouted vessels appeared as early as in the Pottery Neolithic period (especially the Wadi Rabah phase; Garfinkel 1999: Fig. 68, Type A3), and became common in the Chalcolithic repertoire, mainly in the form of

A square courtyard: 560 sqm in area (28 m x 20 m), bounded by 80-100 cm thick stone walls. Most of the courtyard has not yet been excavated. The remains of several stone-built installations and working areas were found in a few soundings. Two rooms: The first, a broad room (62 sqm; 4 m x 15.5 m), was found abutting the southeastern section of the courtyard. The room was divided into two large cells, and its entrance faced east. At least five successive beaten-earth floors were detected—all abutting the room’s walls, implying a long period of habitation. The second room was excavated in the western part of the courtyard. It was 60 sqm in area (4 m x 15 m), and was divided into two large cells. An entrance flanked by two standing monoliths was set at the southern part of the room. This room was built in the early phase of Stratum 2, and in the later phase of this stratum it went out of use, becoming part of the main courtyard. The basket-handled vessel was found crushed on the floor of this room, close to its entrance (Fig. 1).

The basket-handled teapot The basket-handle vessel from Fazael 2 (Fig. 2) was dated to the 1st century of the 4th millennium BC and 36

S. Bar: A Basket-Handled Teapot from Fazael 2

Figure 1. The location of the crushed basket-handled teapot near an entrance to a room (S. Bar).

designated as a teapot because it has both a handle and a spout. The vessel is handmade of the local clay common to the Fazael pottery assemblage. Its body is in the form of a deep globular bowl, and the diameter of its rim (16.5 cm) is slightly larger than that of the base (12.3 cm). The maximum width of the body is 22.8 cm (27.7 cm including the spout), and its maximum height is 24 cm (18.8 cm without the handle). Its base is flat; its spout, with an average external diameter of 3.6 cm, is attached to the upper part of the body. The basket handle is attached to the uppermost part of the rim, above the spout. The handle is nearly aligned with the spout, giving the vessel the shape of a watering can. Since the handle is not perfectly aligned with the spout, this notable offset renders a somewhat sloppy appearance to the vessel. The vessel is not decorated at all, though the upper part of its body shows signs of a pre-firing smoothing treatment.

stand out. The most notable difference is the position of the basket handle relative to the spout. While the Chalcolithic examples are vertical and aligned with the spout, the Early Bronze Age handles are horizontal. Is the Fazael teapot evidence of a continuity of Chalcolithic pottery traditions into the early phases of the Bronze Age, as suggested by Amiran (Amiran 1986: 84)? This is probably not the case. The very rare appearance of this vessel type in the Chalcolithic period cannot support such an assumption. It seems that the similar traits do not a represent continuity. They seem rather to point to an autonomous tradition developed by the Bronze Age artisans. Another question is whether there is a link between the basket-handled vessels from the Chalcolithic mortuary contexts (Kissufim Road and Shiqmim) and the vessel from Fazael. While the appearance of basket handles is very rare in the Chalcolithic repertoire—that may have indicated a link between these occurrences—there are many differences between them, the most notable of which is the absence of a spout in the mortuary items, suggesting a completely different utilization from the Fazael vessel. Other significant differences between the vessels are the morphology of the base (pedestal in the mortuary items) and the lack of painting on the Fazael (and Amiran) items.

The only good parallel for the Fazael 2 teapot is Amiran’s (Amiran 1986) unprovenanced vessel. Both vessels have a similar body morphology, flat bases and vertical basket handles that are located above the spout and aligned with it, and neither are painted. The differences between them include the bird figurines on the Amiran vessel and the different placement of the handle (which is lower in the Fazael teapot). A comparison between the Fazael specimen and later Early Bronze Age I teapots shows similar traits (flat bases, morphology of the body, low basket-handles and limited decoration). However, some differences

Is this a ritual vessel type, as suggested by Amiran? On the one hand, the very rare appearance supports a non37

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 2. The Fazael 2 basket-handled teapot.

daily use, and the bird figurines on the Amiran item certainly support such an assumption. On the other hand, these were never found in a Chalcolithic cultic context.

Bar, S., Bar Oz, G., Ben Yosef, D., Boaretto, E., RabanGerstel, N. and Winter, H. 2013. Fazael 2, one of the latest Chalcolithic sites in the Jordan Valley? Report of the 2007-2008 excavation seasons. Journal of the Israel Prehistoric Society 43: 148-185. Bar-Adon, P. 1980. The Cave of the Treasure: The Finds from the Caves in Nahal Mishmar. Jerusalem, Israel Exploration Society. Ben-Tor, A. 1975. Two Burial Caves of the Proto-Urban Period at Azor, 1971. Qedem Monographs 1. Jerusalem, The Hebrew University of Jerusalem Institute of Archaeology. Dunand, M. 1973. Fouilles de Byblos, V. Paris, Librairie d’Amerique et d’Orient, A. Maisonne. Garfinkel, Y. 1999. Neolithic and Chalcolithic Pottery of the Southern Levant. Jerusalem, The Hebrew University of Jerusalem Institute of Archaeology. Goren, Y. 2002. The pottery assemblage. In Y. Goren and P. Fabian (eds), Kissufim Road, A Chalcolithic Mortuary Site: 21-41. Jerusalem, Israel Antiquities Authority. Levy, T. E. and Alon, D. 1987. Excavations in Shiqmim Cemetery 3: final report on the 1982 Excavation.

It appears that we must wait until additional spouted vessels are discovered in Chalcolithic contexts to explore this intriguing phenomenon further. References Amiran, R. 1971. The Ancient Pottery of Eretz Yisrael. Jerusalem, The Bialik Institute and the Israel Exploration Society. Amiran, R. 1986. A new type of Chalcolithic ritual vessel and some implications for the Nahal Mishmar Hoard. Bulletin of the American Schools of Oriental Research 262: 83-87. Bar, S. 2013. Yogevim VeNokdim I: Reports on Archaeological Excavations in the Jordan Valley and the Desert Fringe of Samaria (2006-2001). Haifa, Seker. Bar, S. 2014. The Dawn of the Bronze Age. Leiden and Boston, Brill. 38

S. Bar: A Basket-Handled Teapot from Fazael 2 In T. E. Levy (ed.), Shiqmim I: Studies Concerning Chalcolithic Societies in the Northern Negev Desert, Israel: 335-355. Oxford, British Archaeological Reports International Series 356.

Marquet-Krause, J. 1949. Les Fouilles de ‘Ay (et-Tell) 19331935. Paris, Geuthner. Sussman, V. and Ben-Arieh, S. 1966. Ancient burials in Giv’atayim. ‘Atiqot (Hebrew Series) 3: 27-39.

39

The Contributions of Early French Scholars to Levantine Prehistory Ofer Bar-Yosef and François Valla

Opening remarks

Nineteenth Century Pioneers

For many centuries, the Holy Land attracted generations of travelers, pilgrims, clergymen and others. They came to this country searching for places mentioned in the Biblical stories or recorded in the New Testament. A minor change but with far reaching implications occurred in early 19th Century when prehistoric sites were discovered in Western Europe. The knowledge that many such sites are embedded in river gravels, exposed in quarries or caves providing a record of past human presence motivated research in other regions of the Old World. These revelations attracted scholars who were either curious investigators or sent on mission by various institutions, to conduct similar surveys and research in the Holy Land. Those who came to look for similar finds in the Levant were already familiar with the prehistoric stone tools uncovered at the time mainly in French sites. Their studies during the late 19th and early 20th centuries laid the foundations of our current knowledge. This paper is not intended as a complete history of all those who made contributions to the study of Levantine prehistory. We felt that there is a need today to tell the story of French researchers in the Levant, briefly describe the motivation of the 19th century clergypersons, who searched for evidence for events and places mentioned in the Bible, and summarize their activities without ignoring the presence of others. The scholars involved were mainly British and Germans, as well as a few Americans with similar hopes, and scholars from other countries, who conducted their research at the same time. Many of the sites that were recorded, or even surface collections, in the course of these explorations, became targets of systematic excavations in the 20th century. Many sites disappeared due to the rapid developments of urban centers, establishment of new villages, building roads, etc. that took place at the end of the Ottoman days and especially after the First World War. W. F. Albright, the father of Biblical Archaeology, summarized the early discoveries in his 1949 volume ‘The Archaeology of Palestine’. Undoubtedly, a fuller record of the history of prehistoric research in this region is needed. However, in the 21st century when English is the dominant language of the academic world, we believe that it is most appropriate to pay attention to the pioneers who spoke and reported in French.

The Nineteenth century witnessed the development of geology, human biology and evolution. The archaeology of the prehistoric periods became the tool to study the past prior to the events mentioned in the Bible. However, this mission could not be achieved before treating intertwined issues. The first was how to fit the new discoveries into the short time scale held by biblical scholars and generally accepted that established the creation of the world some 6000 years ago. The second issue was the relationship between the new findings and the biblical record. It took many years of field research and of scholarly debates in order to get over these difficulties. Indeed, it was not before the 1930’s that Prehistory gained the full recognition of its longer time scale, and its independence vis a vis Biblical Archaeology. Ironically, this development resulted at least partly from the work of faithful French Catholics. During the second half of the nineteenth century, France witnessed a Catholic revival. In this context, the Holy Land attracted a growing interest. Believers were encouraged to visit Palestine in organized pilgrimages. Missionaries were sent to sustain and develop local Christian communities, and they built schools and hospitals. Scholars undertook a thorough exploration of the Land of the Bible to get a better understanding of the Scriptures with apologetic motivations. Main actors in this process were priests belonging to a variety of Congregations: Dominicans, Augustins of the Assumption (Assumptionists), Jesuits and many others. Generally speaking, in spite of intermittent conflicts between anticlerical republicans and the Church at the turn of the century, the French state supported the move for geopolitical reasons. The travelers, who came mainly from Western Europe and later from the USA looking for historical antiquities, recorded prehistoric sites and surface finds, employing European terminology. Quite often, these early researchers landed in Beirut that was the main harbor during this time. Among them were C. L. Irby and J. Mangles, English scholars, who were the first in 1817-1818 to note the presence of dolmens in Trans-Jordan. Another example was the famous French Mesopotamian researcher E. Botta (1802-1870), 40

O. Bar-Yosef and F. Valla: The Contributions of Early French Scholars to Levantine Prehistory a diplomat, and his travel companion, Dr. Hedenborg from Sweden. On their way to Tripoli in 1838, they identified prehistoric breccia in Wadi Antelias (north of Beirut) and in Ras-el-Kelb.

John Lubbock (1865). Arcelin stressed the importance of vegetal food in the human diet as evidenced by the scarcity of animal bones. Later, Moretain’s collection, like many others, was shipped to France and distributed among various museums.

Here we need to stress that the concept of Prehistory emerged only during these early years. P. Tournai (1833), a French geologist was the pioneer who suggested calling the period that preceded the biblical floods période antédiluvienne (pre-flood period). However, when visiting Egypt he saw no evidence in Thebes for the floods, and he changed the term to be antéhistorique (‘pre-historic’; Chippindale 1989). The term ‘prehistory’ in English was actually suggested by D. Wilson in 1851.

Among the prominent 19th century scholars was Father Vincent J. J. Germer-Durand (1845-1917), an Assumptionist, who was perhaps the most systematic and enthusiastic researcher in the region (Fig. 1). In the environs of Jerusalem, he collected thousands of stone artifacts. The major localities were the valley of Baqa’a-Rephaim in Jerusalem that drains westward, Zur-Bahr and Artas near Bethlehem. His collections were kept for many years in the Hostelry Notre Dame de France in Jerusalem. Durand built this hostelry in order to host the pilgrims, and there he opened a museum in 1897. In January 1896, Durand summarized the prehistory of the country in a public lecture in the Ecole Biblique de St. Etienne in Jerusalem. His conclusions were that prehistoric humans were able to occupy only the hilly areas and the Trans-Jordanian plateau because the waters of the Mediterranean Sea covered the Jordan Valley and part of the coastal plain. He attributed the bifaces he collected in Baqa’aRephaim to the oldest known lithic industry at this

Among the first French investigators was Victor Guérin (1821-1890). From 1852 he served at the French School of Athens, and he made frequent trips to Turkey, Egypt and the Levant until 1888. His seven volumes on the geographical history of Palestine were published from 1854 to 1888. These volumes summarize his field records concerning many sites and ancient ruins that he identified employing name places from the Bible. This project received the prize of the French Academy of Sciences. Second was F. de Saulcy (1807-1880), who in 1850-1851 and in 1863 excavated in numerous places and sent his finds to the Louvre. Perhaps the most interesting locality of his operation was what he called the Tombs of the Kings, a megalithic site north of Jerusalem that many years later was recorded in detail by M. Stekelis. The French geologist L. Lartet (1840-1899) visited the same sites in 1864-1865. It was he who later excavated the rockshelter of Cro-Magnon where he uncovered the remains of five skeletons of modern humans in 1868. Lartet was part of a multidisciplinary expedition organized by the Duke H. de Luynes in order to study the Dead Sea and its surroundings (Luynes 1871). Lartet was the first to identify the remains of the Pleistocene Lake Lisan in the Jordan Valley. He conducted some soundings in the prehistoric cave of Jiita (Lebanon) and noted breccia with artifacts and bones near Kfar Kana in the Lower Galilee (Lartet 1869). During the 1890’s Father J. Moretain, who lived in Beit Sahour, a village near Bethlehem (Buzy 1928) made a series of surface collections. His finds included numerous axes, serrated blades, arrowheads, knives and even a bone point. Moretain first described these objects in a lecture delivered to the International Congress for Anthropology and Prehistory in Paris. The first report by P. Cazalis de Fondouce, an eminent French archaeologist, attributed these artifacts to a historic period. However, A. Arcelin (1838-1904) corrected the mistake by assigning this collection to the Neolithic period, using the recently designed terminology of Sir

Figure 1. Father Germer-Durand.

41

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead time, the ‘Chellean’ as identified in the Somme Valley gravels, later incorporated into the Acheulian. He also conducted surface collecting in Trans-Jordan, and he noted a Neolithic site near Jaffa. The geomorphological reconstruction was based on the work of E. Hull (18291917). The latter served as the geologist of the Survey of Western Palestine (Hull 1885, 1889).

Jesuit St. Joseph University in Beirut. He was the first to excavate a prehistoric site in the Levant when he conducted in 1890 a series of excavations in the caves near Adlun (Zumoffen 1900). His book Phoenicia before the Phoenicians is also based on his surveys between Adlun, south of Beirut and Nahr-Ibrahim cave north of the city. Since his basic academic training was in geology, he also published a book on the geology of Lebanon (Zumoffen 1926). Zumoffen also cooperated with Father P. Bovier-Lapierre who taught in Egypt. Moreover, Zumoffen replaced Bovier-Lapierre in Beirut from 1915 to 1942. Bovier-Lapierre surveyed some areas in Lebanon, the Galilee, and Mt. Carmel. He cooperated with Father A. Mallon, another Jesuit, who was influenced by him and became interested in prehistory from 1919 onward. The cooperation between the two led to the publication of the first systematic list of sites by periods (Mallon 1925). The list of sites in the 1925 paper divided the known finds according to relative chronology on the basis of typology. It was arranged as follows: Paleolithic I with handaxes, Paleolithic II without handaxes, Mesolithic – including sites with adzes, Neolithic – characterized by polished axes and blades with parallel edges and finally the Eneolithic – known today as Chalcolithic.

At the same time Father Germer-Durand justified the collections of old objects by writing the following: ‘The study of this civilization when flint played a major role, was divided into two periods. One we call Paleolithic and the second Neolithic… In reality, the Neolithic period merged into History or at least with the Proto-History… At the time of Abraham, the inhabitant of Palestine did not know other tools, or other weaponry beside these rocks and the Hebrew people used such knives and arms of stone during the time of Joshua and Saul’ (Germer-Durand 1887; our translation). It indicates the difficulties that the clergy had in reconciling the discoveries from the Stone Age with chronology based on the Bible. In this debate Father Germer-Durand was a conservative, who held that the Acheulian material could not be older than 6000 years. The increasing research activities in the Holy Land did not escape the attention of G. de Mortillet, the first editor of Matériaux pour l’histoire primitive et naturelle de l’homme, a journal dedicated to prehistoric archaeology and ethnography. In his book (1883), he mentions the Moretain collection, and he cites the ‘Chellean’ artifacts near Beit Sahour and in the vicinity of Tiberias collected by Abbé Richard.

Father Alexis Mallon (1875-1934; Fig. 2) was an influential figure in creating the conditions for the second phase of research, the excavations of selected sites. In his early career, he wrote the first published Grammar of the Coptic language (1904). He was sent by his Order to Jerusalem in 1913 with the mission to found a Pontifical Biblical Institute affiliated to the one just created in Rome. However, he spent the troubled years of the First World War in Cairo, and he came back to Jerusalem only in 1919. His original mission was

Another active French researcher was Father G. Zumoffen (1848-1928), who served as a teacher in the

Figure 2. Father A. Mallon.

42

O. Bar-Yosef and F. Valla: The Contributions of Early French Scholars to Levantine Prehistory accomplished in 1927 when he laid the foundations of the Pontifical Biblical Institute in Jerusalem.

benefitted immensely from the work carried out by Father Alexis Mallon mentioned above. An update of the Mallon list was published by Neuville (1929).

The growing interest in prehistoric research led Father Mallon to join R. Neuville in the excavations at Oumm Qatafa in 1928, the same year he also did a sounding in Shuqbah cave. On the basis of his finds, he offered D. Garrod who arrived that year in Jerusalem after her excavations in Gibraltar, to dig the site. In his enthusiastic energy Father Mallon discovered Teleilat el-Ghassul in 1929, and he started digging there with the hope to find the evidence for the destruction of Sodom and Gomorrah. He dated Teleilat el-Ghassul to 2000 BCE. He conducted five field seasons with his colleague Father R. Koeppel before his untimely death in 1934. Father Koeppel continued the field work until 1938. Upon completing the excavations, Koeppel published the site report to which R. Neuville (see below) contributed the chapter on the lithics.

René Neuville (1899-1952) René Neuville was born in 1899 in Gibraltar where his father was the French consul. René started his own diplomatic service at an early age, first in Vintimille in Italy (Fig. 3a). In 1926 he was appointed as Chancelier du Consulat de France in Jerusalem where he served until 1937. His early studies of Egyptian epigraphy led him to archaeology while A. Mallon introduced him to prehistory upon his arrival in Jerusalem. Following the discovery of the Zuttiyeh skull by F. Turville-Petre in 1925 Neuville had started his own research in the country. A devout Catholic, he worked for the French Consulate in Jerusalem, an institution traditionally considered as the protector of the Holy Places. Consequently, he had close ties with Father A. Mallon and with Fathers D. Buzy and P. Duvigneau, two members of the Congregation des Prêtres du Sacré Cœur de Betharam, who were based at the Carmelite monastery in Bethlehem. Due to their connections with the local Bedouins, Neuville became interested in the sites along Wadi Khareitoun that descends to the Dead Sea. The Bedouin who lived in the Judean Desert emptied or cleaned out small caves and rockshelters in order to turn them into storage facilities, dwelling places, or modified goat pens. Artifacts, bone tools

Four political entities – Palestine, Lebanon, Syria and Jordan – were created after the First World War, when the region was controlled by Britain and France. However, the new boundaries did not prevent scholars who worked in Palestine to cooperate with their colleagues in Syria and Lebanon, or visit their excavations. Although the first excavations in Lebanon preceded those in Palestine, additional striking discoveries were yet to come from the sites of Wadi Antelias. First, a small cave known as Abri Bergy produced Upper Paleolithic assemblages. It was particularly, the excavations of Ksar ‘Akil rock-shelter by A. Day from the American University of Beirut and later by Jesuit Fathers G. Doherty and F. Ewing from Boston College that exposed the deepest sequence of Upper Paleolithic layers in the entire Levant with a few Mousterian layers at the base (Ewing 1947). Not less important for understanding the Levant as the crossroads between Africa and Asia are the excavation of Bovier-Lapierre. He carried out the dig in the Lower Paleolithic site in Abassieh, today a suburb of Cairo, prior to the First World War. Currently it is the northeast African station of makers of Lower Paleolithic stone tools closest to the Levant (Bovier-Lapierre 1926). The years after the end of the First World War and prior to the Second World War were the ‘Golden Years’ of Palestinian archaeology of all periods. Several newcomers laid the foundations to the prehistory of the region including F. Turville-Petre (Bar-Yosef and Callendar 1997) and D. Garrod, who came from Britain (Bar-Yosef and Callander 2004), R. Neuville from France and M. Stekelis from Russia (Bar-Yosef and Callander 1997). As the work of the first two researchers and the last one were described elsewhere, the following summary refers only to R. Neuville and his associates. These four young scholars, in their mid-thirties,

Figure 3a. R. Neuville when incorporated as a diplomat.

43

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead and occasional figurines, which they found, were sold to the monasteries near Bethlehem. In the Carmelite monastery, Neuville identified an erotic figurine, found, according to his informant, in the small cave of Ain Sakhri in Wadi Khareitoun (Neuville 1929). In the following visit to the site, Neuville sieved the dumps left by the Bedouin, and he uncovered a Natufian assemblage.

collections including the bifaces from Baka’a – Rephaim, and he recommended further excavations in this locality. This wish was later fulfilled by Stekelis, who published his findings in 1948. Neuville also recognized the importance of the sites that Father Buzy reported from his surface collections in Wadi Tahouneh. The lithic industry was then labelled ‘Tahounian’, and it was attributed to the Mesolithic (Buzy 1928). It is dated today to the period known as ‘Pre Pottery Neolithic B’.

In addition, Neuville was fortunate to discover by himself in his excavations at Umm ez-Zuweitina cave in Wadi Djaher (Nahal Tekoa) the broken Natufian figurine of a ‘kneeling gazelle’ (Neuville 1926). Since then these two figurines became renowned as representing the mobile art of the Natufian culture, although the archaeological context of the first one remains uncertain.

The close relationship between R. Neuville and Father A. Mallon, led the former to write the report on the lithics from this site. Continuing his interest in the lithics of later periods, Neuville described Bronze Age lithics (Neuville 1934b), including the assemblage from Minet Dalhia (in Lebanon). That site was originally considered by R. Desribes as ‘Solutrean’. However, Neuville correctly assigned it to the ‘Enéolithique’ (or early Chalcolithic; Neuville 1933; Cauvin 1968).

Father D. Buzy (1883-1965) was interested, like other clergypersons in finding the traces of biblical events. In search of tracing the Israelites journey during the Exodus through the Sinai desert, he initiated a trip with R. Neuville to Kadesh Barnea (near Qseime, in Wadi Qudeirat). There, he collected a surface scatter of Upper Paleolithic artifacts published by him as ‘Magdalenian’ following the French terminology (Buzy 1929). While travelling by car from Jerusalem to Sinai he noted the potential of the Negev for prehistoric research, stating that, ‘as a matter of fact, the whole area of the Negev from Beer-Sheva to el-Qseimeh deserves a special study from a prehistoric point of view. As our car took us smoothly along the German-Turkish road, recently restored by the government of Palestine, we saw tens and hundreds of tracks on the fields and hills, covered by flints glinting in the sun, lighting our way. One could say that this whole region, measuring some few hundreds of kilometers, is an enormous lithic station. We are pointing it out for eventual researchers’. (Buzy 1929:364-365, our translation).

During 1928-1932 Neuville excavated in Wadi Khareitoun the rockshelter of Erq el-Ahmar (the ‘red cliff ’) where he identified several Upper Paleolithic phases. Many years later I. Gilead (1981) employed the name ‘Ahmarian’ for the particular Upper Paleolithic bladey assemblage discovered in this site. This industry is accepted today as evidence for the expansion of modern humans into Europe. Smaller excavations in the same area by R. Neuville included Umm Qala’a with Natufian remains and the Mousterian sites of Tor Abu Sif, Sahba, and Ghar, and etTabban. There he noted the transition from the Middle to the Upper Paleolithic. However, his largest project was the cave of Umm Qatafa, which mostly featured Acheulian assemblages. At the time, Neuville was unable to finish the excavations, and he completed them after the Second World War when he was appointed as the French consul in Jerusalem. Among his other works was a small, less well-known excavation, in the dark karstic cave called Teomim (‘Twins’ in Hebrew), on the western slopes of the Judean hills.

The information and artifacts in the collections of the Bethlehem monastery motivated Neuville, whose work was always supported by A. Mallon, to initiate an excavation in Umm Qatafa cave and the other sites in the Judean Desert. During these years, Neuville established contacts with M. Stekelis, who later excavated with him the Erq el-Ahmar and Qafzeh caves. Neuville also worked with L. Picard, M. Avnimelech, Y. Ben-Tor, and G. Haas from the Hebrew University, founded in 1925. He also made contact with the Institut de Paléontologie Humaine in Paris where he met the major French scholars of the times such as M. Boule (1861-1942), Abbé H. Breuil (1877-1961) and R. Vaufrey (1890-1967). These scholars provided Neuville with scientific recognition and financial funding.

While visiting Nazareth, members of the clergy drew Neuville’s attention to the cave in Jebel Qafzeh. Together with M. Stekelis, he excavated there from 1933-1935. The human remains uncovered in the Mousterian layers of this cave became immediately famous, but they remained inadequately published until a later date. They were noted by Boule and Vallois (1946) and by other paleoanthropologists. However, only with the renewed excavations in Qafzeh cave by B. Vandermeersch were the skeletal remains studied in detail (Vallois and Vandermeersch 1972; Vandermeersch 1981; Tillier 1999).

This led to the important visit in Palestine in 1933 of Abbé H. Breuil, a leading figure in world prehistory for several decades. Breuil examined the available

Neuville’s activities were interrupted by the outbreak of the Second World War. He spent these years studying 44

O. Bar-Yosef and F. Valla: The Contributions of Early French Scholars to Levantine Prehistory Vandermeersch 1972), and a fuller revision including the old collections was done later (Bar-Yosef and BelferCohen 2004). The two skulls saw the light recently (Vandermersch et al. 2013). Neuville passed away in 1952 shortly after his book (Neuville 1951) was published in Paris. Following the news of his death, the Bedouin who worked with him excavated two small caves in the northern Judean Desert, Mazraq en-Naj and el Quseir, and they smuggled the material to Jerusalem, crossing the border between the states of Jordan and Israel created in 1948 (Perrot 1952). Discussion The research conducted by the 19th century pioneers facilitated the decisions as to where to start digging in 1925, first by F. Turville-Petre, then by R. Neuville and D. Garrod. The survey reports that were produced, especially those summarized by A. Mallon (1925) and Neuville (1929), followed the European terminology of the prehistoric periods. Whether small surface collections of open-air localities or cave sites, they were ordered according to the sequence from the Paleolithic to the Neolithic and Eneolithic. The ‘Golden Years’ of archaeology between the two world wars provided ideal conditions for field research despite few good roads and poor living conditions. The knowledge and insights of Neuville and Garrod are clearly expressed in their publications. However, their first task was to build a periodical and cultural sequence based on excavated sites. The lack of radiometric dates meant that they had to use the relative chronology of stratigraphy and its geological context as provided by geologists and in particular by Leo Picard. Picard was a professor of geology in the Hebrew University, who came to the country from Switzerland in order to look for water sources for the new Jewish settlements in the Jezreel and Jordan Valleys. Being trained in the Alpine glacial chronology employed by Quaternary geologists, he recognized three major Pluvial periods in the country. This scheme, in addition to the paleoecological reconstruction based on the known behavior of the hunted species, served as support for the relative chronology and paleoclimatological reconstructions.

Figure 3b. R. Neuville in Jerusalem as the French Consul in Israel.

the early prehistory of Morocco with A. Ruhlman, and he came back as the French consul to Jerusalem after the war (Fig. 3b). He was able to renew his excavations at Umm Qatafa and reach the bedrock. He entrusted the analysis of his collections from the excavations at the Terrace of El-Khiam to J. Perrot who had just arrived in Israel. The report was included in the book of Neuville that was published in Paris in 1951. Unfortunately, Neuville did not manage to analyze the lithic collections from Qafzeh cave. Only a brief chapter, mostly about the stratigraphy of the site, was included in his 1951 volume. The original Qafzeh lithic assemblages were not published by M. Stekelis, who ‘inherited’ the collections after Neuville’s untimely death in 1952. The main part of the collections with the bulk of the debitage, following the division made by the Department of Antiquity in 1936, remained in the storage of the Rockefeller Museum erected in 1938. The Middle Paleolithic assemblages were later studied in 1968 by B. Vandermeersch and O. Bar-Yosef, who did not publish their results. Better luck faced the material recovered during the new excavations conducted by B. Vandermeersch since 1965. These were analyzed and published in detail (Hovers 2009).

It is interesting to compare briefly the first chronological scheme offered by Neuville in his 1934 paper with Garrod’s chronology published in her book on the Mount Carmel excavations (Garrod and Bate 1937), and then examine the changes Neuville introduced in his 1951 volume. In 1934, Neuville recognized that the evolution of the stone industries began with two parallel traditions in making artifacts. One tradition is represented by bifaces, namely the Chellean –Acheulian in Umm Qatafa. The other tradition, a core and flake industry began with the Tayacian, and from the base

The Upper Paleolithic layers were not published by R. Neuville, nor were the two skulls attributed to this period. The first publication of the Upper Paleolithic lithics was based on the new excavation in the cave conducted by B. Vandermeersch since 1965 (Ronen and 45

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead of the latter site, it evolved into the Levallois tradition. Then came the typical Mousterian and the transition to the Upper Paleolithic, within which Neuville designated six phases. The Mesolithic included the Kebaran, hardly known at the time, which preceded the Natufian, subdivided into four phases, from Natufian I to IV. This was followed by parallel developments of the Tahounian (I and II) with the Ghassulian I and II ending with the Canaanian (early Bronze) stone tools and pottery.

Upper Paleolithic or the Epi-Paleolithic. In addition, they provided the first reference to cultures as entities within a given period. The pioneering studies and technological determinations influenced the following generations from the 1950’s onwards. Terms in use today such as Natufian or Kebaran were first coined during the pioneering days. So also were technotypological definitions such as ‘Levalloiso-Mousterian’. The latter designation was based on the subdivisions in France where, broadly speaking, the Middle Paleolithic was called Levalloisian in the northern part and in the southern region, Mousterian.

In his 1951 volume (Table 2, page 263) Neuville reiterated what was also known in Western Europe as the co-existence of the Tayacian (or Clactonian) and the Acheulian. Within the latter phase he recognized the late Acheulian named Micoquian (as did Garrod) when he noted also the first appearance of the early Levalloisian. The Levalloisian evolved into Mousterian, and then came the Upper Paleolithic with six phases, followed by the Natufian and the Tahounian (Neuville 1951, p. 178). We should note that his Paleolithic chronology took into account the Mediterranean Sea levels that were studied in Lebanon, the results of the excavations in Mt Carmel caves, already available in Garrod and Bate volume (1937), as well as the results published by A. Rust (1950) from his pre-war excavations at the rock-shelters of Yabrud (Syria). In so doing, Neuville expressed his views about the relationship between Palestinian prehistory and that of Western Europe. However, he refused to push forward any conclusions concerning the rate of relative evolution of one region against the other (Neuville 1951, p. 262).

Both Garrod and Neuville recognized that most of Middle Paleolithic assemblages in Palestine resemble both industries, and they coined the term LevalloisoMousterian. The impact of F. Bordes terminological approach was to reserve the term ‘Levalloisian’ for the technique employed in making several of the Middle Paleolithic assemblages, and to employ the term ‘Mousterian’ for the generally contemporary different cultures of the Middle Paleolithic. In recent years, we note a tendency to use periodic terms to replace cultural designations. The Pre-pottery Neolithic A and Pre Pottery Neolithic B periods become general cultural terms in spite of the many regional differences. Other scholars refer to Epi-paleolithic as a cultural term divided into three subperiods ignoring cultural terms such as Kebaran, Nebekian and others. Research tradition in the study of European Paleolithic demonstrates that a ‘prehistoric culture’ with its own material attributes (artifacts made of stone, bone, antler and more) has a particular territory and chronological boundaries. Each of these cultures deserves a special name such as Aurignacian, Solutrean, Badegoulian and more. This procedure was introduced into the Levant by the pioneer prehistorians when they considered that the amount of information justified a special name. This procedure continues in recent publications (e.g., Garrard and Byrd 2013). The approach of employing periodic terms as equivalent to cultural names is a failure to note that by recognizing ‘prehistoric cultures’ we distance ourselves from the makers of the stone tools. Apparently, this means that in spite of over 100 years of prehistoric research and modern use of advanced dating techniques, members of our community are in the same place where the pioneers of the 19th century stood. However, an indepth discussion of this issue from a wider historical view is beyond the scope of this paper.

We should note that none of these chronological schemes could be built without deep soundings in the caves, or potential correlation with sea levels. In reviewing the excavations of that period, one can see the advantages of digging down quite fast to the bedrock of each site in comparison with the slow, careful pace of today. Conducting this process with hired workers and often without a grid, Neuville and Garrod were able to expose many layers (e.g., in Qafzeh, Umm Qatafa, Tabun, el-Wad). As the dig proceeded through the unexcavated area, they were able to correct some of the obvious mistakes in attributing the finds to the observed different layers. This is evident in the excavations of Qafzeh (Bar-Yosef and Belfer-Cohen 2004) where the system of dividing the cave into sections without a common grid had certain advantages. Another example is Garrod’s response to Bordes’ review of her excavations at Tabun cave (Garrod 1956) where she tried to indicate by publishing the sections of the different arbitrary spits, how many Amudian assemblages were within the complex of Tabun layer E.

Acknowledgements We thought that this paper is appropriate for celebrating I. Gilead’s career and contribution to the studies of the Upper Paleolithic of the southern Levant, as well as other periods and that his naming ‘Ahmarian’

In sum, we need to remark that terminologies brought by the pioneers from Europe still heavily influence our current use whether it is for the Lower, Middle and 46

O. Bar-Yosef and F. Valla: The Contributions of Early French Scholars to Levantine Prehistory industry commemorates R. Neuville excavations in Erq el-Ahmar. O. Bar-Yosef would like to acknowledge the notes on the history of prehistoric research received many years ago from the archives collected by his colleague, the late D. Gilead who passed away in 1974. We are grateful to A. Belfer-Cohen and M. Gruber for comments and editorial assistance. All shortcomings are ours.

Hull, E. 1886. The Survey of Western Palestine: Memoir on the Physical Geology and Geography of Arabia Petraea. London, Adelphi, Publishers: The Committee of the Palestine Exploration Fund. Lartet L. 1869. Suivi de Propositions de géologie et de botanique données par la faculté. In V. Masson et fils (eds), Essai sur la géologie de la Palestine et des contrées avoisinantes: telles que l’Égypte et l’Arabie, comprenant les observations recueillies dans le cours de l’expédition du duc de Luynes à la Mer Morte: 292. Paris, E. Martinet. Luynes, H. (Duc de). 1871. Voyage d’exploration à la Mer Morte, à Petra et sur la rive gauche du Jourdain; oeuvre posthume publiée par ses petits-fils sous la direction de M. le comte de Vogüe. Paris. A. Bertrand. Neuville, R. 1929. Additions à la Liste des Stations Préhistoriques de Palestine et Transjordanie. Journal of the Palestine Oriental Society 9: 114-121. Neuville, R. 1930. Notes de Préhistoire Palestinienne. Journal of the Palestine Exploration Society 10: 193221. Neuville, R. 1931. L’Acheuléen supérieure de la grotte d’Oumm Qatafa (Palestine). L’Anthropologie 41: 13-51, 249. Neuville, R. 1933. Statuette érotique du désert de Judée. L’Anthropologie XLIII: 558-560. Neuville, R. 1934. Le Préhistorique de Palestine. Revue Biblique 43: 237-259. Neuville, R. 1934. Objets en silex. In A. Mallon, R. Koeppel and R. Neuville (eds), Teleilat Ghassul: 55-65. Rome, Pontifical Biblical Institute Press. Neuville, R. 1951. Le Paléolithique et le Mésolithique de Désert de Judée. Paris, Masson. Ronen, A. and Vandermeersch, B. 1972. The Upper Paleolithic sequence in the Cave of Qafzeh (Israel). Quaternaria XVI: 189-202. Rust, A. 1950. Die Hölenfunde von Yabrud (Syrien). OffaBücher 8. Neuenmünster, Karl Wachhöltz. Tillier A-M. 1999. Les enfants moustériens de Qafezh, interprétation phylogénétique et paléoanthropologique. Cahiers de Paléoanthropologie. Paris. Le Centre national de la recherche scientifique Editions. Tournal, P. 1833. Considérations générales sur le phénomène des cavernes à ossements. Annales de Chimie et de Physique 55:161-181. Vallois, H. V. and B. Vandermeersch 1972. Le crâne moustérien de Qafzeh (Homo VI). L’Anthropologie. 76: 71-96. Vandermeersch, B. 1981. Les hommes fossiles de Qafzeh (Israël). Cahiers de Paléoanthropologie. 8: 455. Vandermeersch, B., Arensburg, B., Bar-Yosef, O. and Belfer-Cohen, A. 2013. Upper Paleolithic human remains from Qafzeh Cave. Journal of the Israel Prehistoric Society (Mitekufat Haeven) 43: 7-21. Zumoffen, G. 1900. La Phénicie avant les phéniciens: l’âge de la pierre. Beyrouth, Imprimerie Catholique. Zumoffen, G. 1926. Géologie du Liban. Paris, H. Barrère.

References Bar-Yosef, O. and Belfer-Cohen, A. 2004. The Qafzeh Upper Paleolithic Assemblages: 70 Years Later. Eurasian Prehistory 2: 145-180. Bar-Yosef, O. and Callander, J. 1997. A forgotten archaeologist: The life of Francis Turville-Petre. Palestine Exploration Quarterly 129: 2-18. Bar-Yosef, O. and J. Callander 2004. Dorothy Annie Elizabeth Garrod 1892-1968. In G. M. Cohen and M. S. Jukowsky (eds), Breaking Ground: Pioneer Women Archaeologists: 380-424. Ann Arbor, University of Michigan Press. Boule, M. and Vallois, H. V. 1946. Les hommes fossiles: éléments de Paléontologie humaine. XII588p. Paris, Masson. Buzy, D. 1927. Les Stations Lithiques d’el Qeseimeh. Revue Biblique 36: 90-92. Buzy, D. 1928. Une Industrie Mésolithique de Palestine. Revue Biblique 37: 558-578. Buzy, D. 1929. Une Station Magdalénian dans le Négev ‘Ain el-Qudeirat. Revue Biblique 38: 364-381. Callander, J. and Bar-Yosef, O. 2000. Saving Mount Carmel Caves: A Cautionary Tale for Archaeology in Our Time. Palestine Exploration Quarterly 94-112. Chippindate, C. 1989. The invention of words for the idea of prehistory. Proceedings of the Prehistoric Society 54:303-314. Day, A. E. 1926. The Rock Shelter of Ksar Akil near the Cave of Antilias. Palestine Exploration Fund 158-160. Ewing, J. F. 1947. Preliminary Note on the Excavations at the Palaeolithic site of Ksar ‘Akil, Republic of Lebanon. Antiquity 21: 186-196. Garrard, A. N. and Byrd, F. B. 2013. Beyond the Fertile Crescent: Late Palaeolithic and Neolithic Communities in the Jordanian Steppe. Oxford, Oxbow. VIII-430 p. Garrod, D. A. E. 1956. Acheuléo-jabrudien et préAurignacien de la Grotte de Taboun, Mont Carmel: Etude stratigraphique et chronologique. Quaternaria 3: 39-59. Garrod, D. A. E. and Bate, D. M. A. 1937. The Stone Age of Mount Carmel, Excavations at the Wadi Mughara, 1. Oxford, Clarendon Press. Germer-Durand, J. 1897. L’âge de la pierre en Palestine. Revue Biblique 439-449. Germer-Durand, J. 1881. Un musée palestinien. Paris: Maison de la Bonne Presse 32. Hovers, E. 2009. The lithic assemblages of Qafzeh Cave. New York, Oxford University Press. 47

The Riddle of the ‘Aurignacian’ in the Negev: The Lithic Assemblage from Nahal Neqarot in the Central Negev, Israel Anna Belfer-Cohen and A. Nigel Goring-Morris To Itzik, with thanks for the friendship, sharing, laughter, and great memories over the years

atop Har Harif as such an ‘admixture’ representing two separate assemblages, one Aurignacian (K9A) and the other Epipalaeolithic (K9), based on raw material types and typological considerations.2

Introduction1 Excavations at Nahal Neqarot rock-shelter (NQR) in the central Negev were conducted with a view to investigate the initial interpretation upon discovery that the site contained both in situ Upper Palaeolithic and Epipalaeolithic components (Belfer-Cohen et al. 1991, 1993). The rock-shelter appeared to provide a unique opportunity to investigate contentious issues connected with the nature of the Upper Palaeolithic cultural sequence in the more arid margins of the Southern Levant (Belfer-Cohen 1996; Gilead 1981a, 1981b; Goring-Morris 1987; Marks 1977, 1981).

With this in mind, Nahal Neqarot rock-shelter appeared to be an excellent test candidate, since it was then the only known rock-shelter occupation in the Negev with significant in situ sediments. The test excavations detailed herein revealed that the 60 cm thick sequence of cultural deposits was exclusively Epipalaeolithic, comprising mostly Middle Epipalaeolithic ‘Ramonian’ occupations (Belfer-Cohen 1996; Belfer-Cohen et al. 1991; Goring-Morris et al. 1998).

At the time it was assumed that the Levantine Aurignacian findings reported from and defined in the Mediterranean regions of the Levant (e.g. Belfer-Cohen and Bar-Yosef 1981; Copeland 1975) were present also in the desertic regions of the Negev and Sinai. This assumption was based on reports of sites claimed to reflect the presence of two distinct industries; the microlithic component of the retrieved assemblage was assigned to the local Middle Epipalaeolithic, while the macrolithic component with frontally flat carinated scrapers as well as scrapers on ‘Aurignacian’ blades was used as evidence for the existence of an Upper Palaeolithic local ‘Aurignacian’, similar to the Aurignacian assemblages in the north. Indeed, while Noy (Yizraeli 1967) presented Ramat Matred M190 and M141 as surface sites of Epipalaeolithic attribution containing both microlithic and macrolithic components, Bar-Yosef (1970) considered those assemblages as admixtures of Epipalaeolithic and Upper Palaeolithic industries. Subsequently, Marks and his associates (Larson and Marks 1977: 173; Marks and Simmons 1977) defined a surface concentration

Here we present techno-typological details of the Ramonian chipped stone assemblage at Nahal Neqarot rock-shelter (NQR), followed by a discussion as to its character and its contribution to our understanding of prehistoric developments in the Negev. Stratigraphy The rock-shelter, at 640 m asl, is strategically located on one of the uppermost, northern tributaries of Nahal Neqarot, close to one of the main passes north into the nearby Maktesh Ramon (Fig. 1). Access to the west and south is also readily available by way of the Nahal Neqarot, Nahal Oded and Nahal Lotz watersheds. The closest permanent water sources available at present are located some 5 km away to the southwest at Be’erot Oded. The rock-shelter faces to the east, and the occupation area is permanently in the shade. Occupation layers at NQR have been extensively eroded due to the fluvial action from a gully above the rockSee also remarks by Phillips (1987). Another instance of such an ‘admixture’ is the report on five similar surface assemblages with supposedly ‘mixed’ Upper Palaeolithic (Levantine Aurignacian) and Epipalaeolithic components in Wadi Sudr, western Sinai (Baruch and Bar-Yosef 1986). We observed a similar phenomenon of surface concentrations with two raw materials and a clear distinction between microliths and medium-sized tools on chalcedony and macro-tools on cherty flint elsewhere on Har Harif (Goring-Morris 1992). 2 

Nahal Neqarot rock-shelter (OIG 12960/99364; UTM: 67169/37926) was discovered in the 1980s within the framework of the Emergency Archaeological Survey of the Negev (Survey of Israel Site 223(1299/93/4); Rosen 1994). A one week season of excavations was conducted in 1991 (License #120/91), directed by the authors, I. Gilead and S. A. Rosen under the auspices of the Israel Antiquities Authority, Ben Gurion University of the Negev and the Hebrew University of Jerusalem. 1 

48

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev

Figure 1. Map of the central Negev, with Nahal Neqarot south of Maktesh Ramon marked.

shelter (Figs. 2-3). Field observations indicate that the cultural sediments likely extend along the parabolashaped wall of the rock-shelter for at least 20 m. Following the Epipalaeolithic occupation of the rockshelter, huge segments of the roof collapsed. These massive blocks, covering the sediments, apparently contributed to the preservation of the anthropogenic levels. The dip of the occupation levels relative to the present talus slope clearly indicates that a considerable area was truncated and eroded. It thus appears that the occupation was originally substantial, perhaps extending over 150-200 m2.

grey powdery, ashy sediments with lenses of lighter and darker colour, sometimes with a greenish hue. Stones, especially towards the base of the occupation, seem to represent the flaking off and disintegration of the rock-shelter wall. An 8 cm thick hearth (square Q10b/d, 217-225 cm below datum), ca. 1.0 m in diameter, had a dark grey/black fill with reddish base. Small quantities of charcoal and numbers of rodent burrows were noted throughout the sequence. The intensity of occupation, as reflected by the lithic densities, was considerable from the bedrock up through the entire sequence, densest towards the top.

Some 20 m2 were collected and excavated in two areas, revealing that occupation sediments reached a depth of at least 60 cm to bedrock (Figs. 3-4). A grid of 0.5 x 0.5 m squares was excavated in arbitrary 5 cm spits where levels could not be distinguished. Heights were measured below an arbitrary datum. The occupation layers were almost entirely anthropogenic, comprising

Charcoal Charcoal samples for dating were collected particularly from the lower portions of the section. Three samples from the main test pit reported here derive from grid squares Q/R9 (Figs. 3-4) had provided the following results (Fig. 5): 49

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 2. Nahal Neqarot rockshelter from the southeast. The Ramonian occupation layer underlies the massive fallen blocks.

Figure 3. Plan and general sections of Nahal Neqarot rockshelter.

50

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev

A

B Figure 4. Close-up of the two excavation areas and drawn section of grid squares Q/R along the 9 line. Note that the occupation layer is located directly on bedrock and is overlain by the massive roof collapse.

The C values indicate that the samples derive from C3 vegetation, indicating somewhat more humid conditions than today. While the results correspond to the stratigraphy, only the uppermost sample (OxA4794), deriving from midway up the section, appears to relate to the Early Ramonian occupation determined

as such through the characteristics of the lithic assemblage. Nearly 400 charcoal specimens (ca. 90% of all charcoal recovered at NQR) were analyzed for taxonomic identification (Baruch 1999; Baruch and Goring-Morris 51

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Square

Elevation

Date

sd

Cal BC*

sd

Lab. #

Material

δ13C ‰

NQR

Q9d/R9c

175-185

13,060

100

14,006

413

OxA-4794

charcoal

-22

NQR

R9c/Q9d

195-200

14,760

110

16,138

325

OxA-4795

charcoal

-22

NQR

Q9c

210-215

17,860

160

19,484

415

OxA-4796

charcoal

-19

*calibrated using CalPal

Figure 5. Radiocarbon dates from Nahal Neqarot.

1997). The species represented include: Juniperus (81%), Rhamnus (9%), Ephedra (6%), Tamarix (0.3%), Zygophyllum (0.3%) and chenopods (1%). Except for juniper, all other identified taxa still grow in various habitats in the general vicinity of the site. However, the dominance of juniper in the assemblage is striking since this taxon does not occur today anywhere in the central Negev highlands. At present, it is found only in the Edom mountains (ca. 850-1300 m asl) east of the Rift valley and in the anticlines of northern Sinai (ca. 300-900 m asl). This discontinuous distribution pattern indicates that in the past Juniperus phoenicea should have occurred in intermediate localities between these two regions. The findings at NQR indeed demonstrate that this was the case, since the central Negev highlands are situated midway between the two areas. It may be thus visualized that it was in these same habitats that Juniperus phoenicea could be encountered in Epipalaeolithic times. The absence of Pistacia atlantica is somewhat surprising, given its prevalence locally in other somewhat later Late Natufian and Harifian sites (Baruch and Goring-Morris 1997).

is at least one item that is large and costate (Q10b middle level). There is also a single complete, dorsally perforated cowrie shell with ochre (N8/9 upper level). Most faunal remains derive from the upper part of the sequence. Groundstone tools include: a short limestone pestle (Fig. 6; K12, surface) and three small hard limestone discs with burnished edges, some with evidence of flaking around the circumference (Fig. 6; M7c upper with signs of ochre [complete]; Q10b, lower [broken, polished on the edge]; Q10c lower, burnished all around). They also display slight signs of crushing on the edges and faces, perhaps reflecting use as anvils and retouchers, and/ or for the application of the microburin technique and backing the microliths (and see discussion). A few small lumps and crumbs of ochre were found in both areas. Additionally a couple of probable Byzantine

Material Culture With but a few exceptions (see below), the vast majority of the chipped stone assemblage can be attributed to the Ramonian industry. In addition to the chipped stone assemblage, other finds included: a small faunal assemblage (which remains to be studied); some marine mollusks and a few groundstone artefacts, as well as a single small fragment of a polished bone point (square M8a, upper layer), broken longitudinally. The marine mollusks include six Antalis scaphopods (dentalia); most are smooth-bore (squares Q10d and R10c lower levels; R9c base of upper level; Q9c upper level), though there

Figure 6. Pestle and other groundstone tools from Nahal Neqarot. Both scales in cm.

52

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev and Bedouin ceramic sherds (N9d surface; R9c upper) were recovered.

mostly for medium and macrolithic items, is flint (from different geological sources), which is sometimes cherty. Interestingly, some of the larger tools on such material display double patina (and see below).

Chipped Stone Assemblages Lithic densities of debitage and tools were considerable throughout the sequence in both excavation areas. All chipped stone artefacts from both excavation areas were studied in detail. However, since there are minor indications of earlier and later Epipalaeolithic elements, the results presented here comprise only the finds from the part of the eastern area (Q9/R9), considered to be less affected by post-depositional processes. We arbitrarily subdivided the material in these squares into three units, from bedrock to the surface in order to investigate possible changes through the sequence. With the exception of a few rectangles and trapezes, as well as a couple of small lamelles scalènes/triangles that probably can be attributed to the earlier Geometric Kebaran and the Nizzanan, respectively (GoringMorris 1987), and a single Neolithic arrowhead, the vast majority of the lithics can be confidently assigned to the Epipalaeolithic Ramonian entity. The lithics (from both areas) are in pristine condition. Several possibilities for conjoins were noted. The majority of the artefacts (cores, debitage and tools) are made on translucent chalcedony. As noted elsewhere in the Negev, while chalcedony has excellent flaking properties, the nodules available in local outcrops are usually small (fist-sized and smaller) and commonly contain hair-line cracks. Accordingly, they are suitable only for the production of mediumsized and smaller, microlithic blanks. An additional component,

Cores (Fig. 7) The 26 cores include relatively even distributions between the use of flint (14) and chalcedony (12), with more flint cores in the lower level (Fig. 8). Most are single, unfacetted bladelet cores, three being pyramidal. In the upper level one is on a flake, and one was a flake core. In the middle level two are flake cores, one is mixed, and one is on a flake. In the lower level seven are flake, and four are mixed flake/bladelet cores, with only one for bladelets. Two are on flakes.

Figure 7. Core types and raw materials at Nahal Neqarot.

53

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Core type

variability displayed by other Ramonian assemblages (Goring-Morris 1987: table VII-3).

Lower Middle Upper Total

Single unfacetted

6

6

6

18

Single facetted

1

0

0

1

Double, opposed, same face

3

0

1

4

Two platforms at 90°

1

1

0

2

Roughout & amorphous

1

0

0

1

Total

12

7

7

26

Flint

9

3

2

14

Chalcedony

3

4

5

12

Total

12

7

7

26

Microburin technique (Fig. 11) The microburin technique was used habitually at Nahal Neqarot as indicated by the general and restricted indices that center around 30; higher values for the lower level likely reflect the small sample size (Fig. 11). The other values also accord well with those reported from other Ramonian assemblages (Goring-Morris 1987: table VII: 4-5). The vast majority comprises distal microburins and almost 85% are ‘dexter’, reflecting the pronounced standardization in the application of the technique as in most other Ramonian assemblages (Figs. 12-14). The distal shapes of the blade/let blanks on which the microburins are made, are mostly blunt, pointed, and oblique, including overshot items (Figs. 15-16). This reflects the fact that the configuration of the distal tips of the blanks was not important, since the shape of the microlithic tool derived from the application of the mbt (see Belfer-Cohen and GoringMorris 2002).

Raw material

Figure 8. Core types and raw materials by level at Nahal Neqarot.

The debitage counts indicate that most knapping was conducted on-site (Fig. 9). Debitage: Core (20-135) and Tool: Core (6-60) ratios vary significantly in the different levels. The ratios progressively increase up the section (Fig. 10), representing two extremes of the

Level

PE

Flakes

Blades

Bdlts

RB

CTE

BS

Total

Upper

138

332

163

Middle

71

196

85

302

-

4

4

943

164

2

7

3

528

Lower

37

85

56

49

1

6

2

236

Total

246

613

304

515

3

17

9

1707

Key: PE, Primary elements; Bdlts, Bladelets: RB, Ridge blades; CTE, Core trimming elements; BS, Burin spalls.

Figure 9. Debitage counts at Nahal Neqarot.

Level

Cores

Debitage

Tools

Debit: Core

Tools: Core

Upper

7

943

414

134.7

59.1

Middle

7

528

267

75.4

38.1

Lower

12

236

76

19.7

6.3

Total

26

1707

757

65.7

29.1

Figure 10. Artefact counts and ratios at Nahal Neqarot.

Level

P1

P2

P3 (Pt)

P4 Resharp

Total

Imbt

Imbtr

Upper

165

8

4

2

179

30.2

32.0

Middle

103

3

1

5

112

30.0

31.5

Lower

45

4

-

-

49

39.2

45.4

Total

313

15

5

7

340

31.0

33.2

Key: P1, Microburin; P2, Krukowski microburin; P3, Piquant-trièdre; P4, resharpening microburin. Imbt: Microburins x 100, divided by Microburins + All Tools Imbtr: Microburins x 100, divided by Microburins + Tool Types G1-3, I1-30, J1-14, K1-6

Figure 11. Microburin indices at Nahal Neqarot.

54

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev

Level

n

P

M

D

I

DL

DR

PL

PR

R

L

Upper

179

31

-

137

10

110

15

8

16

126 [84.6]

23 [15.4]

Middle

112

21

1

88

2

79

9

9

12

91 [83.5]

18 [16.5]

Lower

49

21

-

24

4

20

4

6

15

35 [77.8]

10 [22.2]

Total

340

73

1

249

16

209

28

23

43

252 [83.2]

51 [16.8]

Key: P, Proximal; M, Medial; D, Distal; I. Indeterminate; DL, Distal left; DR, Distal right; PL, Proximal left; PR, Proximal right; L, Sinister; R, Dexter. Frequencies in parentheses.

Figure 12. Observations on microburins by level at Nahal Neqarot.

Figure 13. Scans of Microburins.

Type of notch on mbt

n

Regular Inverse Helwan

Upper level

179

177

2



Middle level

112

111

1



Lower level

24

24

-



Total

315

312

3



Distal shape of mbt

n

pointed

oblique

blunt

other

Upper level

137

40

19

65

13

Middle level

87

28

24

24

11

Lower level

24

7

5

8

4

Total

248

75

48

97

28

Figure 15. Distal shapes of microburins by level at Nahal Neqarot.

Figure 14. Types of notch on microburins by level at Nahal Neqarot.

55

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 16. Scans of Ramon points at Nahal Neqarot.

56

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev Tools

where necessary (Fig. 17). In addition we examined all the recovered material irrespective of the integrity of the excavated units; it can be stated that no apparent differences were noted in comparison to the units we present below.

Since there appear to be few differences in the basic composition of the tool classes in all three levels, the following description relates to the total assemblage, making reference to the different stratigraphic units, Typology

Lower

Middle

Upper

N

N

N

A2

Scraper on retouched flake

A7

Scraper on blade

A8a

Scraper on retouched blade

A8b

Scraper on Aurignacian blade

A12

Scraper, varia

1

B2a

Scraper, carinated, broad

1

B2b

Scraper, carinated, broad, on Aurign. bl.

C3

Burin on break/natural face

%

1

0.1

1

2

0.3

1

1

0.1

1

1

0.1

1

0.1

1

2

0.3

1

1

0.1

1

0.1

1 1

Total N

1

C15

Burin, varia

1

1

0.1

D1

Burin/scraper

1

1

0.1

D2

Multiple, other

1

1

0.1

1

E1

Partially retouched blade

E2

Completely retouched blade

E3

Blade retouched on both edges

E4

Inversely/alternately ret. blade

E6

Backed knife

2

1

4

0.5

2

2

0.3

1

1

0.1

1

2

0.3

1

1

0.1

1

E8

Retouched/backed blade, varia

2

3

0.4

E9

Retouched/backed blade, fragment

7

7

0.9

E10

Aurignacian blade

1

1

0.1

G1

Straight truncation

2

2

0.3

G2

Concave truncation

1

1

0.1

1

3

0.4

2

0.3

2

3

0.4

1

1

0.1

4

0.5

1

G3

Oblique truncation

1

1

I4

Inversely retouched bladelet

1

1

I5

Alternately retouched bladelet

1

I7

Completely retouched bldelet

I18

Scalene bladelet

2

I19

Scalene bladelet, basal modification

1

1

2

0.3

I20

Arched backed bladelet

1

1

1

3

0.4

I22

La Mouillah point

2

6

3

11

1.5

I23

La Mouillah point, basal modification

2

0.3

I24

Ramon point

28

114

205

347

45.8

I25

Ramon point, basal modification

2

6

11

19

2.5

I26

Atypical Ramon point

1

1

2

0.3

I29

Helwan bladelet

1

1

2

0.3

2

2

I30

Ret/backed bladelet, varia

I31

Ret/backed bladelet, fragment

13

J2

Rectangle

2

J4

Trapeze

1

J9

Helwan lunate

4

J10

Atypical Helwan lunate

J14

Atypical triangle

2

2

4

0.5

95

142

250

33.0

2

0.3

1

1

0.1

4

9

1.2

2

0.3

3

11

1.5

2 1

7

Figure 17. Typological counts by level at Nahal Neqarot (Type list after Goring-Morris 1987).

57

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Typology

Lower

Middle

Upper

N

N

N

N

Total %

2

0.3

10

6

19

2.5

2

2

0.3

1

3

0.4

3

6

0.8

L2

Borer

2

M1a

Retouched notch (blade/bladelet)

3

M1b

Retouched notch (flake)

M2a

Multiple notches (blade)

N4

Massive battered piece

3

N7

Heavy Duty, varia

2

2

0.3

O1

Retouched flake

2

2

4

0.5

O5

Varia

1

1

2

0.3

267

414

757

100.0

2

Total

76 Figure 17. Continued.

Scrapers (Figs. 18-20): Nine scrapers were recovered with most deriving from the upper level. They are mostly made on thick, robust blade blanks clearly originating from a reduction sequence that differs from that used for the production of microlith blanks. This is apparent in the choice of raw materials for the macrolithic elements (i.e. flint, as opposed to chalcedony for microliths). Hardly any of the items were made on primary elements, and only a few retained some dorsal cortex. This indicates that the larger component in the assemblage does not derive from the initial decortication of nodules. Notable is the presence of double patina on several items, indicating the recycling (and curation?) of larger blade blanks. This phenomenon has been noted in other Ramonian assemblages, e.g. Shunera XXI (Goring-Morris 1987). Several have lateral ‘Aurignacian’ retouch, and flat (broad) carination also appears.

retouched splayed bladelet (see below). Most microliths can be assigned to the Ramon point category, as the majority of the numerous broken microliths are proximal, i.e. basal and medial fragments of Ramon points. Indeed, the Ramon point varieties predominate in the assemblage (48.6%). This includes a few La Mouillah points, all with the mbt scar at the distal extremity. Relatively few of the Ramon points display basal treatment and 129 are broken, i.e. distal parts (Fig. 17). Metric observations of the Ramon points indicate a high degree of homogeneity between the three stratigraphic units (Fig. 22): mean lengths of 30-31mm, mean widths of 7.4-7.8 mm, and mean thicknesses of 2.5-3.0 mm. The shaping of the Ramon points is mostly achieved through semi-abrupt and abrupt retouch. There is also use of fine retouch, with only a few points displaying bipolar and mixed backing (Fig. 23). The backs of the points are mostly on the left side of the blank and have straight or slightly concave outlines (Fig. 24). The point tips are overwhelmingly distal (Fig. 25).

Burins: the few burins are unremarkable (Fig. 20:1). Multiple Tools: although rare, both are scraper combinations, one a burin and the other a truncation; the former is double patinated.

Of interest are signs of often quite intense and distinctive use/battering on many of the unretouched lateral edges of the Ramon points. These factors result in flat removals on the ventral and/or dorsal surfaces. The removal sometimes takes the form of shallow notches, although a few are deep, whether single or multiple, and even serrated (Fig. 16:13-17, 23-25). Such use-wear is unusual in other Ramonian assemblages (pers. obs.). A few points display signs of burning. One Ramon point bears signs of a brown/black material along the back, presumably the remains of hafting mastic.

Retouched Blades: this class forms a heterogeneous group, although there is one ‘Aurignacian’ blade, as well as a backed knife. One partially retouched blade has ochre stains. Truncations: these are mostly oblique on broken blades; several may be items broken during the manufacture of microliths. Microliths (Figs. 16, 22-26): there are a few apparently earlier types based on raw material, morphology and size, distributed through the stratigraphic section. They include small scalene bladelets, sometimes with basal modification, and arch backed bladelets. These are reminiscent of early Epipalaeolithic ‘Nizzanan’ items elsewhere in the Negev (see Goring-Morris 1995). There are also two Helwan bladelets and an inversely

There is a high degree of correlation between the handedness of the Ramon points and the microburins, according well with the results from most Ramonian assemblages (Figs. 25-26; Goring-Morris 1987: table VII5). 58

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev

Figure 18. Scans of endscrapers on thick blades at Nahal Neqarot – note ‘Aurignacian-type’ retouch.

59

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 19. Thick endscrapers with ‘Aurignacian-type’ retouch at Nahal Neqarot.

Geometric Microliths: this category is barely represented. It includes three flint trapeze-rectangles. There are also 11 Helwan lunates (Fig. 21:8-10) and 11 atypical triangles that derive from all excavation units, one being miniature.

Perforators: this class is represented by a single microlithic borer. Notches & Denticulates (Fig. 21: 7, 11): these items are mainly on broken blade/lets. Most of the broken single 60

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev

Figure 20. Tools from Nahal Neqarot – a burin and scrapers.

notches likely represent unsuccessful application of the microburin technique. However, several of the single notches and denticulates are wide and shallow of the

‘spoke-shave’ variety, reminiscent of the typical Early Natufian tool types (and see Gassin et al. 2013; Rozoy 1978 for similar items in the West European Mesolithic). 61

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 21. Tools from Nahal Neqarot – Ramon points, denticulates and a chisel/retoucher.

Massive Tools: they are not numerous, and most derive from the lower unit. They include several massive battered blades and flakes, and a massive denticulate/ scraper on a Core Trimming Element. A couple of the items display a double patina.

Varia: this category is represented by a few retouched flakes. Of interest (though not from the area reported here) was the presence of a chisel/retoucher (Fig. 21:12). 62

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev

Level

n

L

SD

W

SD

Th

L/W

L.W

Upper

115

30.0

4.3

7.6

0.7

3.0

3.9

228.0

Middle

79

30.7

4.9

7.8

0.9

3.1

3.9

239.5

Lower

14

31.5

6.0

7.4

0.5

2.5

4.3

233.1

Total

208 Figure 22. Mean metric attributes of Ramon points by level at Nahal Neqarot (length, width and thickness in mm).

Level

n

Bipolar

Mixed

Abrupt

Semi-abrupt

Fine

Upper

115

2 (1.7)

-

37 (32.3)

24 (20.9)

24 (20.9)

Middle

79

-

4 (5.1)

30 (38.0)

34 (43.0)

11 (13.9)

Lower

14

-

-

3 (21.4)

8 (57.1)

3 (21.4)

Total

208

2 (1.0)

4 (1.9)

70 (33.7)

66 (31.7)

38 (18.3)

Indet.

Figure 23. Retouch/backing types of Ramon points by level at Nahal Neqarot (frequencies in parentheses).

Level

n

Straight

Concave

Wavy

Convex

Upper

115

41 (35.6%)

38 (33.0%)

16 (13.9%)

20 (17.4%)



Middle

79

48 (60.8%)

32 (40.5%)

2 (2.5%)

2 (2.5%)

4 (5.1%)

Lower

14

11 (78.6%)

3 (21.4%)







Total

208

100 (48.1%)

73 (35.1%)

18 (8.6%)

22 (10.6%)

4 (1.9%)

Figure 24. Shape of backed edge of Ramon points at Nahal Neqarot (frequencies in parentheses).

Level

n

Distal

Proximal

Left

Right

mbt

Upper

115 (71)

115 [100%]



103 (64) [89.6%]

12 (7) [10.4%]

51 (4)

Middle

79 (43)

76 [96.2%]

3 [3.8%]

65 (41) [82.3%]

14 (2) [17.7%]

24 (16)

Lower

14 (15)

12 [85.7%]

2 [14.3%]

12 (12) [85.7%]

2 (3) [14.3%]

9 (4)

Total

208

203 [97.6%]

5 [2.4%]

180 [86.5%]

28 [13.5%]

84

Figure 25. Location of Ramon point tips, retouch related to tip, evidence for presence of microburin scar (in parentheses) at Nahal Neqarot.

Level

Points n

D

Upper

115

Middle

79

Microburins

P

L

R

n

D

P

R

L

100.0%

-

89.6%

10.4%

149

83.9%

16.1%

84.6%

15.4%

96.2%

3.8%

82.3%

17.7%

109

80.7%

19.3%

83.5%

16.5%

Lower

14

85.7%

14.3%

85.7%

14.3%

45

53.3%

46.7%

77.8%

22.2%

Total

208

97.6%

2.4%

86.5%

13.5%

303

78.2%

21.8%

83.2%

16.8%

Figure 26. Correlation between Ramon point and microburin alignments attributes at Nahal Neqarot.

63

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Discussion

The most interesting aspect of the Nahal Neqarot Ramonian assemblage is the presence of numbers of broad flat carinated scrapers on blade blanks with lateral ‘Aurignacian’ retouch. They are made on flint and chert, different from the chalcedony used for the microlithic component, and often exhibit a double patina. Actually, the investigation of the site was initiated through the surface findings of such items on the talus of the site. These findings led us to investigate whether this element might represent an earlier, Upper Palaeolithic component, predating the obviously Ramonian microlithic one. Our definitive conclusion is that both components form an integral part of the Ramonian assemblage. Indeed, there are several other Ramonian sites that display similar patterns as noted in the introduction. These include the deflated sites in the Negev highlands at Har Harif K9, Sha’on Hol Site 14, Ramat Matred 190 and RM141, which contain similar microlithic and macrolithic components (Belfer-Cohen 1996; Larson and Marks 1977:173; Marks and Simmons 1977; Yegorov et al. 2015; Yizraeli 1967). In addition, the in situ site of Shunera XXI in the western Negev also contains a similar configuration (Goring-Morris 1987). Here it is of interest to note the phenomenon of much earlier (Levallois) items usually recycled as scraper blanks. They are present in several Terminal Ramonian/Early Natufian sites, including Azariq XV and Upper Besor VI (Goring-Morris 1987, 1998).

The stratigraphic sequence at Nahal Neqarot from bedrock to the top is primarily anthropogenic in nature, consisting of quantities of fine-grained powdery sediments and ashes, as well as hearths. Also notable are post-depositional processes including the presence of rodent burrows, as well as the collapse of a large 2 m thick portion of the rock-shelter roof. The collapse most probably dating to the early Holocene had actually contributed to the preservation of the occupation sediments. The combination of collapse and rodent activities may account for what appears to be some degree of admixture. Still, the vast majority of the assemblage described here, from bedrock to the surface, clearly relates, based on its techno-typological configuration, to the Middle Epipalaeolithic Ramonian culture (Goring-Morris 1987). However, although the dates are in stratigraphic order, only one of the three dates, 14,006±413 cal BC (OxA-4794), falls more or less within the established Ramonian time-frame (based on stratigraphic considerations and other radiocarbon dates, GoringMorris 1995). However, it has a large standard deviation. The other dates are earlier, corresponding to the Early and the beginning of the Middle Epipalaeolithic. One possibility is the presence of old wood. Another is that, as noted above, there are also indications of the very ephemeral presence of earlier elements such as arch backed items and rectangles. These may represent the presence of Nebekian/Nizzanan and Geometric Kebaran mixed in with the Ramonian ashy deposits. These items differ from the main corpus of the lithic assemblage, not only typologically but also in their raw material and dimensions. In addition, we note the presence of a small Helwan retouched component of lunates and other microliths, again throughout the sequence. On the basis of other assemblages in the Negev and Sinai, it seems that these could relate to one of two entities, the Terminal Ramonian or Early Natufian (Goring-Morris and Belfer-Cohen 1997, 2013 and references therein). Both entities are assigned to the beginning of the Late Epipalaeolithic, with the Terminal Ramonian viewed as developing directly out of the local Ramonian. Assemblages lacking any obvious Ramonian components have been viewed as short-term forays into the region by Early Natufian groups based in adjacent Mediterranean zone regions, most notably across the Rift valley to the east. The presence in the Nahal Neqarot assemblages of a few small ‘atypical’ Ramon points may indicate that we are dealing also with a minor Terminal Ramonian component post-dating the principle Ramonian occupation. Additionally, in the western excavation area, at the top of the section, immediately underlying the roof collapse, a few large abruptly backed lunates were noted. These indicate a minor Late Natufian presence.

In summary, the investigations at Nahal Neqarot rock-shelter unequivocally demonstrate that what was thought to be an earlier, Upper Palaeolithic level underlying the main Ramonian occupation is, in fact, integral to the Epipalaeolithic Ramonian use of the locality. This, naturally, has ramifications concerning the dating and cultural attribution of other supposedly mixed assemblages elsewhere in the Negev and Sinai. Acknowledgements The lithic artefact scans were conducted in the Computerized Archaeology Laboratory of the Institute of Archaeology at The Hebrew University of Jerusalem. References Baruch, U. 1999. The Contribution of Palynology and Anthracology to Archaeological Research in the Southern Levant. In S. Pike and S. Gittin (eds), The Practical Impact of Science on Near Eastern and Aegean Archaeology: 17-28. Weiner Laboratory Monographs 3, London, Archetype Publications. Baruch, U. and Bar-Yosef, O. 1986. Upper Paleolithic Assemblages from Wadi Sudr, Western Sinai. Paléorient 12(1):69-84. Baruch, U. and Goring-Morris, A. N. 1997. The Arboreal Vegetation of the Central Negev Highlands, Israel, at the End of the Pleistocene: Evidence from 64

A. Belfer-Cohen and A. N. Goring-Morris: The Riddle of the ‘Aurignacian’ in the Negev Archaeological Charred Wood Remains. Vegetation History and Archaeobotany 6:249-259. Bar-Yosef, O. 1970. The Epi-Palaeolithic Cultures of Palestine. Unpublished PhD thesis, Hebrew University of Jerusalem. Belfer-Cohen, A. 1996. Problems in Defining a Prehistoric Culture: An Example from the Southern Levant. In M. Otte (ed.), Nature et Culture: 247-260. Liège, Etudes et Recherches Archéologiques de l’Université de Liège: 68. Belfer-Cohen, A. and Bar-Yosef, O. 1981. The Aurignacian in Hayonim Cave. Paléorient 7 (2):19-42. Belfer-Cohen, A., Gilead, I., Goring-Morris, A. N. and Rosen, S. A. 1991. An Epipalaeolithic Rockshelter at Nahal Neqarot in the Central Negev. Mitekufat Haeven – Journal of the Israel Prehistoric Society 24:164-168. Belfer-Cohen, A., Gilead, I., Goring-Morris, A. N. and Rosen, S. A. 1993. Nahal Neqarot. Excavations and Surveys in Israel – Hadashot Archeologiot 13/100:118119. Belfer-Cohen, A. and Goring-Morris, A. N. 2002. Why Microliths? Microlithization in the Levant. In R. G. Elston and S. L. Kuhn (eds), Thinking Small: Global Perspectives on Microlithic Technologies: 57-68. Archeology Papers of the American Anthropological Association, AP3A 12. Arlington, Virginia, American Anthropological Association. Copeland, L. 1975. The Middle and Upper Paleolithic of Lebanon and Syria in the Light of Recent Research. In F. Wendorf and A. E. Marks (eds), Problems in Prehistory: North Africa and the Levant: 317-350. Dallas, Southern Methodist University Press. Gassin, B., Marchand, G., Claud, E., Guéret, C. and Philibert, S. 2013. Les lames à coches du second Mésolithique: des outils dédiés au travail des plantes? Bulletin de la Société Préhistorique Française 110 (1):25-46. Gilead, I. 1981a. Upper Palaeolithic in Sinai and the Negev: Sites in Gebel Maghara, Qadesh Barnea and Nahal Zin. Unpublished PhD thesis, Hebrew University of Jerusalem. Gilead, I. 1981b. Upper Palaeolithic Tool Assemblages from the Negev and Sinai. In J. Cauvin and P. Sanlaville (eds), Préhistoire du Levant: 331-342. Paris, Editions Du Centre National Dela Recherche Scientifique. Goring-Morris, A. N. 1987. At the Edge: Terminal Pleistocene Hunter-Gatherers in the Negev and Sinai. British Archaeological Reports, International Series 361, Oxford, British Archaeological Reports. Goring-Morris, A. N. 1992. Har Harif. In D. N. Freedman (ed.), The Anchor Bible Dictionary, vol. III: 56. New York, Doubleday. Goring-Morris, A. N. 1995. Complex hunter-gatherers at the end of the Paleolithic (20,000-10,000 BP). In T. E. Levy (ed.), The Archaeology of Society in the Holy Land: 141-168. London, Leicester University Press.

Goring-Morris, A. N. 1998. Mobiliary Art from the Late Epipalaeolithic of the Negev, Israel. Rock Art Research 15 (2):81-88. Goring-Morris, A. N., Baruch, U., Belfer-Cohen, A. and Rosen, S. 1998. Epipalaeolithic occupations in Nahal Neqarot Rockshelter, Negev Israel: Radiocarbon dating and identification of charred wood remains. Geoarchaeology 13 (2):219-232. Goring-Morris, A. N. and Belfer-Cohen, A. 1997. The Articulation of Cultural Processes and Late Quaternary Environmental Changes in Cisjordan. Paléorient 23 (2):71-93. Goring-Morris, A. N. and Belfer-Cohen, A. 2013. Ruminations on the Role of Periphery and Centre for the Natufian. In O. Bar-Yosef and F. R. Valla (eds), The Natufian Foragers in the Levant. Terminal Pleistocene Social Changes in Western Asia: 562-583. Ann Arbor, International Monographs in Prehistory. Larson, P. A. and Marks, A. E. 1977. Two Upper Paleolithic Sites in the Har Harif. In A. E. Marks (ed.), Prehistory and Paleoenvironments in the Central Negev, Israel. Volume II. The Avdat/Aqev Area, Part 2, and the Har Harif: 173-190. Dallas, Southern Methodist University Press. Marks, A. E. 1977. Introduction: A Preliminary Overview of Central Negev Prehistory. In A. E. Marks (ed.), Prehistory and Paleoenvironments in the Central Negev, Israel, vol. II. The Avdat/Aqev Area, Part 2 and the Har Harif: 3-34. Dallas, Southern Methodist University Press. Marks, A. E. 1981. The Upper Paleolithic of the Levant. In J. Cauvin and P. Sanlaville (eds), Préhistoire du Levant: 369-374. Paris, Éditions du Centre National de la Recherche Scientifique. Marks, A. E. and Simmons, A. H. 1977. The Negev Kebaran of the Har Harif. In A. E. Marks (ed.), Prehistory and Paleoenvironments in the Central Negev, Israel. Volume II. The Avdat/Aqev Area, Part 2, and the Har Harif: 233-270. Dallas, Southern Methodist University Press. Phillips, J. L. 1987. Sinai during the Paleolithic: the Early Periods. In A. E. Close (ed.), Prehistory of Arid North Africa: Essays in Honor of Fred Wendorf: 105-121. Dallas, Southern Methodist University Press. Rosen, S. A. 1994. Map of Makhtesh Ramon (204). The Israel Archaeological Survey, Jerusalem, Israel Antiquities Authority. Rozoy, J.-G. 1978. Typologie de l’Épipaléolithique (Mésolithique) franco-belge. Société archéologique champenoise (n° spécial du Bulletin de la Société archéologique champenoise, juillet 1978), Charleville, Société archéologique champenoise. Yegorov, D., Yaroshevich A., Vardi, J., and Birkenfeld, M. 2015. Sha’on Hol, Site 14 (HG14): A new Late Epipalaeolithic site in the Central Negev Highlands. Journal of the Israel Prehistoric Society – Mitekufat Haeven 45:77-96. Yizraeli, T. 1967. Mesolithic hunters’ industries at Ramat Matred (The Wilderness of Zin), first report. Palestine Exploration Quarterly 99:78-85. 65

Forging a Link: Evidence for a ‘Lost Horizon’ – The Late Chalcolithic to EB 1 Transition in the Southern Levant Eliot Braun Abstract Because of an extraordinarily low level of visibility in the archaeological record, the transition from Late Chalcolithic (LC) to Early Bronze I (EB 1) in the southern Levant was once thought to be evidence for a thoroughgoing and sometimes even a complete break in cultural continuity. While minor increments in knowledge in past decades have somewhat modified those extreme perceptions, only recently has significant new light has been cast on what is a barely understood episode in human activity. Although recent discoveries at select sites proffer definitive evidence for a far more gradual transition than previously perceived, there remain significant problems in identifying relevant segments of the archaeological record pertinent to that time span, dated to between c.3800 BCE and c.3600 BCE. While recognizing challenges and difficulties inherent in teasing out information on a rather poorly documented segment of the archaeological record, this paper reviews some definitive evidence for a significant degree of continuity between LC and EB 1, while critically examining select scholars’ claims for a far greater degree of continuity and persistence of traditions between these periods than is actually discernible from extant information. Keywords: South Levantine late prehistory, Chalcolithic-Early Bronze 1 Transition, ‘Lost Horizon’, Modi’in Buchman, Ashqelon Cluster, Adeimeh, Tell esh-Shuna, Hujayrat al-Ghuzlan and Jebel Mutawwaq

A short history of research on the LC to EB 1 transition

unfounded and that it was more properly assigned to EB 1. Accordingly, they concluded that the two periods were primarily defined by distinctly different ceramic types.

More than eight decades separate us from the first substantial publication of the quintessential south Levantine Chalcolithic site of Teleilat Ghassul (Ghassul) with its distinctive material culture (Mallon 1932; Koeppel 1940) and the discoveries of additional early pottery horizons at Megiddo (Engberg and Shipton 1934), Beth Shan (FitzGerald 1934; 1935) and Afula (Sukenik 1936; 1948), also termed ‘Chalcolithic’ and, additionally, ‘Early Bronze’ (Wright 1937). Those early excavations rightfully assumed a chrono-cultural sequence of Chalcolithic to EB 1. However, it was early days, and the archaeological record of the southern Levant had not been extensively revealed. In addition, excavation techniques were not particularly refined. Consequently, the level of understanding of the relationship between those chrono-cultural periods was, and for a long time thereafter, remained slight (Braun and Roux 2013).

Despite Wright’s new LC paradigm, for decades some scholars (e.g., Mellaart 1966:29,35, 66; de Vaux1 1970:531-536; 1971; Perrot 1972:439; Elliott 1978:5051; Hanbury-Tenison 1986) continued to link GBW with an LC phase and also to relate it to early EB 1 ceramic traditions. Notably, that linkage was based on purported evidence from Meser (Dothan 1957; 1959; 1971), where GBW was attributed to three successive occupations, the earliest of which (Stratum 3) yielded hallmark Chalcolithic (so-called ‘Ghassulian’) pot types, and subsequent occupations (Strata 2 and 1) associated with distinctively EB 1 pottery. However, GBW and other related types (Braun 2012a) have never been found in situ in association with occupations dated by pot-types that define LC. Furthermore, examination of LC assemblages makes it eminently clear that GBW, the product of full-blown craft specialization, had no recognizable antecedents in south Levantine Chalcolithic potting traditions. Additional definitive evidence of GBW having been divorced from Chalcolithic traditions derives from a broad exposure of Stratum II at Yiftah’el II, replete with

Based on what was then known of the archaeological record, G. E. Wright (1937) in his seminal PhD thesis posited a Late Chalcolithic (LC) phase associated with a particular type of pottery, now commonly known as Gray Burnished Ware (GBW; Braun 2012a:6-11), associated with bowl types then recognized as LC hallmarks. However, several decades of observation of the revealed archaeological record later indicated to Wright (1958), and also to Amiran (1963; 1969:2234) that the LC ascription of the distinctive GBW was

R. de Vaux (1970, 530; 1971) was somewhat of an exception as he defined GBW as LC, but he rejected a claim of any association with Ghassulian Chalcolithic. 1 

66

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’ examples of early GBW found in situ on floors of early EB 1 houses. Not a single Chalcolithic pot type was associated with that occupation, although some similarities of features indicated a modicum of inspiration likely derived from the earlier period (Braun 1989a; 1997). That significant disparity with LC traditions was further emphasized by distinctive, curvilinear architecture at the site that has no parallels in the LC of the southern Levant.2 Accordingly, I postulated an intermediate or transitional phase between the latest recognized LC and the Yiftah’el II Phase of early EB 1 (Braun 1997), which I somewhat whimsically labeled the ‘lost horizon’ (Braun 1996; 2011).

real insights into the nature of the transition from an obscure and elusive LC to EB 1 period. Unfortunately, chronological phasing of the Chalcolithic period and especially its latest manifestations remain poorly defined as none of the type-sites was occupied in early EB 1, leaving open the question of what the latest manifestation of Chalcolithic is. Some scholars’ understanding of Ghassul as a typesite, eventually, following R. Neuville (1930), came to equate the eponymous ‘Ghassulian’ with Chalcolithic. Later, with the discovery of the Beersheva Valley Chalcolithic sites, the term ‘Ghassulian-Beersheva’ came into vogue. While those terms were often used to designate LC, they are known to be problematic as the earliest levels at Ghassul should be understood as considerably earlier and possibly even Neolithic (Hennessy 1969), while the latest occupations at the Negev sites could not be definitively defined as the latest manifestations of Chalcolithic culture, although some radiocarbon dates from Shiqmim and even Gilat (however unlikely), could have suggested such a possibility (Gilead 2011:14, passim). More recently, Gilead (2011) has defined ‘Ghassulian’ as representative of a relatively late facies of the period, effectively excluding earlier phases associated with different material culture hallmarks. However, to date, there is no clear picture of precisely what constitutes LC, although a cluster of sites at Fazael in the Jordan Valley is yielding some significant evidence for that phase (Bar et al. 2015). The following review in some ways helps to clarify the archaeological record of this period.

Thus, GBW was understood to be associated with an early, but not the earliest phase of EB 1. Meser, with its sparse evidence presented only in preliminary publications, with no accompanying data on provenience of objects, proved to be a ‘red herring’ for chrono-cultural associations of GBW (Braun and Roux 2013:16-17).3 In retrospect it was clear the site had yielded only non-discrete, chronologically mixed artifact assemblages, as an in-depth study of the stratigraphic and architectural evidence suggested (Braun 1989b; 1996). Rather unfortunately, the result of Dothan’s false linkage of GBW to LC was, for decades, to effectively stifle research on the transition between those chrono-cultural phases, while sidetracking attempts at discerning material culture associated with an LC to EB 1 interface. From late in the last century attempts have been made to investigate the transition (e.g., Amiran 1977; 1985; Hanbury-Tenison 1986; Braun 1989a; 2000a; Joffe and Dessel 1995). More recently, the application of radiocarbon dating to the archaeological record, and especially its recently improved methodology and widespread use, have provided many new data (e.g., Regev et al. 2012; Braun et al. 2013) that have greatly altered understanding of the chronology of the Chalcolithic and EB 1 periods and prompted scholars’ new interpretations of the evidence (e.g., Gilead 1994; 2011; Bourke 2001; Bourke et al. 2001; 2004; Braun 2001; Burton and Levy 2001; 2011; Blackham 2002; Bourke and Lovell 2004; Klimscha 2009; Rowan and Golden 2009). However, until rather recently there was little in the published excavated archaeological record that offered

The LC to EB 1 transition or the ‘lost horizon’ Within the past decade excavations at several sites (Fig. 1) have yielded new data on an elusive LC-EB 1 transition, making it quite clear that some human occupation of the southern Levant, however limited, continued, with no break in human activity between LC and EB 1. The most convincing evidence comes from Modi’in Buchman (van den Brink 2013), while that from a cluster of sites on the Mediterranean Littoral in modern Ashqelon (the Ashqelon Cluster) is far more equivocal and problematic. Rather surprisingly, two arid-zone sites, Tall Huyjarat al-Ghuzlan and Tell Magass (Lutfi and Schmidt 2009) in southern Jordan near Aqaba, apparently associated with copper production, may have actually experienced a floruit of activity in that time span. In addition, there is also a modicum of evidence for this same period from older excavations at the central Jordan Valley site of Tell esh-Shuna and in a unique burial near Jericho, while a cluster of sites at Fazael on the western side of the Great Rift Valley has yielded definitive evidence of LC (Bar et al. 2015). Following are brief descriptions of the evidence from these sites and evaluations of the validity

Comparanda with the enéolithique récent of Lebanon do, however, suggest associations with the Chalcolithic of that region (Braun 1989b). 3  Similarly misleading data are from Qatar Damiyeh in the Jordan Valley where Kaptijn and de Vrees (2008) claim to have discovered pottery definitively dated to the LC-EB 1 transition. That claim, however, is merely based on a pottery assemblage derived from a surface collection at a site not then excavated. Thus, it offers no definitive evidence for chronological associations between what they claim to be LC and EB 1 pot types. Their claim is particularly suspect as those scholars cite comparanda from Gilat (Gilead and Fabian 2010: *96-*97; Gilead 2011), a primarily early Chalcolithic site, as supportive of their purported paradigm. 2 

67

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead series of radiocarbon determinations allowing us to bracket the transition to between ca. 3800 and 3600 BC (Braun et al. 2013; Roux, van den Brink and Shalev 2013; see also below). Stratum 4 at Modi’in-Buchman, assigned by the excavator, E. C. M. van den Brink, to the latest Chalcolithic (his LC 2 Phase4), is notable for rectilinear architecture, as is the earliest EB 1 occupation of Stratum 3. There is also a modicum of continuity in morphology of ceramic vessels between those periods. However, more telling are micro-studies on the pottery assemblages of those strata (Roux, van den Brink and Shalev 2013) that have vastly refined our knowledge of potting traditions by indicating significant changes in clay sources and details of manufacture reflecting chronological developments. They signal greater cultural changes that took place with a waning of LC traditions and the onset of what is traditionally understood to be early or initial EB 1 (Braun and Gophna 2004), sometimes known as EB 1A.5 The sequence from Modi’in Buchman, which brackets the transitional LC to EB1 between Strata 4 and 2, Roux’s detailed ceramic studies cited above, and radiocarbon dates, all enable us to use the successive ceramic assemblages from this site as reliable benchmarks for comparison of this aspect of material culture in determining additional LC-EB 1 transitional sites. Evidence from Tell esh-Shuna North In two small soundings at the central Jordan Valley site of Tell esh-Shuna, H. de Contenson (1960a; 1960b; 1961) and C. Gustavson-Gaube (1985; 1986) documented deep, stratified deposits that appear to include evidence of the LC to EB 1 transition. In de Contenson’s paradigm that might be a transition between Levels I and II, with GBW appearing in the later level. In a sequence excavated by Gustavson-Gaube (1985:49) her comments suggest that the ceramic aspect is likely analogous to that at Modi’in. She noted ‘…persistent use of established potterymaking traditions throughout the excavated sequence, independent of what the archaeologist may wish to call Chalcolithic or Early Bronze 1.’ Notably GBW appeared c. one third of the way through that continuum, indicating it was a somewhat tardy development. Thus, the first third of that sequence seems likely to contain evidence for the LC – EB 1 transition. Significant evidence of architectural remains in those soundings, some of which is curvilinear, suggests affinities with Chalcolithic and early EB 1 sites. Unfortunately, however, both soundings were too limited in scope to offer more than mere hints

Figure 1. Map of southern Levant with principal LC to EB 1 transition deposits.

of interpretations offered by scholars who excavated them. Evidence from Modi’in Buchman This site in the Shephela (the western piedmont of the Judea-Samaria Incline) has yielded extraordinarily relevant results concerning the transition from LC to EB 1, in particular in a sequence of superimposed ‘Deep Deposits’ associated with ceramic assemblages and architectural remains that throw intense light on the nature of the transition (Roux, van den Brink and Shalev 2013; van den Brink 2013). Studies of the pottery and architecture from this site offer a detailed view of change and development along a time trajectory; one that clearly evinces an LC to earliest EB 1 interface. Most fortunately, that site and others have yielded a

This is the later of two LC phases. I particularly eschew this, and its parallel, later EB 1 designation, EB IB, as neither has been properly defined. Neither are these designations even reasonably accurate enough for any detailed discussion for c. 800 years of human activity represented by the LCEB 1 and EB 1 periods. 4  5 

68

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’ of the nature of the ‘lost horizon’. However, G. Philip, at a recent ICAANE workshop in Vienna in April 2016, suggested there might actually have been a hiatus in occupation at the site. Thus, future publication of data may alter the present interpretation.

be considered definitive evidence for any hypothetical reconstruction of the chrono-stratigraphy of the sites. While some of those samples may indeed reflect their purported contexts, others may represent specious associations and cause error in interpretation of the archaeological record of the cluster. As that record is potentially central to any discussion of the LCEB 1 transition, following is a detailed discussion of problematic aspects of the interpreted record. In that regard one should keep in mind that all sites in the Ashqelon Cluster were artificially leveled prior to their excavation (see below).

Evidence from Yesodot Excavations at this Judean Shephela site yielded material culture that shares aspects of LC and EB 1, which caused the excavators (Paz and Nativ 2013) to cite it as ‘postGhassulian’.6 The architecture of the site is rectilinear, while the pottery assemblage has features that suggested to the excavators it should be considered a ‘precursor’ of EB 1 traditions. Notably absent are Chalcolithic hallmark pot types such as cornets and churns. The chipped stone assemblage also indicates new features associated with EB 1 traditions, including a sizable number of prismatic (i.e., ‘canaanean’) flint blades, with what are described as Chalcolithic type modifications, suggesting the adoption of a new technology altered by older traditions. An infant jar burial, rarely encountered at EB 1 sites (see below) seems also to be a continuation of a Chalcolithic tradition at Yesodot (cf. Bar 2014:316317). Thus, the sum of the features at this site suggest it should be assigned to the LC-EB 1 transition.

Afridar Select precincts of this neighborhood were intermittently subjected to salvage excavations7 during a lengthy period beginning in 1968 (Gophna 2004). Pertinent to the present discussion are results from work at locales designated as Areas A, B, C, D, E, F, G, J and M. Some of those sites revealed early EB 1 curvilinear buildings and very early types of EB 1 pottery (Braun 2000a; 2000b) as well as Chalcolithic artifacts (Braun and Gophna 2004:226-227; Golani 2004; Baumgarten 2004; Khalaily 2004; Gophna 2004:3-4), while others offered evidence of metallurgical activity. Barnea

The Ashqelon Cluster of Sites

Excavations by Golani and Nagar (2011) in select precincts (Fig. 3) of a large area (c. 1 ha) unearthed primarily rectilinear structures associated with jar burials, a circular structure probably for storage, and a separate cemetery of cist burials. Finds included pottery types of different phases of EB 1 as well as a modicum of artifacts that have excellent comparanda at Chalcolithic sites. Most unfortunately, none of the jars from the burials has been published and no comparanda for them have been cited by the excavators.

The Mediterranean Littoral at modern Ashqelon was the location of a cluster of late prehistoric sites situated along the waterfront for several kms between Tel Ashqelon in the south to at least as far north as the Barnea neighborhood. Numerous excavations have unearthed a patchwork of occupations dating to several phases of late prehistory in two neighborhoods, Afridar and Barnea (Baumgarten 2004; Braun and Gophna 2004; Golani 2004; 2008; 2013; 2014; Gophna 2004: Fig. 1; Golani and Nagar 2011; Golani and Rosenberg 2012; Khalaily 2004).

Reevaluation of the Evidence from the Ashqelon Cluster

Available data are interpreted in particularly variant ways, offering very different reconstructions of the chrono-cultural record of the region, but a series of radiocarbon dates (Fig. 2) strongly suggests that some remains should be dated to the LC-EB 1 transition. Most unfortunately, the ascription of particular radiocarbon samples to specific archaeological deposits and of other remains by purported associations claiming to indicate continuity from the LC to EB I, are highly problematic. While the 14C samples recovered are probably accurate in themselves, their contexts are of the lowest level of validity (Braun et al. 2013:31) and thus, there is no assurance they actually represent the archaeological deposits they purport to date. Therefore, they may not

Despite claims in recent publications (Golani and Nagar 2011; Golani and Rosenberg 2012; Golani 2013) that together with Afridar, the Barnea finds represent a complete and continuous EB 1 sequence, there is no definitive evidence for such a stratigraphic paradigm; merely written claims of such. Those scholars claim the earliest EB 1 was found at Afridar and that later phases, which purportedly lasted until the very end of EB 1 were encountered at Barnea, and that the entire sequence represents continuity from the Chalcolithic period. However, they do not offer any definitive evidence for their claims. Instead they merely based those claims on a highly subjective interpretation of the evidence Usually these were small, individual building lots parceled out on flattened surfaces prepared for development, but Area E, intended for a public building, was much larger. 7 

Ghassulian in this sense indicates late Chalcolithic as expressed by Gilead (2011). 6 

69

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Site

Lab Code

Age

Calibrated 1 δ

Calibrated 2 δ

Location & Description

4241 BC (93.9%) 3760 BC 3742 BC (1.5%) 3714 BC

Area E1; debris within pit; olive stones

Afridar

RT-2634

4223 BC (2.6%) 4209 BC 4156 BC (4.3%) 4132 BC 5170±100 4065 BC (43.3%) 3906 BC 3880 BC (18.1%) 3801 BC

Afridar

RT-2256

5057±75

3954 BC (68.2%) 3785 BC

3986 BC (94.4%) 3694 BC 3680 BC (1.0%) 3666 BC

Area E1: debris within pit; charcoal, olive wood

Afridar

RT-2272

4890±70

3768 BC (68.2%) 3635 BC

3932 BC (3.9%) 3877 BC 3805 BC (77.5%) 3619 BC 3611 BC (14.0%) 3521 BC

Area E1: debris within pit; charcoal, olive pits

Afridar

RT-2255

4800±65

3652 BC (68.2%) 3520 BC

3704 BC (81.7%) 3497 BC 3456 BC (13.7%) 3377 BC

Area E1: debris within pit; charcoal, olive wood

Afridar Area

RT-2254

5065±45

3944 BC (22.3%) 3905 BC 3880 BC (45.9%) 3801 BC

3966 BC (94.6%) 3764 BC 3723 BC (0.8%) 3716 BC

Area E1: debris within pit.; charcoal, olive wood, olive pits

3499 BC (15.4%) 3435 BC 3379 BC (38.1%) 3264 BC 3242 BC (41.9%) 3103 BC

Area F: Stratum I within Building 150,; charcoal

Afridar

RT-2567

4575±45

3493 BC (9.8%) 3468 BC 3375 BC (28.2%) 3328 BC 3217 BC (15.9%) 3178 BC 3159 BC (14.3%) 3122 BC

Afridar

RT-2644b

4945±45

3769 BC (52.7%) 3692 BC 3685 BC (15.5%) 3661 BC

3906 BC (3.2%) 3880 BC 3801 BC (92.2%) 3643 BC

Area G: Strata I-II, fill within and below circular building; charcoal

Afridar

RT-2645b

4890±30

3695 BC (68.2%) 3647 BC

3712 BC (95.4%) 3637 BC

Area G: Strata I-II, fill within and below circular building; charcoal

Afridar

RT-2647

4855±30

3693 BC (8.4%) 3684 BC 3663 BC (59.8%) 3635 BC

3705 BC (85.3%) 3631 BC 3561 BC (10.1%) 3537 BC

Area G: Strata I-II, fill within and below circular building

Afridar

RT-2469b

4990±45

3906 BC (10.4%) 3880 BC 3801 BC (57.8%) 3705 BC

3942 BC (23.8%) 3857 BC 3843 BC (0.7%) 3837 BC 3821 BC (70.9%) 3659 BC

Area E2: debris within pit; charcoal

3892 BC (0.4%) 3885 BC 3798 BC (90.9%) 3631 BC 3578 BC (0.4%) 3572 BC 3567 BC (3.7%) 3536 BC

Area E2: debris within pit; charcoal

Afridar

RT-2258a

4900±45

3759 BC (7.1%) 3743 BC 3714 BC (61.1%) 3639 BC

Afridar

RT-2447

4840±50

3700 BC (38.6%) 3620 BC 3580 BC (29.6%) 3530 BC

3720 BC (95.4%) 3510 BC

Area E2: debris within pit; olive stones, charcoal

Afridar

RTT-4672

4806±40

3645 BC (15.5%) 3629 BC 3585 BC (52.7%) 3530 BC

3692 BC (0.5%) 3686 BC 3661 BC (93.9%) 3517 BC 3396 BC (1.0%) 3386 BC

Area M: Stratum II, debris upon surface; olive stone

Afridar

RTT-4673

4780±40

3637 BC (8.4%) 3627 BC 3597 BC (59.8%) 3526 BC

3648 BC (88.2%) 3512 BC 3425 BC (7.2%) 3382 BC

Area M: Stratum II, debris upon surface; olive stone

Afridar

RTT-4674

4703±40

3624 BC (10.9%) 3602 BC 3524 BC (14.7%) 3497 BC 3457 BC (42.6%) 3377 BC

3632 BC (23.6%) 3560 BC 3537 BC (21.3%) 3486 BC 3474 BC (50.6%) 3371 BC

Area M: Stratum II debris upon surface; olive stone

Afridar

RT-2157

4945±55

3771 BC (68.2%) 3660 BC

3936 BC (9.0%) 3873 BC 3809 BC (86.4%) 3639 BC

Area E1: debris above pit, hearth; charcoal, olive wood

Afridar

RT-2219

4755±45

3635 BC (66.8%) 3519 BC 3393 BC (1.4%) 3390 BC

3641 BC (75.4%) 3497 BC 3452 BC (20.0%) 3377 BC

Area M: debris within pit; charcoal, olive wood

Afridar

RT-2441/ 2442

4925±45

3761 BC (11.6%) 3741 BC 3731 BC (3.0%) 3726 BC 3715 BC (53.6%) 3652 BC

3792 BC (95.4%) 3640 BC

Area J: Stratum V, within a jar; charcoal

3089 BC (10.0%) 3051 BC 3031 BC (58.2%) 2887 BC

3331 BC (9.2%) 3214 BC 3187 BC (1.7%) 3156 BC 3129 BC (81.2%) 2862 BC 2807 BC (2.8%) 2758 BC 2718 BC (0.5%) 2707 BC

Area A2: Stratum III, stones recovered from debris directly upon a sealed mudbrick floor within a large circular mudbrick granary (W11) associated with Stratum III; olive stones

3488 BC (1.3%) 3472 BC 3372 BC (94.1%) 2926 BC

Area B: Stratum V, a lump of charcoal recovered from fill of loose sands mixed with clayey lumps directly beneath a Stratum IV surface (L254B)

Barnea

Barnea

RTT-5434

RTT-5428

4335±75

4500±80

3348 BC (68.2%) 3096 BC

Figure 2. Radiocarbon dates from the Ashqelon Cluster of sites.

70

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’

Site

Barnea

Barnea

Barnea

Barnea

Lab Code

RTT-5436

RTT-5437

RTT-5431

RTT-5432

Age

Calibrated 1 δ

Calibrated 2 δ

Location & Description

4300±75

3084 BC (3.9%) 3066 BC 3028 BC (61.5%) 2872 BC 2801 BC (1.6%) 2792 BC 2786 BC (1.1%) 2780 BC

3316 BC (1.4%) 3273 BC 3266 BC (1.6%) 3237 BC 3169 BC (0.2%) 3164 BC 3111 BC (76.0%) 2834 BC 2818 BC (15.8%) 2663 BC 2648 BC (0.4%) 2636 BC

Area D: Stratum II, stones from a floor within partial remains of stone and mudbrick structure (W50) associated with Stratum II; olive stones

4300±75

3084 BC (3.9%) 3066 BC 3028 BC (61.5%) 2872 BC 2801 BC (1.6%) 2792 BC 2786 BC (1.1%) 2780 BC

3316 BC (1.4%) 3273 BC 3266 BC (1.6%) 3237 BC 3169 BC (0.2%) 3164 BC 3111 BC (76.0%) 2834 BC 2818 BC (15.8%) 2663 BC 2648 BC (0.4%) 2636 BC

Area D: Stratum II; olive stones upon a surface directly over a Stratum III mudbrick wall (W156) ; olive stones

4590±75

3508 BC (21.7%) 3427 BC 3381 BC (18.6%) 3321 BC 3273 BC (0.9%) 3268 BC 3236 BC (15.2%) 3170 BC 3163 BC (11.8%) 3113 BC

3627 BC (2.4%) 3593 BC 3527 BC (92.4%) 3090 BC 3046 BC (0.6%) 3034 BC

Area G: from lower portion of debris upon a floor found filling an underground silo (W103) probably associated with Stratum IV; olive stones

4595±75

3514 BC (25.0%) 3423 BC 3404 BC (0.8%) 3399 BC 3384 BC (18.5%) 3323 BC 3235 BC (13.6%) 3172 BC 3162 BC (10.3%) 3117 BC

3629 BC (3.3%) 3586 BC 3530 BC (92.1%) 3091 BC

Area G: Stratum IIIB; olive stones

Barnea

RTT-5435

4250±75

2927 BC (31.3%) 2841 BC 2814 BC (36.9%) 2678 BC

3084 BC (1.2%) 3066 BC 3028 BC (93.8%) 2620 BC 2606 BC (0.3%) 2601 BC 2592 BC (0.2%) 2589 BC

Area J: Stratum IIIA; L2047; B20106; the surface make-up associated with a Stratum III wall (W205). Theoretically the surface could also have been associated with Stratum IIIB. Remains of Stratum IV were not found in this area; olive stones

Barnea

RTT-5433

4315±75

3090 BC (6.0%) 3060 BC 3030 BC (62.2%) 2870 BC

3350 BC (95.4%) 2650 BC

Area K: Stratum IIIB, from debris upon a surface associated with Stratum IIIB; olive stones

3500 BC (6.1%) 3430 BC 3380 BC (87.5%) 3010 BC

Area L: Stratum IV, from debris upon a surface associated with a Stratum IV wall (W620); olive stones

Barnea

RTT-5430

4530±75

3361 BC (27.1%) 3265 BC 3242 BC (41.1%) 3104 BC

Barnea

RTT-5429

4775±85

3646 BC (54.6%) 3509 BC 3426 BC (13.6%) 3382 BC

3703 BC (95.4%) 3369 BC

Area M: Stratum IVA stones recovered from debris upon Stratum IV surface, olive stones

Afridar

RT2247/8

4545±105

3495 BC (4.8%) 3466 BC 3375 BC (62.1%) 3090 BC 3044 BC (1.2%) 3036 BC

3620 BC (0.6%) 3610 BC 3522 BC (94.8%) 2927 BC

Square D4, Stratum I, within curvilinear Building 150; sterile sand layer separates between earlier EB 1A pits; charcoal

Afridar Area F

RT-2451

-

Modern Figure 2. Continued.

of material culture, particularly that which is clearly Chalcolithic. For example, Golani and Nagar (2011:91) claim:

of the early EB I at Ashkelon. This is to be expected when a large degree of cultural continuity can be recognized in the transition between these two periods, and does not necessarily indicate Chalcolithic habitation at this site or in its vicinity’.

Thus the Chalcolithic ceramic remains found within the EB 1 strata at the site of Ashkelon Barnea and in other excavations that recovered remains of the EB 1 at Ashkelon are here interpreted as representing a Chalcolithic element within8 the material culture

That interpretation, and that of Golani and Rosenberg (Golani 2004; Golani and Rosenberg 2012) suggest any and all features which they, themselves identify as ‘Chalcolithic’ in the Ashqelon cluster were contemporary with EB 1. Accordingly, they claim a

The italic font appears in their original text, where it denotes special emphasis. 8 

71

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead inexplicable hiatus in LC traditions at Afridar, which Golani and colleagues claimed to have lasted for several centuries, but ‘continue’ at the post Afridar, Barnea site. How such a great disparity in interpretation of the archaeological record is possible may be explained by the artificially, bulldozed, truncated topography of the Ashqelon Cluster and particularly in the manner in which the Barnea locale was excavated by Golani and colleagues. Their chrono-stratigraphic paradigm for the Ashqelon Cluster is based on absolute elevations, which fail to take into account a topography that had once preserved significant archaeological deposits above the newly created surfaces in large precincts of the excavated locales. It further ignores compelling evidence of comparanda that argue for far different interpretations (see quotation above). Those scholars frame their argument in a manner that leaves no room for alternative interpretation (see above) because it is a tautology that quite simply claims that because something was found within what one declares to be an EB 1 context, ipso-facto, it must be EB 1 in date. As such it ignored very considerable and significant evidence for Chalcolithic occupation from Afridar (Braun and Gophna 2004) as well as comparanda for Chalcolithic mortuary and other material culture traditions from the Barnea excavations. Topography of the Ashqelon Littoral and its Relationship to Stratigraphy Figure 3. A plan of the excavation locales at Barnea, based on Golani and Nagar 2011, Fig. 7.1.

great deal of continuity from LC to the following period. However, there is actually overwhelming evidence to contradict that claim. Indeed, there is copious evidence to indicate intermittent Chalcolithic (some of it considerably earlier than LC) and EB 1 occupations within the excavated archaeological record of the Ashqelon cluster. This non-continuous paradigm of occupation explains the radiocarbon dates from it that range throughout most of the 4th millennium (Fig. 2). It also accounts for what would otherwise be an 72

Although today the topography of the Afridar and Barnea neighborhoods where excavation was undertaken is completely flat, those locales were once covered by sand dunes, which obscured long ridges several meters high, fossilized dunes of calcareous sandstone or kurkar as it is locally known (Orni and Efrat 1980:41; Ravikovitch 1981:XIV, 23) paralleling the coast on land and just offshore. Archaeological deposits existed on segments of those ridges and between them in soil-filled ‘troughs’ (Gophna 1997; 2004:1) or swales. Evidence of those topographic features exists in photographic documentation (Figs. 4-14,17) and in excavation reports (e.g., Baumgarten 2004:161, 164:

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’

Figure 4. Aerial view of sand dune covering a portion of the Afridar neighborhood c. 1960. Photo courtesy of Professor Ram Gophna’s personal archive.

2004; Braun and Gophna 2004:187-10; Khalaily 2004: Fig. 1, upper right background; Golani 2008:19) although, most unfortunately, not even a single excavation in the entire cluster preceded leveling to what I call ‘ground zero’9, newly bulldozed surfaces (Figs. 5-11,17), today at the street level of the flattened littoral at Ashqelon. Thus, no deposits above ‘ground zero’ were ever archaeologically investigated, which severely limited the recoverable archaeological sequence of occupation anywhere within the Ashqelon Cluster.

a modern surface, while other deposits lay at much higher elevations. In Afridar Area A, Brandl excavated late EB 1 remains in a very low-lying area at elevations analogous to those of Area G, with remains dated to much earlier EB 1 and the Chalcolithic period. Remains of the earlier topography were actually extant during the excavation of Barnea and are photographically documented (Figs. 7-14, 17), but have since been destroyed. Those photographs show vestiges of one kurkar ridge sectioned lengthwise by bulldozers, which rose to a considerable height above ‘ground zero’ (Figs. 11-12).10 Eminently visible in one very long section was a crushed pot clearly in situ at an elevation of ca. 1.5 m above ‘ground zero’ (Fig. 13). That pot and an eminently visible layer of soil bearing many sherds and flecks of carbonized material are positive evidence of archaeological deposits that once existed well above the newly bulldozed surface.

Area G (Fig. 6) was located on the east side of one kurkar ridge and concealed two strata as well as earlier deposits below. Towards the west the deposits were much shallower and clearly were located on a significantly higher portion of the ridge. Anything above ‘ground zero’ was removed with the leveling of the ridge. Area F (Khalaily 2004:121, Fig. 1) had three strata, and contiguous Area J (Baumgarten 2004) yielded five strata, all below ‘ground zero’. In Area M Golani (2008:19) found some remains lying well beneath

Apparently unexcavated in Golani’s salvage project, they clearly negate Golani’s (2014:19) claim that

This is an artificially flattened surface, today more or less at street levels ranging between c. 21 m ASL in Afridar and c. 27 m ASL in Barnea. 9 

This is more or less equivalent to the modern street level, perhaps with a maximum difference of c. 20 cm. 10 

73

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Golani and colleagues to misidentify LC deposits and artifacts and conflate all late prehistoric deposits they encountered into a single, extraordinarily long, purportedly uninterrupted EB 1 sequence. That is because their chrono-stratigraphic paradigm begins at ‘ground zero’ and counts every deposit below as if the cluster were merely a single site with only one sequence of deposits. Thus, strata were assigned to chrono-cultural horizons solely on the basis of absolute elevations, with no consideration that each excavated sequence might well unearth different horizons. Elements at Barnea, which could be seen to be clearly above the newly excavated surface from which all excavation was begun, are not considered in the excavators’ stratigraphic paradigm. The veracity of my interpretation is greatly strengthened by a recent large-scale salvage excavation in the summer of 2018 on the Ashqelon Littoral, c. 3 km southeast of Tell Ashqelon in the Agamim Quarter. There, Abadi-Reis and Varga (2018) and colleagues unearthed a large, LC settlement of three superimposed strata, including rectilinear houses, pottery typical of the Beersheva facies and a significant number of prismatic ‘canaanean blades’ (Abadi-Reis 2018). That find illustrates, beyond doubt, additional LC occupation of the Ashqelon Littoral. Figure 5. Inspector for the Israel Department of Antiquities and Museums, Dov Meiron, stands well below the modern surface in a bulldozer trench cut into a kurkar ridge in the Afridar neighborhood of Ashqelon, 1968. Note the modern ground level next to the high-rise buildings in the background. No trace of that ridge remains today (Fig. 3). Photo courtesy of Professor Ram Gophna’s personal archive.

Observations on Soil Deposits From architectural remains encountered in the Ashqelon Cluster, the use of stone seems to have been confined to low wall foundations of chunks of kurkar, while mudbrick apparently accounted for the mass of materials from which walls were constructed. If one accepts Golani’s and colleagues’ estimate of the time span of EB 1 occupation, one might suspect the depth of deposits at those sites to have been considerable. At Afridar one would look for between 300 and 200 years of an accumulation of mudbrick detritus; at Barnea similar debris would represent another seven centuries of occupation. However, the several excavations show relatively scant evidence of fills bearing occupation deposits.

archaeological remains were only in ‘a natural trough’. While it is impossible to ascribe that vessel to a particular chrono-cultural horizon, (it apparently remained unexcavated and is unpublished) its fabric is similar to those from Afridar dated to late prehistoric horizons.11 Clearly, at least one ridge, unfortunately not documented in preliminary reports on the excavations, had originally covered significant portions of the c.1 hectare area of the Barnea excavation locale, probed only in select, newly flattened precincts as indicated by the published reports. Thus, the documentation of the site’s archaeological record is only partial and restricted to low-lying elevations; ‘ground zero’ and below.

It is noteworthy that the Mediterranean Littoral is subject to heavy rains and squalls coming off the sea, and while there is no rule of thumb to measure earthen accumulations deriving from c. 10 centuries of mudbrick detritus, it should be pointed out that sites occupied for lengthy periods such as nearby Tel Ashqelon in the same topographic niche, are very often the result of continuing decay and subsequent rebuilding, especially in mudbrick. Thus, merely on the basis of the physical evidence presented in the publications, there is some reason to question the purported paradigm of continuous settlement throughout a millennium of occupation.

Failure to take topographic information into account, particularly at the Barnea and Afridar Marina locales (‘Atiqot 45: cover illustrations) has, I believe, caused It was definitively dissimilar to hard-fired Roman and Byzantine fabrics also found in immediate the region (Milevski and Krokhmalnik 2010: Fig. 5). 11 

74

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’

Figure 6. Afridar, Area G during excavation. Note the modern bulldozed surface (‘ground zero’) and low-lying excavated trenches in which early EB 1 remains were found at the base of a truncated kurkar prominence. Photo courtesy of the Israel Antiquities Authority.

Figure 7. View facing south from the excavation area of Barnea of the newly flattened littoral. The high-rise buildings in the distance are at the border of the Afridar neighborhood. Numbers 1-3 indicate three locales excavated at the Barnea site. Note remains of a ridge, sectioned lengthwise by a bulldozer at the right. Beyond it is the sea.

Mortuary Traditions

The Date of the Barnea Cist Cemetery

Three types of burials are associated with the Ashqelon sites. Notably, the burials were found in different arrays and locales. Those in the Afridar locations are rare and different from the large number found at Barnea in two precincts. Most, or even all of them suggest chrono-cultural associations for significant deposits in the cluster that may well not date to EB 1.

This concentration of burials (Fig. 3: Area E Cemetery) represents a formalized cemetery12 of cists, all sans grave goods, some single and in some instances placed end to end in ladder-like formations (Fig. 14:A). They are dated by the excavators to an advanced phase of EB 1 according to ‘proximity to the site’ and similarity in 12 

75

Defined as a cluster of graves in an area devoted solely to burials.

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 8. A view of the bulldozed littoral from the Barnea locale during excavation, facing north. Note hills in the background, remains of the natural topography of the region.

Figure 9. Remains of a kurkar prominence at the northwest extent of ‘ground zero’, Barnea. Beyond it is the sea. The numbers in the foreground mark different excavation locales. Details of archaeological deposits in it are documented in Figs. 12-14.

Figure 10. A partial view of the Barnea locale during excavation, facing south-southwest. Note remains of a kurkar prominence above ‘ground zero’ (Fig. 6) in the center background. The buildings of Afridar are visible in the left background below the shade net and the sea is visible in the right background.

76

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’

Figure 11. View of a bulldozer sectioned segment of a kurkar ridge that parallels the beach at the western border of the Barnea excavation locale (see also Fig. 4). It is covered by grass growing in a layer of sand, below which is a thick layer of soil (the top of which has been emphasized by a white line) replete with archaeological deposits, in situ (see Figs. 13-14), well above ‘ground zero’. The thin, straight white line is string, an aid for drawing.

Figure 12. View of a segment of a bulldozer section at the western edge of the excavation area in Barnea (see Fig. 12). The arrow points to what was apparently a complete pot, in situ, sectioned by the bulldozer.

77

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 13. Closeup view of an in situ pot sheared off by a bulldozer in an archaeological deposit c. 1.5 m above ‘ground zero’ in Barnea.

Figure 14. Comparison of ladder-like arrangements of cist burials from two sites. A: A photograph of a chain of burials in the Barnea cemetery, purported to be EB 1 in date, taken by the author, courtesy of the Israel Antiquities Authority; B: A photograph of a similar ‘ladder-like’ chain of Chalcolithic burials on a kurkar ridge at Palmahim, taken by the author, courtesy of A. Gorzalczany, Israel Antiquities Authority.

78

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’ construction materials and features to buildings there.13 These cists unearthed directly below ‘ground zero’ (Golani and Nagar 2011: Fig. 7.6) are claimed as the latest level in the excavators’ unified stratigraphic paradigm. Although in the excavator’s tautological chronostratigraphic paradigm proximity, absolute elevations and an uppermost position in a sequence must be EB 1 in date, they are only circumstantial ‘evidence’ for dating. Golani’s and colleagues’ interpretation does not consider the possibility that the cemetery was not part of a single sequence of activity at the site, but could be much earlier than the nearby EB 1 occupation, especially when there is evidence of missing deposits above, preserved in the nearby ridge.

for any association between the EB 1 site excavated by me at Palmahim Quarry (Braun 200a; 2000b) and the Barnea cemetery. There are LC burials at the quarry site, but they are in man-made caves hewn from the kurkar bedrock, while the EB 1 burials represent much later utilization of the same caves. At Adeimeh, just northeast of the Dead Sea in Jordan, 168 cist graves in ladder-like formations, all devoid of grave goods, were excavated by M. Stekelis (1935) in a cemetery thought to be associated with the nearby, quintessential Chalcolithic site of Ghassul. Although, a recent excavator of Ghassul, S. Bourke (2001:149) suggested that ‘...the ceramic evidence for Adeimeh remains slight’,17 and that the cemetery may represent a ‘“post-Ghassulian” (i.e., post Chalcolithic) utilization, those statements are clearly refuted by Stekelis’ publication.

Comparanda indicate this cemetery does not date to EB 1, but should be assigned to a much earlier period. Single units in the grave cluster were sometimes of mudbrick, while the ‘ladder-like’ burials were constructed of flat slabs and chunks of kurkar. ‘Ladder-like’ cist burials are a very distinctive type known elsewhere at only two additional sites in the southern Levant; both definitively Chalcolithic in date. Other cist burials, known from many regions of the southern Levant and from many prehistoric and historic periods, are not relevant comparanda (contra Golani and Nagar 2011:9394) as they bear only passing resemblances to the very distinctive ‘ladder-like’ constructions in this cemetery.14

Several of the Adeimeh cists were found below later burial structures (called ‘tumuli’) that are definitively associated with Early Chalcolithic and possibly earlier artifacts. Cists 32-33 were below Tumulus I, Cists 36-39 below Tumulus II, Cists 98 and 99 below Tumulus III and Cist 155 below Tumulus IV (Stekelis 1935:54, 55, 60, 64) and therefore must belong to an earlier use of the cemetery. An abundance of associated artifacts from the tumuli, primarily pottery (Stekelis 1935: Figs. 19, 20, Pls. IV and V) is documented, albeit published only in a limited number of representative illustrations (Figs. 15, 16). Those artifacts, some of clear chrono-cultural associations, are termini ante quem for the dating of the ‘ladder-like’ cist graves. Notably, none of the Adeimeh grave goods is even similar to EB 1 types, while most parallels for them are dated considerably earlier than LC18 (see below: Excursus). Good ceramic parallels presently known argue well for an association between the pre-LC people occupying Ghassul (Lovell 2001) or contemporary peoples and the Adeimeh cemetery population.

The most compelling parallels to the Barnea cemetery are at another coastal site on a kurkar ridge near Kibbutz Palmahim, c. 30 km north of Ashqelon (Fig. 14:B), where cists and ‘ladder-like’ cist graves, all also devoid of grave goods, were found, some directly below circular stone-built tombs with stone ossuaries15 ascribed to LC by associated pottery (Gorzalczany 2007). Significantly all known ‘ladder-like cist tombs’ are devoid of grave goods. Although the Chalcolithic cemetery is located c. 1 km distant there is no reason to accept Golani’s and Nagar’s (2011:93) suggestion, wholly without basis,16

While a lack of grave goods in all the ladder-like cist burials in the three cemeteries at Barnea, Palmahim and Adeimeh cannot be cited as definitive evidence for a specific dating, it is a clear indication of their association with a singular tradition19 particularly common in early phases of the south Levantine

It merely makes use of materials available at the site, slabs of kurkar and mud for bricks. 14  Rectangular cists are a simple way of delineating graves. While some examples are known from the Chalcolithic period within tumuli at Adeimeh (e.g., Stekelis 1935:51-65) and circular grave structures at Palmahim (e.g., Golani and Nagar 2011: Fig. 7.8), similar arrangements are noted in graves of the Early Bronze Age at Ala Safat in the Jordan Valley (Stekelis 1961), at Deir Ain Abata in the Arava Valley, Jordan (Braun, Moth and Politis 2012) and in tumuli particularly associated with arid zones where they are known from Neolithic times (e.g., Avner 2002:154-155) until well into the second millennium (e.g., Haiman 1992; Yannai et al. 2013) and possibly later. Virtually all those burials (except some in the very arid zones) are distinguished from the Barnea cists by their associations with grave goods. 15  Stone ossuaries are a distinctly Chalcolithic mode of burial (e.g., Perrot and Ladiray 1980: Fig. 115.1; (Fabian, Sheftelowitz and Gilead 2015) and, to the present, unknown in association with EB 1 mortuary contexts. 16  There are no tombs of this type at the Palmahim Quarry site and no Ghassulian pottery was associated with any EB 1 occupation deposits there (Braun 2000a; 2000b). The only Chalcolithic pottery found in situ was in 11 LC cave tombs in the immediate vicinity of the EB 1 13 

settlement (Gophna and Lifshitz 1980). The stratigraphic associations of these tombs is obscure as all were encountered during quarrying, which destroyed any stratigraphic context for them. 17  This refers to the published evidence as the excavator clearly stated that ceramic finds from the tumuli were numerous (Stekelis 1935:45). 18  Stekelis’ (1935:77-78), writing early in the 20th century may be excused for dating finds to the ‘Ghassulian’ period, citing comparanda from Ghassul, as he was clearly unaware the earliest pottery at the site was either early Chalcolithic or possibly even Neolithic (Hennessy 1969; 1982). 19  Exceptions may be in arid zones where artifacts, especially pottery, are encountered in low frequencies.

79

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 15. Pottery and a limestone vessel (n) from tumuli superimposed on ladder-like cist burials at Adeimeh, Jordan. New renderings after Stekelis 1935: Fig. 19.

Chalcolithic (e.g., Smith et al. 2006:337; Rowan and Golden 2009:67), and in stark contrast to EB 1 mortuary traditions of many types including shaft tombs, circular charnel houses at Bab edh Dhra (e.g., Rast and Schaub 1974; Schaub and Rast 1989), narrow troughs, some stone-lined with cobbles and rectangular, semisubterranean charnel houses at en-Naq, Ghor es Safi (Waheeb 1995), dolmen fields (Stekelis 1961; Prag 1995), burial caves and small tumuli (e.g., Deir ‘Ain Abata; Braun, Moth and Politis 2012), which were invariably associated with grave goods, most commonly ceramic vessels (e.g., de Vaux and Steve 1949 passim; Rast and Schaub 1974; 1989 passim). The Date of the Barnea Jar Burials and other Mortuary Evidence from the Ashqelon Cluster Some pot burials encountered within a multi-phased settlement at one locale in Barnea were embedded within walls of buildings, while others were beneath floors associated with habitations (Fig. 17). The

Figure 16. An early Chalcolithic vessel in the Amman Jordan Citadel Museum, probably from the Adeimeh Cemetery (photo by the author; no scale but approximately 15 cm high).

80

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’

Figure 17. A jar burial in Barnea with remains of a wall of kurkar stone just below the modern bulldozed surface; photo by the author, courtesy of A. Golani, Israel Antiquities Authority.

excavators’ argument for their dating is, as for remains described above, tautological. Because those buildings were found just below the surface and in proximity to definitively EB 1 structures, then all elements must be contemporary, because of a purported lack of any default possibility. If there are no Chalcolithic remains then, ipso facto, the pot burials must be non-Chalcolithic and by default EB 1.

al-Ghuzlan in the immediate vicinity of Aqaba, Jordan, where they are described as ‘some burials of children in pithoi’ (Klimscha 2013:53). While the exact dates of those burials are not cited, they should be within the 3950-3550 BCE range suggested for the site (Klimscha 2013:38). This could indicate a possible ascription to the LC-EB 1 transition. A simple burial of a child in a flexed position within an EB 1 house at Tell Iktanu (Prag 1989:37), without grave goods is clearly dated EB 1. However, it appears to have been a unique occurrence as it is without EB 1 comparanda in the southern Levant.

However, given the problematic nature of the stratigraphic paradigm (see above), that ascription may also be called into question, especially in light of parallel evidence of mortuary customs, which elsewhere in the southern Levant are virtually unknown in EB 1. Most glaringly, they are absent at nearby Afridar, which is dated by Golani et al. to the earliest EB 1. That hiatus of several centuries in the purported LC to EB 1 continuity paradigm is inexplicable and indicative of its incorrectness.

A lone infant burial in Area M at Afridar (Golani 2004:48), beneath a large piece of pottery, is representative of a quite different tradition, as are two partial burials in jar fragments from Area G (Braun and Gophna 2004:198199). In those instances the burials were well below the levels of EB 1 surfaces. Similar type burials are found at many Chalcolithic sites. Some of them are earlier than LC (e.g., Macdonald 1932:7; Koeppel et al. 1940: Pl. 40; Mallon, Koeppel and Neuville 1934: Pls. 24-25; Philip and Baird 1993:17; Blackham 1999:37 with references; Lovell et al. 1997:362; Lovell 2010:113-115; Nagar and Eshed 2001; Smith et al. 2006:331; Golani and Nagar 2011: Table 7.1 with citations; Eshed and Bar 2012), while none is known in an EB 1 context.

Burials within EB 1 occupations are extremely rare, and they may suggest some continuity from earlier traditions, as they are invariably associated with early EB 1. Pot burials are usually associated with pre LC occupations, although there are some rare exceptions. One is from a very early EB 1 context at Modi’in (E. C. M. van den Brink personal communication) on the ‘Plateau’, where it was found above an LC occupation, which may well date to the LC-EB 1 transition. Another was found below a floor at Yesodot (see above), and appears to belong to the same chrono-cultural horizon. Likely additional examples are from the site of Hujayrat

The jar burials at Barnea are also quite different from the relatively few EB 1 jar burials in the southern Levant discovered to date. The sole instance of an EB 1 jar burial in the immediate region of Barnea is an 81

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead adult in a pithos found at nearby Nizzanim (Yekutieli and Gophna 1994), where it is ascribed to early EB 1 (the excavators’ EB 1A). Additional evidence for infant jar burials at sites in the southern Levant in EB 1 is limited to more northerly locales. Three are known from Tel Te’o (Kahila Bar-Gal and Smith 2001:165), at least three from Kabri (Scheftelowitz 2002:28-29) and one from ‘En Esur (Yannai 2006:59, Fig. 4.73:3), all in holemouth vessels. Additional examples from Bet Yerah (Greenberg and Paz 2004:9-10, Fig. 8) derive from an early phase of Early Bronze Age 1 (the excavators’ EB 1A, Period A, Unit 9). Notably, the typology of at least one of the jars with its broad, widely splayed strap handles (Greenberg and Paz 2004: Fig. 8) is reminiscent of pre-EB 1 vessels. Its origin and those of a similar vessel (Greenberg et al. 2006: Fig. 4.3:9) from an early excavation with relatively limited stratigraphic control (Greenberg and Paz 2004; Braun 2008; Braun 2012b) could admit of the possibility of a pre-EB 1 date for it.

situ in clear EB 1 contexts, do not constitute definitive evidence of their continued use in EB 1. However, it is not unlikely that some of the materials recovered could date to the LC-EB 1 transition (Roux, van den Brink and Shalev 2013). Ground Stone Artifacts Fragments of basalt vessels (Fig. 20), products of a well-developed LC industry, which differ significantly from EB 1 types (Braun 1990), are additional evidence of residual artifacts in the Ashqelon Cluster. The most notable type is the fenestrated, pedestaled bowl (e.g., Fig. 20:6, 8, 9; Braun and Gophna 2004: Fig. 24:1; Golani 2008: Fig. 11:3) incised with rectangles or chevrons. In EB 1 that type of incised decoration is unknown,20 while the basic morphology of the vessel was apparently copied, but only in ceramics, most of which are associated with northern sites. Similarly, straight-sided bowls with flat bases, of basalt and sometimes dark gray phosphorite, bearing this same type of incised decoration are also found at numerous Chalcolithic sites (e.g., Fig. 20:2, 4, 7; Scheftelowitz 2004: Figs. 4.3, 4.7) where they are correctly considered hallmarks of the horizon. Additional examples, including pedestaled types, have been found at several of the Ashqelon sites (e.g., Rowan 2004: Fig. 4: Braun and Gophna 2004: Fig. 24:1; Golani 2008: Fig. 11:3; Rosenberg and Golani 2012:35). Haematite maceheads from Area F are likely additional evidence of a Chalcolithic occupation as noted by Sebbane (2009:152), whose study indicates that such objects have only been found in situ in Chalcolithic contexts. Once again, there is more than a chance that some of these objects could derive from an LC to EB 1 transition.

Thus, the ascription of infant jar burials at Barnea to EB 1, and especially to late in the period, for which there are no known parallels, is highly questionable. The practice is an exponent of a long-lived tradition of burying infants in jars or in portions of them that in the southern Levant began in the Pottery Neolithic period, continued through Chalcolithic times (e.g., Philip and Baird 1993:17) and was only rarely practiced early in EB 1. It would have been an extraordinarily parochial practice, unique at Barnea, if, after a hiatus of several centuries when Afridar was occupied and the custom fell into disuse, that it was reinstituted at Barnea as an isolated phenomenon for c. half a millennium. A more reasonable interpretation suggests that the practice was likely related to a Chalcolithic occupation, or, if the Modi’in and Yesodot parallels are indicative, to an LC – EB 1 transition, possibly indicated by some of the radiocarbon dates from the Ashqelon Cluster. In one sense, however, Barnea appears to be unique. At least for the present, it has the largest concentration of late prehistoric infant jar burials in the southern Levant.

A ‘violin’ type figurine (Rosenberg and Golani 2012:35) from the Barnea locale, although apparently found in an EB 1 context, is clearly earlier in date. It is more than likely, considering the obvious date of the cist cemetery, that it is a residual object derived from a relatively early Chalcolithic occupation. Such figurines are virtually unknown in the Beersheva cluster of LC sites, but they have been recovered in considerable quantities at the earlier Chalcolithic site of Gilat (Commenge et al. 2006; Gilead and Fabian 2010:*96-*97; Gilead 2011).

The Evidence of Ceramic Artifacts from the Ashqelon Cluster At Ashqelon and other southern coastal sites, high concentrations of salt and other factors have been extremely detrimental to the preservation of ceramics, especially those of late prehistoric horizons. That has ensured that only a few complete vessels have been recovered from excavations in the Ashqelon Cluster. Because of similarities in LC and early EB 1 forms it is often impossible to definitively distinguish between fragmentary pots of these horizons (Figures 18-19) (Braun 2000a). Thus, ceramic remains of a poorly preserved occupation of the Chalcolithic period may be obscured or invisible to some excavators, while fragments of cornets and churns from purportedly EB 1 deposits, pottery types that have never been found in

Chipped Stone Artifacts Bifacial and other types of flint tools found at the Ashqelon sites, unknown in EB 1 contexts, are additional evidence of activity in the Chalcolithic period (e.g., Khalaily 2004:147, Fig. 22:5; Khailaly 2008: Fig. 14:4; Zbenovich 2004a:65-66; Zbenovich 2004b:266). Although an EB 1 basalt bowl industry indicates a modicum of continuity from LC, there are significant typological changes and in the discontinued use of incised decoration that distinguish its products from those of Chalcolithic predecessors (Braun 1990). 20 

82

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’

Figure 18. Select Pottery from the Ashqelon Cluster and Chalcolithic Sites: 1. Straight-sided bowl (so-called ‘v-shaped’) from Afridar, Area J, after Baumgarten 2004: Fig. 10:1. 2. Straight-sided bowl (so-called ‘v-shaped’) from Afridar, Area J, after Baumgarten 2004: Fig. 10:2. 3. Straight-sided bowl (so-called ‘v-shaped’) from Safadi, after Commenge-Pellerin 1990: Fig. 18:3. 4. Bowl with ‘pie-crust’ decoration on rim from Safadi, after Commenge-Pellerin 1990: Fig. 24:1. 5. Bowl with ‘pie-crust’ decoration on rim from Safadi, after Commenge-Pellerin 1990: Fig. 24:4. 6. Bowl with ‘pie-crust’ decoration on rim from Safadi, after Commenge-Pellerin 1990: Fig. 24:6. 7. Bowl with ‘pie-crust’ decoration on rim from Afridar Area F, after Baumgarten 2004: Fig. 12:6. 8. Bowl with ‘pie-crust’ decoration on rim from Afridar Area F, after Baumgarten 2004: Fig. 12:2. 9. Bowl with ‘pie-crust’ decoration on rim from Afridar Area F, after Baumgarten 2004: Fig. 12:4. 10. Bowl with ‘pie-crust’ decoration on rim from Gilat, after Levy 2006: Plate 10.8:3. 11. Bowl with ‘pie-crust’ decoration on rim from Gilat, after Levy 2006: Plate 10.8:4. 12. Bowl with ‘pie-crust’ decoration on rim from Gilat, after Levy 2006: Plate 10.8:2. 13. Cornet fragment from Area J Afridar, after Baumgarten 2004: Fig. 13:1. 14. Cornet fragment from Area J Afridar, after Baumgarten 2004: Fig. 13:2. 15. Cornet fragment from Area J Afridar, after Baumgarten 2004: Fig. 13:3. 16. Cornet fragment from Area J Afridar, after Baumgarten 2004: Fig. 13:4. 17. Cornet fragment from Area E Afridar, after Golani 2004: Fig. 30:1. 18. Cornet fragment from Area E Afridar, after Golani 2004: Fig. 30:1. 19. Cornet fragment from Area E Afridar, after Golani 2004: Fig. 30:1. 20. Cornet fragment from Area E Afridar, after Golani 2004: Fig. 30:1. 21. Cornet fragment from Area E Afridar, after Golani 2004: Fig. 30:2. 22. Cornet fragment from Chalcolithic Grar, after Gilead et al. 1995: Fig. 4.9: 6. 23. Cornet fragment from Chalcolithic Grar, after Gilead et al. 1995: Fig. 4.9:16. 24. Cornet fragment from Chalcolithic Grar, after Gilead et al. 1995: Fig. 4.9:17. 25. Cornet fragment from Chalcolithic Grar, after Gilead et al. 1995: Fig. 4.9:8.

83

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 19. Select Pottery from the Ashqelon Cluster and Chalcolithic Sites: 1. Jar rim from Afridar Area J, after Baumgarten 2004: Fig. 16:3. 2. Jar rim from Afridar Area J, after Baumgarten 2004: Fig. 10:7. 3. Incised, decorated jar neck from Afridar, Area J, after Baumgarten 2004: Fig. 10:6. 4. Incised, decorated jar neck from Abu Matar, after Commenge-Pellerin 1987: Fig. 36:8. 5. Jar neck from Afridar, Area J, after Baumgarten 2004: Fig. 16:3. 6. Jar neck from Afridar, Area J, after Baumgarten 2004: Fig. 10:7. 7. Chalcolithic jar rim from Abu Matar, after Commenge-Pellerin 1987: Fig. 33:6. 8. Chalcolithic jar rim from Grar, after Gilead et al. 1995: Fig. 4.15:3. 9. Chalcolithic vessel from Bir Safadi, after Commenge-Pellerin 1990: Fig. 37:6. 10. Oval-shaped ceramic vessel (churn?) from Afridar, Area J, after Baumgarten 2004: Fig. 10:15. 11. Churn fragment from Teleilat Ghassul (‘Middle Chalcolithic’), after Lovell 2001: Fig. 4.23.

Metallurgy

and Kamenski 2004: Table 1, Areas G, H:6), made in the lost wax (cire perdue) technique, is almost certainly of Chalcolithic origin, as for the present there is no evidence whatsoever of objects from EB 1 contexts made in that fashion (Sebbane 2014:116; Sariel Shalev personal communication).

Much has been made of the copper-associated industry at Afridar (Golani 2004; 2014), for which no definitive date has been forthcoming.21 Radiocarbon samples from his Afridar excavations may well be associated with LC or LC – EB 1 occupations, but they have no definitive associations with specific archaeological deposits. Significantly, Segal, Halicz and Kamenski (2004:324) indicate that: ‘... the frequency distribution of several elements in the copper objects from Ashqelon Afridar corresponds to those of the Chalcolithic period rather than the EB 1’. A notable artifact from Ashqelon, a single piece of a copper tube from Area G (Segal, Halicz

An Enhanced Stratigraphic Paradigm for the Ashqelon Cluster Occupation Sequence Golani’s (2014:120) and colleagues’ global, four stratum stratigraphic paradigm for the Ashqelon Cluster encompassing the entirety of EB 1 (including an implied LC to EB 1 transition indicated by their ‘continuity’ argument), which they claim lasted ca. 1000 years, lacks corroborative evidence for the radical redefinition of the archaeological record it proposes. According to that

The evidence presented is merely noted as dating to EB 1 with no specific time span within the c. 800 year long period. 21 

84

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’

Figure 20. Chalcolithic ground stone vessel fragments from the Ashqelon Cluster and Grar: 1. From Afridar Area E, after Rowan 2004: Fig. 4:3. 2. From Grar, after Gilead et al. 1995: Fig. 7.2:9. 3. From Afridar Area E, after Rowan 2004: Fig. 4:2. 4. From Grar, after Gilead et al. 1995: Fig. 7.2:10. 5. From Afridar Area E, after Rowan 2004: Fig. 4:1. 6. From Afridar Area E, after Rowan 2004: Fig. 6:2. 7. From Grar, after Gilead et al. 1995: Fig. 7.2:4. 8. From Afridar Area E, after Rowan 2004: Fig. 6:1. 9. From Afridar Area G, after Braun and Gophna 2004: Fig. 24:1.

paradigm the earliest EB 1 occupations represented at the Afridar sites would be equivalent to a Stratum V at Barnea (if it were to have existed), while the latest phases, beginning with ‘late EB 1A’ and ending with ‘late EB 1B’, are Strata IV-II at Barnea. That paradigm conflates a purported sequence of an entire millennium into a continuum of only three strata, which ignores distinctions between the material cultures of different phases of the Chalcolithic and EB 1 periods. It further ignores at least one radiocarbon date (Fig. 2: RT-2634) that suggests human activity at the site as early as the 5th millennium.

to the end of EB 1, albeit at different locales. Whether occupation of the cluster during that long time span was continuous remains obscure, as the record as understood to date has too many gaps for such a determination. Radiocarbon evidence (Fig. 2) suggests the high likelihood of occupation along the littoral for long periods, including activity likely to be attributed to the LC-EB 1 transition. Remains of that period could be the earliest occupations in Area G, perhaps those below Stratum II and possibly even Stratum II itself, which according to the evidence of architecture, could well be one of the earliest manifestations of EB 1 curvilinear house types, perhaps signaling the end of the transition. Other possible candidates for this period could be houses in the Barnea locale and their associated pot burials, which do not appear to be paralleled in developed and late EB 1. However, they could also date to some earlier, purely Chalcolithic period, which is not unknown in the region.

A more realistic evaluation, based on available evidence from the presently excavated, and very incomplete archaeological record of the Ashqelon Cluster, suggests there was intermittent activity (possibly with real chronological gaps) on the littoral, probably from relatively early Chalcolithic times (e.g., the cist cemetery and the ‘violin-shaped’ figurine) into the LC (as at the Agamim site), and EB 1 periods, and possibly 85

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Evidence from Hujayrat al-Ghuzlan and Tell Magass

Should the impressive metallurgical associations of this site prove to date to the LC-EB 1 transition, then the site would appear to be something of an anomaly, in that it may have reached its floruit when most sites in the more fertile zones of the southern Levant appear to have been either diminished or even abandoned. It remains for further research to attempt to discern the place of this interesting site within the mosaic of the south Levantine LC-EB 1 transition.

These arid zone sites c. 4 km north of Aqaba in Jordan are notable for their well-preserved architecture and copious evidence of copper production. A range of radiocarbon dates (Klimscha 2009) make it likely that some of the early phases of the sites’ occupations may date to the LC period and also to the LC-EB 1 transition. Unfortunately for the present study, the material culture of those sites, primarily represented by pottery assemblages (Kerner 2009), belongs to a cultural sphere that is apparently related only tangentially to that of the ‘normative’ southern Levant. While pots from Hujayrat al-Ghuzlan share basic forms and some aspects of decoration with generic Chalcolithic and EB 1 types, because they are substantially different and notably much cruder, they cannot, with any confidence, be placed within a linear, chronological sequence of late prehistoric ceramics of the southern Levant. Therefore, the dates of the occupation of this site are based solely on radiocarbon data.

Evidence from Jebel Mutawwaq (Wadi Zarqa, Jordan) Although this site has been dated to EB 1 on the basis of its specialized architecture, simple houses of sausageshape (i.e., structures with parallel walls ending in regularly curving apses), published pottery types (Fig. 21) suggest significant affinities with earlier periods. A squat jar with broad shoulders and narrow neck, of indeterminate size (Fig. 21:1), morphologically similar to some vessels of the Chalcolithic horizon (e.g., Koeppel et al. 1940: Pl. 96:4 upper, 2 lower), is decorated with a horizontal line of a type of incisions often associated with pottery of the Wadi Rabah Period (Khalaily and Kamaisky 2002). Another jar (Fig. 21:6) with a wide, flat, hooked knob (not a ledge handle) is adorned with lines of rope-like decoration arranged as if they were used for suspending the vessel. That type of decoration, so far as this writer is aware, is unknown in EB 1 contexts, but has parallels at Ghassul (e.g., Mallon et al. 1934: Fig. 50:XXIV, Fig. 52:III). A jar (Fig. 21:5) with wide aperture and splayed rim has similar types of hooked knobs at its waist, possibly

Notably, a number of infants buried in what are described as ‘pithoi’ are also associated with the site. It is uncertain whether these burials are analogous to those at Modi’in, Yesodot and Barnea as no details are available to compare them. If the burial vessels were complete and oversized types generally described as ‘pithoi’, then they are, for the present, reminiscent of the Nizzanim burial, which, however, was of an adult.

Figure 21. Select Pottery Types from Jebel Mutawwaq: 1. After PAJM photo posted 2014: 7/4/14. 2. After Álvarez et al. 2014: Fig. 11. 3. After PAJM photo posted 2014: 10/3/14. 4. After Álvarez et al. 2014: Fig. 12. 5. After Álvarez et al. 2014: Fig. 11. 6. After Álvarez et al. 2014: Fig. 12.

86

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’ intended for suspending the vessel. It also has a short, apparently rope-like appliqué placed horizontally at the level of the knobs. Another jar with a smaller opening (Fig 21:4) also has short segments of rope-like decoration obliquely positioned at its waist. In form it is paralleled by a similarly decorated jar from Ghassul (Garfinkel 1999: Photo 125). Three similar segments of incised decoration, also obliquely placed, are found on the neck of a jug with a loop handle attached to its rim (Fig. 21:3). Rope-like decoration of this type, made by rows of shallow, circular (finger-made?) impressions is commonly found in pottery types from Ghassul (e.g., Mallon et al. 1934 Fig. 41:8,12, Fig. 62), and it is known also at Hujayrat al Ghuzlan (Kerner 2009: Figures 1:2, 2:11, 4:8,10). It contrasts with early EB 1 modes of ropelike decorations made by obliquely slashing raised lines of clay. Only the jug (Fig. 21:3) with its high loop handle and a crude, cup-like bowl with small loop handle (PAJM 2014:25/4/2014: no. 2) from the site seem closer to EB 1 ceramic types.

be dated to very late LC or the LC-EB 1 transition (Bar and Winter 2010), in keeping with the radiocarbon dating. The Radiocarbon Question Radiocarbon dates place the end of the Chalcolithic sometime early in the fourth millennium BC and the onset of EB 1 sometime around the middle of that millennium (Braun et al. 2013). It seems pretty clear that there is a zone of transition between LC and what has formerly been designated as early EB 1, which I suggest labeling LC-EB 1. The evidence for Modi’in and Hujayrat al-Ghuzlan suggests those sites were occupied during that transition period. However, the nature of such occupation at both sites remains somewhat obscure as exposures were in the former instance rather limited and in the later, somewhat hard to discern from later occupations. Some radiocarbon dates (Fig. 2) from the Ashqelon Cluster suggest human activity there, and there is no reason to question their inherent veracity. However, their associations with specific deposits remain unclear. Dates that appear to coincide with the transition period are marked in bold font; others that indicate activity in the Chalcolithic period are in italics. Some, in the 2 δ range may be dated to one or the other period, and accordingly they appear in bold, italic font.

Additional suggestions of the early dating of some pottery types from Jebel Mutawwaq are found in bases with mat impressions (PAJM 2014: Posted 25/4/2014: no. 4) unknown in good EB 1 contexts, but common in pre-EB 1 potting traditions (e.g., Mallon et al. 1934:91, Pl. 39; Koeppel et al. 1940: Pls. 83-84). The latest of those at Hujayrat al Ghuzlan may date to the LC-EB 1 period (Kerner 2009: Fig. 1:4, 5). Additional vessels from Jebel Mutawwaq include a spoon (PAJM 2014:26/4/2014) similar to many found at Ghassul (e.g., Mallon et al. 1934: Pl. 44:55-58,64), a small fragment of a jar with two vertical lugs (FernandezTresguerres 1992: Fig. 3:12) common on Chalcolithic type vessels (e.g. Mallon 1934 et al.: Pl. 43:34-36) and a vessel with snake decoration (Fernandez-Tresguerres 2005: Fig. 6) similar to others from Ghassul (Mallon 1934 et al.: Fig. 61:9,10,11,13). Possibly the lack of LC pottery types such as churns and cornets is indicative of an LC-EB 1 date for some phases of occupation at the site. However, information presently available is too scarce for determining what the extent of such a settlement might have been.

Identifying the LC to EB 1 transition period After nearly a century of archaeological exploration in the southern Levant, there still remains a surprising dearth of evidence for an LC to EB 1 transition, although it has been somewhat ameliorated by information from the sites discussed above. There are a number of reasons for the paucity of material. Some may be related to how archaeologists perceive and define transitional phases between chrono-cultural facies, while others are in the very nature of the archaeological record and interpretations derived from it. The Problem of Identifying the LC-EB 1 Transition

A Cave Burial

While major distinctions between artifacts of the Chalcolithic and EB 1 are well known and documented in the literature, it is often far more difficult to definitively identify evidence of continuity and the kinds of change that herald the onset of a transition period and emergence into a new cultural facies. At sites where more than a single occupation existed, when two sequential chrono-cultural facies are represented, separating one from the other can be especially difficult. In my experience there are almost never contexts that may be described as ‘clean loci’ or other chronologically discrete archaeological deposits. Artifacts tend to move up and down due to human,

A singular burial of an adult male in a rock shelter in the Judean Desert near Jericho (Schick 1998) was devoid of normative LC or EB 1 grave goods. Accompanying the male skeleton was an extraordinarily well-preserved assemblage of objects of organic materials, some of which were subjected to radiocarbon dating (Hull et al. 1998). Several dates ascribed to them admit of the possibility of this burial having occurred during the LC to EB 1 transition. However, a lack of comparanda for such objects does not allow for an ascription of the burial to either the LC or EB 1 period. A possible exception is a canaanean flint blade, a type that could 87

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead and animal activity and the effects of the elements on deposits.

after a well-organized, sedentary LC society came to an end at many sites, probably sometime early in the fourth millennium. If, as the radiocarbon data appear to indicate, the period of transition to EB 1 lasted for several centuries, then it seems likely that human society experienced some major rupture that brought about substantial alterations in settlement patterns and ushered in a period of significantly changed material culture. In the case of ceramic technology and style, the Modi’in example and the continuity in forms in the southern region into early EB 1 indicate a somewhat gradual transformation. The situation in the northern region, however, remains obscure as LC there is virtually unrecognized.

By contrast, the single period occupation at Yesodot, and the superimposed sequence of relatively discrete deposits at Modi’in Deep Deposits, offer rare opportunities for identifying transitional phases. The former due to its discreteness; the latter because of its in situ architecture and an intense study of stratified ceramic assemblages of significant size that filter out or diminish ‘chonological noise’. Roux’s and colleagues’ work on the pottery from Modi’in offers remarkable insights into details of the kinds of slow change that can occur in one area of technology over a period of c. two centuries. As such it is a fascinating and important study of gradual trajectory occurring over time. It further suggests, by analogy, that a hiatus in mortuary practices of several centuries as that suggested for the Ashqelon Cluster in Afridar is hardly tenable.

There is also evidence of some continuity in mortuary traditions in the occasional examples of individual jar burials, with Ashqelon, Barnea, possibly an extreme example of this in the south. In the north jar burials at Tel Te’o, possibly at Kabri and at Beth Yerah, may suggest continuity in that region. However, the tradition seems to have been rather weak, and it soon waned. The use of caves for multiple burials over time was a popular tradition that apparently carried over from Chalcolithic times into EB 1.

A Lean Archaeological Record Nearly a century of intensive archaeological research in the southern Levant has failed to unearth more than a handful of sites that have yielded evidence of an LC-EB 1 transition. That is no coincidence as few sites offer real evidence of continuity. That indicates a major break in human occupation of sites and a reorganization of human activity. Some recent studies on ancient climates suggest a possible reason for such a rupture; climatic changes may have ushered in a period of significantly increased aridity as occurred in the Holocene from time to time. One such period seems to have coincided with the LC-EB 1 transition (Bar-Matthews et al. 2003: 3196; Verheydena et al. 2008; Dusar et al. 2011:141 with references; Finné et al. 2011:3162; Clarke et al. 2016). That might explain why such major centers as Ghassul and the LC communities of the northern Negev, today located in rather arid zones, were abandoned.

The tradition of copper working was also transmitted from the Chalcolithic period, although some of the finer aspects such as the ‘cire perdue’ method as was a penchant for creating specialized ‘cultic’ objects, were abandoned. Copper was smelted and worked, apparently in sizable quantities at the sites of Tell Magass and Hujayrat al-Ghuzlan (Klimscha 2013), and worked, perhaps at Ashqelon (Golani 2004; 2014), although precisely in which time span, remains unclear. The reason for such radical changes in human society remain obscure. Did populations shrink along with resources to such an extent that they left such a poor archaeological record for that time span? Was some catastrophic climatic event such as a prolonged drought, likely compounded by starvation, disease and increased competition by populations for available resources responsible for this rupture in the social fabric? Were some population elements driven to arid zone sites such as Feinan, Tell Magass and Hujayrat al-Ghuzlan, where they could profit from mining and processing of copper? And if so, who were the consumers of their yields? Were Chalcolithic immigrants at Buto I in the Nile Delta (Faltings 1998) refugees from such a disaster that befell the southern Levant?

Did that arid period have catastrophic results that severely affected south Levantine populations and cause them to shrink, with a concomitant lessening of archaeological deposits? Such a disruptive event could explain, at least in part, the reason for limited continuity of human activity, known primarily from a sparse archaeological record in the more northerly zones blessed by abundant rainfall and perennial water sources. Summary

Answers to these and other questions on why the archaeological record of the LC-EB 1 transition is so very lean in the early fourth millennium are not yet forthcoming from our present understanding of the archaeological record. Too much remains obscure. We

The present review briefly summarizes available information on the LC-EB 1 transition. No matter how scholars label the period, what is eminently clear is that there is a dearth of evidence for human activity 88

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’ can, however, use what evidence is available to help us suggest what are likely to be fruitful lines of inquiry for future research, which hopefully will yield new light on the ‘lost horizon’. ••

Excursus: Comparanda for the Pottery and Stone vessel from the Adeimeh Cemetery Tumuli •• Cornets (Fig. 15:b) are well attested at Ghassul (e.g., Mallon, Koeppel and Neuville 1934: Pls. 4748; Koeppel et al. 1940: Pl. 95; Hennessy 1969: Pl. 8A) and other Chalcolithic sites such as, Safadi (Commenge-Pellerin 1987: Figs. 37,38; 1990: Figs. 36:1-5,1-11, 55-56) and Modi’in in Chalcolithic Strata 6 and 5 (Roux, van den Brink and Shalev 2013; van den Brink 2013). They are notably absent in the final Chalcolithic phase of Modi’in Stratum 4 and the initial EB 1 phase in Stratum 3. Cornets also are known from relatively early contexts in Wady Ghazzeh (Macdonald 1932: Pl. XXXV), Ghassul (Lovell 2001: Fig. 4.27) and Gilat (Levy 2006; Commenge, Levy and Kansa 2006: Pl. 10.4), while they have not yet been shown to derive from unimpeachable EB 1 contexts.22 •• Narrow-footed goblets (Fig 15:d) are another type also found at Ghassul (e.g., Mallon, Koeppel and Neuville 1934: Pl. 49, Fig. 56:12; North 1961: Fig. 13:8630, 8457, 8663: Hennessy 1969: Fig. 6:9; Lovell 2001: Figs. 4.27, 4.42:2-3) from as early as ‘Middle Chalcolithic’ contexts. High-footed vessels without fenestration (e.g., Stekelis 1935: Fig. 19:g, h, l, m, o, q, r, s, Pl. IV:1-4) are also found at Ghassul (Mallon, Koeppel and Neuville 1934: Fig. 56:7,9,10,11,13-17; North 1961: Figs. 13:8723, 14:8607 and possibly Lovell 2001: Fig. 4.42:4,6). •• Two large, shallow bowls with broad trumpet bases or (Fig. 15:o, s) are paralleled in Late Neolithic (sic!) forms from Jericho (Garfinkel 1999: Fig. 44:9). Several fragments of this type are also known from Ghassul (Mallon et al. 1934: Fig. 36:4, 9-11, 13-17; Koeppel et al. 1940: Pl. 79:2, 3), some of which were fenestrated. Others have been found in Neolithic and Chalcolithic levels at Horbat ‘Uza (Getzov et al. 2009: Figures 2.15:5, 14-15, 2.32:7-8), in the Wadi Ziqlab (Banning, Blackham and Lasby 1998: Fig. 8:21) and Gilat (Commenge, Levy and Kansa 2006: Pls. 10.9:5-8, 10.10:4, 5, 9). Several from Adeimeh are notable for their incised decoration. •• The morphology of small jars or goblets with wide mouths and bulbous bodies (Fig. 15:k, p, t) is attested in Chalcolithic contexts, albeit with different modes of decoration and with pierced handles (e.g., Koeppel et al. 1940: Pl. 78:3,6,8). The

••

••

••

Recovered are only tiny fragments of thick, durable bases of cornets in mixed LC and EB 1 contexts (e.g., Baumgarten 2004: Fig. 10: 16; Braun and Gophna 2004: Fig. 21:1; Golani 2004: Fig. 30:1-2), likely residual from earlier deposits. 22 

89

rope-like decoration on some of the Adeimeh specimens seems to be unusual as are the small ledge-like appurtenances, which are known but not common on Chalcolithic vessels (e.g., Koeppel et al. 1940: Pl. 76:9). Goblets on high, trumpet bases (Fig 15:g, h, l, m, q, r; Fig 15) appear to be parochial types associated with the Adeimeh cemetery. However, their overall morphologies as well as decorative features of some, horizontal, pierced lug handles and incised decorations are widely paralleled at Ghassul (e.g., Mallon, Koeppel and Neuville 1934: Pl. 43:34,35, Pl. 49:95,96; Koeppel 1940: Pl. 78:1,3,4,6-8) and in other pre-EB 1 ceramic assemblages (see also below: appurtenances and decorative devices). Simple cups and bowls with flat bases (Fig. 14:e, f and j) are common to many periods and therefore difficult to assign to any particular cultural horizon, but one deep, cup-like vessel is characteristic of Chalcolithic assemblages (e.g., Bourke 2001: Fig. 4.12:9-10). Additional evidence of the relatively early date of this assemblage is in mat impressions on bases of three published bowls (Stekelis 1935: Pl. V:2, 5, 13). Mat-impressed bases (Stekelis 1935: Pl. V:2, 5, 13) have their roots in the Neolithic period and the tradition continues into the Chalcolithic era. Such bases are not uncommon at Ghassul (e.g., Mallon, Koeppel and Neuville 1934: Pl. 39:1, 2, 4; North 1960: Pl. XI:8705, 8527, 8658) where they seem to be associated with relatively early phases (ibid.: 91-91; Hennessy 1982:57; Lovell 2001: Fig. 4.30). Significantly, they are virtually unknown in the Beersheva Cluster of LC sites but are not uncommon in Late Neolithic and relatively early Chalcolithic contexts (e.g., Vailant 1988: Fig. 78; Garfinkel 1999: Photos 59, 69, Fig. 122:12; Lovell 2001: Figures 4.29:5, 4.30:3). Their latest appearance at Hujayrat al-Ghuzlan and Jebel Mutawwaq could suggest they date to the LC-EB 1 transition (see below). Appurtenances and decorative devices (Fig. 15: passim) on pottery from Adeimeh similarly suggest pre-LC dating. They include small, pierced, horizontal ledges (Stekelis 1935: Fig. 17), types found at Ghassul in relatively early levels (e.g., Mallon, Koeppel and Neuville 1934: Figures 41:5, 50:1, 59:, Pls. 41:28, 43:34,35, 47:81,82,48:83 49:95,96). Incised decorations (Stekelis 1935: Fig. 19:m, r, Pl. V:6) are also common at Chalcolithic sites including Ghassul (e.g., Mallon, Koeppel and Neuville 1934: Figures 60,62; Koeppel et al. 1940:79, Pl. 88), Gilat (e.g., Commenge, Levy and Kansa 2006: Pl. 10.2:1 Pl. 10.30), Abu Hamid (Dollfus and Kafafi 1993: Fig. 2:1). They are even associated with sites of the Raba horizon (Garfinkel 1999: Photos 76-78) such as Munhata

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead (Garfinkel 1992: Fig. 138:8-18), Wadi Ziqlab Site 200 (Banning et al. 1996: Fig. 6:33) and Horbat ‘Uza (Stratum 19; Getzov et al. 2009: Fig. 2.22). •• A shallow high, solid-pedestaled limestone bowl (Fig. 15:n)23 is of a generic type of stone vessel common in Neolithic and early Chalcolithic assemblages. A possible fragment of one is known from Ghassul (e.g., Mallon, Koeppel and Neuville 1934), while others derive from Neolithic and early Chalcolithic contexts (e.g., Fig. 36:7; Braun 1997:129 with references). •• I was unable to find good parallels for a small, two-handled jar with slightly flaring rim and coarse, rope-like decoration in high relief. A generic parallel is from an early occupation at Qatif (Epstein 1984: Fig. 3:1). The style of its decoration is found on a Late Neolithic or Early Chalcolithic vessel from Beth Shan (Braun 2004: Fig. 3.28).

in Wadi Ziqlab. Annual of the Department of Antiquities of Jordan XLII:141-159. Bar, S. 2014. The Dawn of the Bronze Age: The Pattern of Settlement in the Lower Jordan Valley and the Desert Fringes of Samaria during the Chalcolithic Period and Early Bronze Age I. Brill: Leiden, Boston. Bar, S., Cohen-Klonymus, H., Pinsky, S., Bar-Oz, G. and Shalvi, G. 2015. Fazael 5: Soundings in a Chalcolithic Site in the Jordan Valley. Mitequfat Haeven: Journal of The Israel Prehistoric Society 45:193-216. Bar, S., and Winter, H. 2010. Canaanean Flint Blades in Chalcolithic Context and the Possible Onset of the Transition to the Early Bronze Age:  A Case Study from Fazael 2. Tel Aviv 37/1:33-47. Bar-Matthews, M., Ayalon, A., Gilmour, M., Matthews, A. and Hawkesworth, C. J. 2003. Sea-land oxygen isotopic relationships from planktonic foraminifera and speleothems in the Eastern Mediterranean Region and their implication for paleorainfall during interglacial intervals. Geochimic et Cosmochimic Acta 67/17:3181-3199. Bar-Yosef, O. 1992. Chapter 2: the Neolithic Period. In A. Ben-Tor (ed.), The Archaeology of Ancient Israel: 10-39. New Haven and London: Yale University Press and Open University of Israel. Baumgarten, Y. 2004. Excavations at Afridar--Area J. ‘Atiqot 45:161-184. Blackham, M. 1999. Tulaylat Ghassul: An appraisal of Robert North’s excavations (1959-1960). Levant 31:19-64. Blackham, M. 2002. Modeling Time and Transition in Prehistory: The Jordan Valley Chalcolithic (5500-3500 BC) [British Archaeological Reports International Series 1027]. Oxford: British Archaeological Reports. Boaretto, E. 2008. Radiocarbon dating (in Golani 2008). ‘Atiqot 60:45-46. Bourke, S. J. 2001. Chapter 4: The Chalcolithic Period. In B. Macdonald, R. Adams and P. Bienkowski (eds), The Archaeology of Jordan: 107-162 (Levantine Archaeology 1). Sheffield: Sheffield Academic Press. Bourke, S. J. and Lovell, J. 2004. Débats et colloques: Ghassul, chronology and cultural sequencing. Paléorient 30/1:179-182. Bourke, S., Lovell, J., Seaton, P., Sparks, R., Mairs, L. and Meadows, J. 2007. A fourth season of renewed excavation by the University of Sydney at Tulaylat Al-Ghassul (1999). Annual of the Department of Antiquities of Jordan 51:35-80. Bourke, S. J., Lawson, E., Lovell, J., Hua, Q., Zoppi, U. and Barbetti, M. 2001. The chronology of the Ghassulian Chalcolithic Period in the Southern Levant: new 14 C determinations from Teleilat Ghassul, Jordan. Radiocarbon 3/3:1204-1217. Bourke, S., Zoppi, U., Meadows, J., Hua, Q., and Gibbins, S. 2004. The end of the Chalcolithic Period in the South Jordan Valley: new 14C determinations from Teleilat Ghassul, Jordan. Radiocarbon 46/1:315323.

References Abadi-Reiss, Y. and Varga, D. 2018. Lecture: Preliminary Report on the Chalcolithic Site of Agamim Ashqelon (Presented at the Annual Meeting of the Israel Prehistory Society, 28/11/2018). Álvarez, J. R. M., Álvarez Martínez, V., Polcaro, A. and Zambruno, P. S., Polcaro, A. and Zambruno, P. 2014. Jebel Mutawwaq. Veinte años de investigación española en Jordania. Anejos de Nailos: Estudios Interdisciplinares de Arqueología 1:63-91. Amiran, R. 1963. The Ancient Pottery of Eretz Yisrael. Jerusalem: Bialik Institute (Hebrew). Amiran, R. 1969. Ancient Pottery of the Holy Land. Jerusalem: Massada. Amiran, R. 1977. Pottery from the Chalcolithic site near Tell Delhamiya and some notes on the character of the Chalcolithic-Early Bronze I Transition. EretzIsrael XIII:48-56 (Hebrew with English summary). Amiran, R. 1985. The Transition from the Chalcolithic to the Early Bronze Age. In J. Amitai (ed.), Biblical Archaeology Today: 108-147. Jerusalem: Israel Exploration Society. Artin, G. 2010. The necropolis and dwellings of Byblos during the Chalcolithic Period: new interpretations. Near Eastern Archaeology 73/2-3:74-85. Avner, U. 2002. Studies in the Material and Spiritual Culture of the Negev and Sinai Populations, during the 6th3rd millennia BC. Unpublished PhD thesis, Hebrew University of Jerusalem. Banning, E. B., Rahimi, D., Siggers, J. and Ta’ani, H. 1996. The 1992 season of excavations in Wadi Ziqlab, Jordan. Annual of the Department of Antiquities of Jordan XL:29-49. Banning, E. B., Blackham, M. and Lasby, D. 1998. Excavations at WZ 121, a Chalcolithic site at Tubna This same vessel has been mistakenly described as pottery by Garfinkel (1999: Fig. 134:1), who ascribed it to the LC period. 23 

90

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’ Braun, E. 1989a. The transition from the Chalcolithic to the Early Bronze Age in northern Israel and Transjordan: is there a missing link? In P. de Miroschedji (ed.), L’urbanisation de la Palestine à l’âge du Bronze ancien: bilan et perspectives des recherches actuelles, (Actes du Colloque d’Emmaüs: 20-24 octobre 1986): 7-28. British Archaeological Reports International Series 527{i}. Oxford: British Archaeological Reports. Braun, E. 1989b. The problem of the apsidal house: new aspects of Early Bronze I domestic architecture in Israel, Jordan and Lebanon. Palestine Exploration Quarterly 121:1-43. Braun, E. 1990. Basalt bowls of the EB I Horizon in the Southern Levant. Paléorient 16:87-95. Braun, E. 1996. Cultural Diversity and Change in the Early Bronze I of Israel and Jordan: Towards an Understanding of the Chronological Progression and Patterns of Regionalism in Early Bronze Society. Unpublished PhD thesis, Tel Aviv University. https://www.academia. edu/6071150/16_Braun_Ph_D_Refs_ Braun, E. 1997. Salvage and Rescue Excavations at the Late Prehistoric Site of Yiftah’el in Lower Galilee, Israel (IAA Reports 2). Jerusalem: Israel Antiquities Authority. Braun, E. 2000a. Area G at Afridar, Palmahim Quarry 3 and the earliest pottery of Early Bronze I: part of the Missing Link. In G. Philip and D. Baird (eds), Breaking with the Past: Ceramics and Change in the Early Bronze Age of the Southern Levant:113-128. Sheffield: Sheffield Academic Press. Braun, E. 2000b. Post Mortem: A late prehistoric site at Palmahim Quarry. Bulletin of the Anglo-Israel Archaeological Society 17:17-28. Braun, E. 2001. Proto and early dynastic Egypt and Early Bronze I-II of the Southern Levant: uneasy 14C correlations. Radiocarbon 43:1202-1218. Braun, E. 2004. Early Beth Shan (Strata XIX-XIII): G. M. FitzGerald’s Deep Cut on the Tell (University Museum Monograph 121). Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Braun, E. 2008. Deconstructing Y. Garfinkel’s ‘Beth Shean Ware’ and ‘Middle Chalcolithic’: a realistic look at two archaeological chimeras. Journal of the Serbian Archaeological Society 24:81-108. Braun, E. 2011. Chapter 12: The transition from Chalcolithic to EB I in the southern Levant: a ‘lost horizon’ slowly revealed. J. L. Lovell and Y. M. Rowan (eds.), Culture, Chronology and the Chalcolithic: Theory and Transition:160-177 (CBRL Levant Supplementary Monograph Series Vol. 9.) Oxford and Oakville: Oxbow. Braun, E. 2012a. On some south Levantine Early Bronze Age ceramic ‘wares’ and styles. Palestine Exploration Quarterly 144/1:5-32. Braun, E. 2012b. Qiryat Ata: New perspectives on a late prehistoric site in the Southern Levant. Studies in

Ancient Art and Civilization (Jagiellonian University, Krakow) 16:7-38. Braun, E., van den Brink, E. C. M., Regev, J., Boaretto, E. and Bar, S. 2013. Aspects of radiocarbon determinations and the dating of the transition from the Chalcolithic Period to Early Bronze Age I in the Southern Levant. Paléorient 39/1:23-46. Braun, E., Moth, W. E. and Politis, K. D. 2012. Chapter II. Early Bronze Age I Period. Sanctuary of Lot at Deir ‘Ain ‘Abata in Jordan: Excavations 1988-2003. Amman: Jordan Distribution Agency. Braun, E. and Roux, V. 2013. The Late Chalcolithic to early Bronze Age I Transition in the Southern Levant, determining continuity and discontinuity or ‘mind the gap’. Paléorient 39/1:15-22. Braun, E. and Gophna, R. 2004. Excavations at Afridar – Area G. ‘Atiqot 45:185-242. Burton, M. and Levy, T. E. 2001. The Chalcolithic radiocarbon record and its use in Southern Levantine Archaeology. Radiocarbon 43/3:12231246. Burton, M. and Levy, T. E. 2011. Chapter 13: The end of the Chalcolithic Period (4500-3600 BC) in the Northern Negev Desert, Israel. In J. L. Lovell and Y. M. Rowan (eds.), Culture, chronology and the Chalcolithic: Theory and Transition: 178-191 (CBRL Levant Supplementary monograph series Vol. 9). Oxford and Oakville: Oxbow. Clarke, J., Brooks, N., Banning, E.B., Bar-Matthews, M., Campbell, S., Clare, L., Cremaschi, M., di Lernia, S., Drake, N., Gallinaro, M., Manning, S. Nicoll, K., Philip, G., Rosen, S., Schoop, U-D., Tafuri, M. A. Weninger, B. and Zerboni, A. 2016. Climatic changes and social transformations in the Near East and North Africa during the ‘long’ 4th millennium BC: A comparative study of environmental and archaeological evidence. Quarternary Science Reviews 136:96-121. Commenge, C., Levy T. E., Alon, D. and Kansa, E. 2006. Chapter 15. Gilat’s figurines: exploring the social and symbolic dimensions of representation. In T. Y. Levy (ed.) Archaeology, Anthropology and Cult: The Sanctuary at Gilat, Israel: 739-830. London: Equinox. Commenge, C., Levy, T. E. and Kansa, E. 2006. Chapter 10. Gilat’s ceramics: cognitive dimensions of pottery production. In T. E. Levy (ed.), Archaeology, Anthropology and Cult: The Sanctuary at Gilat, Israel: 394-506. London: Equinox. Commenge-Pellerin, C. 1987. La poterie d’Abou Matar et de l’Ouadi Zoumeili (Beersheva) au ive millénaire avant l’ère chrétienne. Paris: Paléorient. Commenge-Pellerin, C. 1990. La poterie de Safadi (Beersheva) au ive millénaire avant l’ère chrétienne. Paris: Paléorient. de Contenson, H. 1960a. Three soundings in the Jordan Valley. Annual of the Department of Antiquities of Jordan IV-V:12-98. de Contenson, H. 1960b. La chronologie relative du niveau le plus ancien de Tell esh Shuna (Jordanie) 91

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead d’après le decouvertes recentes. Mélanges de l’Université Saint Joseph XXXVII:57-75. de Contenson, H. 1961. Remarques sur le chalcolithique récent de Tell esh-Shuneh. Revue Biblique LXVIII: 546-556. Dollfus, G. and Kafafi, Z. 1993. Recent Researches at Abu Hamid. Annual of the Department of Antiquities of Jordan XXXVII:241-262. Dothan, M. 1957. Excavations at Meser 1956: Preliminary Report on the First Season. Israel Exploration Journal 7:217-228. Dothan, M. 1959. Excavations at Meser 1957: Preliminary Report on the Second Season. Israel Exploration Journal 9:13-29. Dothan, M. 1971. The Late Chalcolithic Period in Palestine: Chronology and Foreign Contacts. EretzIsrael X:126-131, xii-xiii. (Hebrew, English summary). Jerusalem: Israel Exploration Society. Dunand, M. 1973. Fouilles de Byblos V: texte et planches. Paris: Maisonneuve. Dusar, B., Verstraeten, G., Notebaert, B. and Bakker, J. 2011. Holocene environmental change and its impact on sediment dynamics in the Eastern Mediterranean. Earth-Science Reviews 108:137-157. Eisenberg, E., Gopher, A. and Greenberg, G. 2001. Tel Te’o: A Neolithic, Chalcolithic, and Early Bronze Age Site in the Hula Valley (IAA Reports 13). Jerusalem: Israel Antiquities Authority. Elliott, C. 1978. The Ghassulian Culture in Palestine: origins, influences and abandonment. Levant 10:3754. Engberg, R. and Shipton, G. 1934. Notes on the Chalcolithic and Early Bronze Pottery of Megiddo (Studies in Ancient Oriental Civilizations 10). Chicago: Oriental Institute, University of Chicago. Epstein, C. 1984. A Pottery Neolithic Site Near Tel Qatif. Israel Exploration Journal 34:209-19. Eshed, V. and Bar, S. 2012. An innovative analysis of infant burials from the Chalcolithic Site of Fazael 2, Israel. Israel Exploration Journal 62/2:129-140. Fabian, P., Sheftelowitz, N. and Gilead, I. 2015. Horvat Qarqar South: report on a Chalcolithic cemetery near Qiryat Gat, Israel. Israel Exploration Journal 65/1:1-30. Faltings, D. A. 1998. Canaanites at Buto in the early fourth millennium BC. Egyptian Archaeology 13:2932. Fernandez-Tresguerres, J. A. 1992. Jebel Mutawwaq: los inicios de la edad del Bronce en la zona de Wadi Zarqa (Jordania). Treballs d’Arqueologia 2:128-143. Fernandez-Tresguerres, J. A. 2005. Jebel al-Mutawwaq (Jordania). Actas de las jornadas de archaeología en Asturias (Abril-Maio, 2005). https://www.academia. edu/229536/Actas_de_las_I_Jornadas_de_ Arqueologia_en_Asturias (last accessed 5/23/2016). Finné, M., Holmgren, K., Sundqvist, H. S., Weiberg, E. and Lindblom, M. 2011. Climate in the Eastern Mediterranean, and adjacent regions, during the

past 6000 years: a review. Journal of Archaeological Science 38:3153-3173. FitzGerald, G. M. 1934. Excavations at Beth-Shan in 1933. Palestine Exploration Quarterly for 1934:123134. FitzGerald, G. M. 1935. Beth Shan: Earliest Pottery. The Museum Journal (The University Museum, University of Pennsylvania) XXIV:5-22. Garfinkel, Y. 1992. The Pottery Assemblages of the Sha’ar Hagolan and Rabah Stages of Munhata [Israel] (Les cahiers du Centre de Recherche Français de Jerusalem 6). Paris: Paléorient. Garfinkel, Y. 1999. Neolithic and Chalcolithic Pottery of the Southern Levant (Qedem 39). Jerusalem: The Hebrew University of Jerusalem. Getzov, N., Lieberman-Wander, R., Smithline, H. and Syon, D. 2009. Horbat ‘Uza: the 1991 excavations. Vol. I: The Early Periods (IAA Reports 41). Jerusalem: The Israel Antiquities Authority. Gilead, I. 1994. The history of the Chalcolithic settlement in the Nahal Beer Sheva Area: the radiocarbon aspect. Bulletin of the American Schools of Oriental Research 296:1-14. Gilead, I. 2011. Chalcolithic Culture History: Ghassulian and other Chalcolithic entities in the Southern Levant. In J. L. Lovell and Y. M. Rowan (eds), Culture, Chronology and the Chalcolithic: Theory and Transition :12-24 (CBRL Levant Supplementary Monograph Series Vol. 9). Oxford and Oakville: Oxbow. Gilead, I., Bar-Yosef, D. E., Ben-Tor, A., Fabian, P., Goren, Y., Grigson, C., Hershman, D. and Marder, O. 1995. Grar: A Calcolithic Site in the Northern Negev (BeerSheva VII). Gilead, I. and Fabian, P. 2010. Pre-Ghassulian sites in the Ramot Neighborhood, Beer Sheva: chronological and cultural perspectives. in S. Yona (ed.), Or LeMayer: Studies in Bible, Semitic Languages, Rabinic Literature and Ancient Civilizations Presented to Mayer Gruber on the Occasion of His Sixty-Fifth Birthday: *89-*107. Beersheva: Ben Gurion University of the Negev Press. Golani, A. 2004. Salvage excavations at the Early Bronze Age site of Afridar, Ashqelon--Area E. ‘Atiqot 45:962. Golani, A. 2008. The Early Bronze Age site of Ashqelon, Afridar—Area M. ‘Atiqot 60:19-52. Golani, A. 2013. The transition from the Late Chalcolithic to the Early Bronze I in Southwestern Canaan-Ashqelon as a case for continuity. Paléorient 39/1:95110. Golani, A. 2014. Ashqelon during the EB I Period – a centre for copper processing and trade. In A. Mączyńska (ed.), The Nile Delta as a Centre of Cultural Interactions between Upper Egypt and the Southern Levant in the 4th millennium BC. Proceedings of the conference held in the Poznan Archaeological Museum, Poznań, Poland, 21-22 June 2013 :119-138 (Studies in African Archaeology Vol. 13). Poznań: Archaeological Museum. 92

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’ Golani, A. and Nagar, Y. 2011. Newly discovered burials of the Chalcolithic and Early Bronze Age I in Southern Canaan: evidence of cultural continuity? In J. L. Lovell and Y.M. Rowan (eds), Culture, Chronology and the Chalcolithic: Theory and Transition: 84-96 (CBRL Levant Supplementary monograph series Vol. 9). Oxford and Oakville: Oxbow. Golani, A. and Rosenberg, D. 2012. Groundstone tools of a coppersmith’s community: understanding stone-related aspects of the Early Bronze Age site of Ashqelon, Barnea. Journal of Mediterranean Archaeology 25/1:27-51. Gopher, A. and Greenberg, R. 1996. Part II: the pottery of the Neolithic Levels. In Biran, A., Ilan, D. and Greenberg, R., Dan I: A Chronicle of the Excavations, the Pottery Neolithic and the Early Bronze Age and the Middle Bronze Age Tombs: 65-80 (Annual of the Nelson Glueck School of Biblical Archaeology, Hebrew Union College-Jewish Institute of Religion). Jerusalem: Hebrew Union College. Gopher, A. and Orelle, E. 1995. The Groundstone Assemblages of Munhatta: a Neolithic Site in the Jordan Valley, Israel (Les Cahiers des Missions Archeologiques Français in Israel 7). Paris: Association Paléorient. Gophna, R. 1997. The southern coastal troughs as EB I subsistence areas. Israel Exploration Journal 47:155161. Gophna, R. 2004. Excavations at Ashqelon, Afridar: introduction. ‘Atiqot 45:1-8. Gophna, R. and Lifshitz, S. 1980. A Chalcolithic burial cave at Palmahim. `Atiqot (English Series) XIV:18. Gorzalczany, A. 2007. Centro y periferia en el antiguo Israel: nuevas aproximaciones a las practicas funerarias del Calcolitico en la Planicie Costera. Antiguo Oriente 5:205-230. Greenberg, R., Eisenberg, E., Paz, S. and Paz, Y. 2006. Bet Yerah: The Early Bronze Age Mound. Volume I. Excavation Reports, 1933-1986 (IAA Reports 30). Jerusalem: Israel Antiquities Authority. Greenberg, R. and Paz, S. 2004. An EB IA-EB III stratigraphic sequence from the 1946 Excavations at Tel Beth Yerah. Israel Exploration Journal 54/1:1-23. Gustavson-Gaube, C. 1985. Tell esh-Shuna North 1984: A preliminary report. Annual of the Department of Antiquities of Jordan XXIX:43-87. Gustavson-Gaube, C. 1986. Tell esh-Shuna North 1985: A preliminary report. Annual of the Department of Antiquities of Jordan XXX:69-113. Haiman, M. 1992. Cairn burials and cairn fields in the Negev. Bulletin of American Schools of Oriental Research 287:25-45. Hanbury-Tenison, J. W. 1986. The Late Chalcolithic to Early Bronze I Transition in Palestine and Transjordan (British Archaeological Reports International Series 311). Oxford: British Archaeological Reports. Hennessy, J. B. 1969. Preliminary report on a first season of excavations at Teleilat Ghassul. Levant I:1-24.

Hennessy, J. B. 1982. Teleilat Ghassul: its place in the archaeology of Jordan. In A. Hadidi ed. Studies in the History and Archaeology of Jordan I :55-58. Jordan: Department of Antiquities. Hull, A. J. T., Donahue, D. J., Carmi, I. and Segal, D. 1998. Chapter 20: radiocarbon dating of finds. In T. Schick. The Cave of the Warrior: A Fourth Millennium Burial in the Judean Desert: 110-111 (IAA Reports 5). Jerusalem, Israel Antiquities Authority. Joffe, A. H. and Dessel, J. P. 1995. Redefining chronology and terminology for the Chalcolithic of the Southern Levant. Current Anthropology 36:507-518. Kahila Bar-Gal, G. and Smith, P. 2001. Chapter 12: the human remains. In Eisenberg, E., Gopher, A. Greenberg, R. Tel Te’o, A Neolithic, Chalcolithic, and Early Bronze Age Site in the Hula Valley (IAA Reports 13), 163-170. Jerusalem: Israel Antiquities Authority. Kaptijn, E. and de Vreeze, M. 2008. Different communities or changing pottery techniques? The transition from the Late Chalcolithic to the Early Bronze Age discovered at Qatar Damiyeh Jordan Valley. Leiden Journal of Pottery Studies 24:75-94. Kerner, S. 2009. The pottery of Tall Hujayrat al-Ghuzlan 2000-2004. In L. Khalil and K. Schmidt (eds), Prehistoric cAqaba I: 127-232 (Orient-Archäologie Band 23). Rahden, Westfalia: Verlag Marie Leidorf GmbH. Khalaily, H. 2004. An Early Bronze Age site at Ashqelon, Afridar—Area F. ‘Atiqot 45:121-160. Khalaily, H. 2008. The Flint Assemblage. In A. Golani, The Early Bronze Age Site of Afridar, Ashqelon— Area E. ‘Atiqot 45:36-40. Khalaily, H. and Kamaisky, E. 2002. The use of sickle blades for decorating pottery in the Wady Rabah Culture: The case of Tel Dover. In E. C. M. van den Brink and E. Yannai (eds), In Quest of ancient Settlements and Landscapes: Archaeological Studies in Honor of Ram Gophna: 57-62 Tel Aviv: Ramot. Khalil, L. and Schmidt, K. (eds) 2009. Prehistoric ‘Aqaba I (Orient-Archäologie Band 23). Deutsches Archäologisches Institut Orient-Abteilung: Westfalia: Verlag Marie Leidorf GmbH. Klimscha, F. 2009. Radiocarbon dates from prehistoric c Aqaba and other related sites from the Chalcolithic Period. In L. Khalil and K. Schmidt (eds), Prehistoric c Aqaba I: 363-419 (Orient-Archäologie Band 23). Rahden, Westfalia: Verlag Marie Leidorf GmbH. Klimscha, F. 2013. Innovations in Chalcolithic metallurgy in the Southern Levant during the 5th and 4th Millennium BC: copper-production at Tall Hujayrāt al-Ghuzlān and Tall al-Magaṣṣ, Άqaba Area, Jordan. In S. Burmeister, S. Hansen, M. Kunst and N. MullerScheesel (eds) Metal Matters: Innovative Technologies and Social Change in Prehistory and Antiquity: 31-63. Rahden, Westfalia: Verlag Marie Leidorf GmbH. Koeppel, R., Senès, H., Murphy, J. W., Mahan, G. S. 1940. Teleilat Ghassul II: Compte Rendu des Fouilles de l’Institut Biblique Pontifical 1932-1936. Rome: Institut Biblique Pontifical. 93

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Levy, T. E. (ed.) 2006. Archaeology, Anthropology and Cult: the Sanctuary at Gilat, Israel. London: Equinox. Lovell, J. L. 2001. The Late Neolithic and Chalcolithic Periods in the Southern Levant. New Data from the Site of Teleilat Ghassul, Jordan (Monographs of the Sydney University Teleilat Ghassul Project 1/BAR International Series 974). Oxford: British Archaeological Reports. Lovell, J. L. 2010. Community is cult, cult is community: weaving the web of meanings for the Chalcolithic. Paléorient 36/1:103-122. Lovell, J. L., Kafafi, Z. and Dollfus, G. 1997. A Preliminary note on the ceramics from the basal level of Abu Hamid. In H. G. Gebel, Z. Kafafi and G. O. Rollefson (eds). Prehistory of Jordan II. Perspectives from 1997:361370 (Studies in Early Near Eastern Production, Subsistence and Environment 4). Berlin: ex-Oriente. Lutfi, Kh. and Schmidt, K (eds). 2009. Prehistoric ‘Aqaba I (Orient-Archäologie Band 23). Berlin: Duetsches Archäologisches Institut. Macdonald, E. 1932. Prehistoric Fara. In Beth-Pelet II:121. London: British School of Archaeology in Egypt. Mallon, A. 1932. La civilisation du IIIe millénaire dans la Vallée du Jourdain: les fouilles de Teleilat Ghassul. Syria XIII:333-344. Mallon, A., Koeppel, R. and Neuville, R. 1934 Teleilat Ghassul I: Compte rendu des fouilles de l’Institut Biblique Pontifical 1929-1932. Rome: Institut Biblique Pontifical. Mellaart, J. 1966. The Chalcolithic and Early Bronze Ages in the Near East and Anatolia. Beirut: Khayata. Milevski, I. and Krokhmalnik, A. 2010. Ashqelon, Barn’ea B-C. Hadashot Arkheologiyot http://www.hadashotesi.org.il/report_detail_eng.asp?id=1395&mag_ id=117 (Last accessed 23/5/2016. Nagar, Y. and Eshed, V. 2001. Where are the children? Age-dependent burial practices in Peqi’in. Israel Exploration Journal 51/1:27-35. Neuville, R. 1930. Notes de préhistoire Palestinienne. Journal of Palestine Oriental Society X:193-221. North, R. 1961. Ghassul 1960: Excavation Report (Analecta Biblica.14). Roma: Pontificio Istituto Biblico. Orni, E. and Efrat, E. 1980. Geography of Israel: Fourth Revised Edition. Jerusalem: Israel Universities Press. PAJM (Projecto Archaeológico Jebel Mutawwaq: Facebook) 2014 https://www.facebook.com/ pages/Proyecto-Arqueol%C3%B3gico-JebelMutawwaq/278792018925526 (last accessed 23/5/2016). Paz, Y. and Nativ, A. 2013. Yesodot, Israel: A case for a post-Ghassulian entity. Paléorient 32/1:83-93. Perrot, J. 1972. Préhistoire Palestinienne—prémices. Vol. 8: Columns 286-446 in L. Pirot, A. Robert, H. Cazelles and A. Feuillet, (eds), Dictionnaire de la bible (Supplément). Paris: Letouzey et Ané. Perrot and Ladiray, D. 1980. Tombes à ossuaires de la région côtière palestinienne au ive millénaire avant l’ère chrétienne (Mémoires et travaux du Centre de Recherches Préhistoriques Français de Jérusalem 1). Paris: Paléorient.

Philip, G. and Baird, D. 1993. Preliminary report on the second (1992) season of excavations at Tell eshShuna North. Levant XXV:13-35. Prag, K. 1989. Preliminary report on the excavations at Tell Iktanu, Jordan. Levant XXI:33-45. Prag, K. 1995. The Dead Sea dolmens: death and the landscape. In S. Campbell and A. Green eds. The Archaeology of Death in the Ancient Near East: 75-83 (Oxbow Monograph 51). Oxford: Oxbow. Rast, W. E. and Schaub, R. T. 1974. Survey of the Southeastern Plain of the Dead Sea, 1973. Annual of the Department of Antiquities of Jordan XIX:5-54. Ravikovitch, S. 1981. The Soils of Israel: Formation, Nature and Properties (Hebrew with English summary). Tel Aviv: Hakibbutz Hameuchad Publishing House. Regev, J., de Miroschedji, P., Greenberg, R., Braun, E, Greenhut, Z. and Boaretto, E. 2012. Chronology of the Early Bronze Age in the Southern Levant: new analysis for a high chronology. Radiocarbon 54/34:525-566. Roux, V., van den Brink, E. C. M. and Shalev, S. 2013. Continuity and discontinuity in the Shephela (Israel) between the Late Chalcolithic and the Early Bronze I: the Modi’in ‘Deep Deposits’ ceramic assemblages as a case study. Paléorient 39/1:63-81. Rowan, Y. M. 2004. The ground stone assemblage from Ashqelon, Afridar--Area E. ‘Atiqot 45:85-96. Rowan, Y. M. and Golden, J. 2009. The Chalcolithic Period of the Southern Levant: a synthetic review. Journal of World Prehistory 22:1-92. Saidah, R. 1979. Fouilles de Sidon--Dakerman: l’agglomération chalcolithique. Berytus XXVII: 2955. Schick, T. 1998. The Cave of the Warrior: A Fourth Millennium Burial in the Judean Desert (Israel Antiquities Authority Reports 5). Jerusalem, Israel Antiquities Authority. Smith, P., Zagerson, T., Sabari, P., Golden, J., Levy, T. E., and Dawson, L. 2006. Chapter 8: death and the sanctuary: the human remains from Gilat. In T. E. Levy (ed.), Archaeology, Anthropology and Cult: the Sanctuary at Gilat, Israel: 327-368. London: Equinox. Schaub, R. T. and Rast, W. E. 989. Bâb edh-Dhrâc: Excavations in the Cemetery Directed by Paul W. Lapp (1965-67). Winona Lake, Indiana: Eisenbrauns. Scheftelowitz, N. 2002. Chapter 4: stratigraphy, architecture and tombs. Area B. In A. Kempinski, N. Scheftelowitz and R. Oren (eds), Tel Kabri: The 1986-1993 Excavation Seasons. Tel Aviv: 19-34 (University Monographs No. 20). Tel Aviv: Institute of Archaeology, Tel Aviv University. Scheftelowitz, N. and Oren, R. 2004. Giv’at Ha-Oranim. A Chalcolithic Site (Salvage Excavation Reports 1). Tel Aviv: Institute of Archaeology Tel Aviv University. Sebbane, M. 2009. The Mace in Israel and the Ancient Near East from the Ninth Millennium to the First: Typology and Chronology, Technology, Military and Ceremonial Use, Regional Interconnections. Unpublished PhD thesis. Tel Aviv University (Hebrew). 94

E. Braun: Forging a Link: Evidence for a ‘Lost Horizon’ Sebbane, M. 2014. The Hoard from Nahal Mishmar, and the Metal-working Industry in Israel in the Chalcolithic Period. In M. Sebbane, O. Misch-Brandl and D. M. Master (eds). Masters of Fire: Copper Age Art from Israel: 115-136, 168-182. Princeton and Oxford: Princeton University. Segal, I., Halicz, L. and Kamenski, A. 2004. The metallurgical remains from Ashqelon, Afridar-Areas E, G and H. ‘Atiqot 45:311-330. Smith, P., Zagerson, T. Sabari, P., Golden, J., Levy, T. E. and Dawson, L. 2006. Chapter 8. Death and the Sanctuary: the Human Remains from Gilat. In T. E. Levy (ed.), Archaeology, Anthropology and Cult: The Sanctuary at Gilat, Israel: 327-365. London: Equinox. Stekelis, M. 1935. Les monuments mégalithiques de Palestine (Archives de l’Institut de Paléontologie Humaine: mémoire 15). Paris: Masson. Stekelis, M. 1961. La necropolis megalitic.de Ala-Safat, Transjordania (Monografias, I, del Instituto de Prehistoria y Arquelogia). Barcelona: Diputacion Provincial de Barcelona. Sukenik, E. L. 1936. Late Chalcolithic pottery from ‘Affuleh. Palestine Exploration Fund Quarterly Statement: 150-154. Sukenik, E. L. 1948. Archaeological investigations at `Affula. Journal of the Palestine Oriental Society XXI:1-7. van den Brink, E. C. M. 2013. A late Chalcolithic to Early Bronze Age I progression at the Buchman South Quarter in Modi’in in the central piedmont (Shephela) of Israel. Paléorient 39/1:47-61. Vaillant, N. 1988. Les techniques de la ceramique. In G. Dollfus, et al. (eds), Abu Hamid: village du 4e millénaire de la vallée du Jourdain: 44-46. Amman: Centre Culturel Français. de Vaux, R. 1970. Palestine during the Neolithic and Chalcolithic Periods. In I. E. S. Edwards, C. J. Gadd and N. G. L. Hammond (eds), Cambridge Ancient History, Third edition, Vol. 1, Part 1:498-535. Cambridge: Cambridge U. Press.

de Vaux, R. 1971. Palestine in the Early Bronze Age. Pp. In I. E. S. Edwards, C. J. Gadd and N. G. L. Hammond (eds), Cambridge Ancient History, Third edition, Vol. 1, Part 1:208-237. Cambridge: Cambridge University Press. de Vaux, R. M. and Steve, A. M. 1949. La deuxième campagne de fouilles de Tell el-Far’ah, près Naplouse. Revue Biblique 56:102-138. Verheydena, S., Naderc, F. H., Chengd, H. J., Edwardsd, L. R., Swennene, R. 2008. Paleoclimate reconstruction in the Levant region from the geochemistry of a Holocene stalagmite from the Jeita cave, Lebanon. Quaternary Research 70/3:368-381. Waheeb, M. 1995. The first season of the An-Naq Project, Ghawr as-Safi. Annual of the Department of Antiquties of Jordan XXXIX:553-555. Wright, G. E. 1937. The Pottery of Palestine from the Earliest Times to the End of the Early Bronze Age. New Haven: American Schools of Oriental Research. Wright, G. E. 1958. The Problem of the Transition between the Chalcolithic and Bronze Ages. EretzIsrael V:37*-45*. Yannai, E. 2006.`En Esur (‘Ein Asawir) I. Excavations at a Protohistoric Site in the Coastal Plain of Israel (Israel Antiquities Authority Annual Reports 31). Jerusalem: The Israel Antiquities Authority. Yannai, E., Gophna, R., Liphshitz, S. and Liphshitz, Y. 2013. A Late Bronze Age Cemetery on the Coast of Palmahim. ‘Atiqot 74:9-57. Yekutieli, Y. and Gophna, R. 1994. Excavations at an Early Bronze Agesite near Nizzanim. Tel Aviv 21:162185. Zbenovich, V. G. 2004a. The Flint Assemblage from Ashqelon, Afridar—Area E. ‘Atiqot 45:63-84. Zbenovich, V. G. 2004b. The Flint Assemblage from Ashqelon, Afridar—Areas G and J. ‘Atiqot 45:263278.

95

Unresolved Pottery Neolithic Chrono-Stratigraphic and Chrono-Cultural issues: Comments on the Beginning and the End of the Pottery Neolithic Period Avi Gopher Introduction

following its stratigraphic position. The question of the Lodian is further explored below.

The Pottery Neolithic (PN) period as a whole emerges from the archaeological record as a clear chronostratigraphic unit, including a series of cultural entities of which the major ones are well defined in time and space and characterized by a recurring set of material culture elements. Nevertheless, some chrono-cultural issues and disagreements still remain.

The scarcity of dates for Wadi Rabah sites (see Gopher 2012, Chapter 41), results in a rather loosely defined 14C range for this entity lasting for some 700-900 years. The earliest site quoted for the Wadi Rabah culture is that of Ard Tlaili, dated to C. 7700 cal BP. However, the more mainstream evidence seems to show that the Wadi Rabah culture made its debut sometime around 76007500 cal BP.

C ages are by far the most convenient tool with which to establish detailed chronological frameworks. While they are straightforward and independent, they are nevertheless not entirely free of flaws (see, for example Carmi and Boaretto 2012; Gopher 2012, Chapters 40 and 41). Moreover, although dates are continuously accumulating, the absolute dating of the PN and its cultural entities suffers an almost chronic lack of 14C dates per cultural entity or site. A list of 14C ages from PN sites compiled from published lists was presented by Gopher (2012, Chapter 41 and references therein), and special attention was given to provide the information as originally published, free of interpretation. 14

Where the Wadi Rabah culture ends is another difficult question on which several takes are found among researchers. As these views concern the end of the PN at large, they are further elaborated on below. It is quite accepted in present day research that the Wadi Rabah culture is not the final PN culture. Considering Tel Ẓaf and Tel Ali, for example, as representatives of postWadi Rabah cultural entities, the Wadi Rabah culture would have ended, in my view, around 6800 cal BP. Although we may assume it started earlier, the Qatifian (according to dates from Qatif Y-3 and Ain Waida in the northern Negev and the Dead Sea area), spanned ca. 7000-6800 cal BP and thus seems to have overlapped the latest centuries of the Wadi Rabah range.

Delimited by the PPNC at its ‘lower’ (older) end and the Chalcolithic Ghassulian at its ‘upper’ (younger) end, the two millennia of the PN, a major component in the later prehistory of the region, spans the period between C. 8500 and 6500 cal BP (for details and discussion see Gopher 2012, Chapter 41).

The post Wadi Rabah-pre Ghassulian (PoWR-PG) entities of central and northern Israel are not very well known except for a small number of cases. These entities are usually known from very few sites or sometimes only a single site, and the 14C inventory of dates is accordingly limited, as well. Conspicuous are certain date series from sites in the Jordan Valley, such as Tel Ali 1b and Tel Ẓaf, as well as the problematic date series of olive stones from Nahal Zehora II. The range of the PoWRPG dates spans some 400 years from 6750-6310 cal BP (for details see Gopher 2012, Chapter 41). Using only dates of clearly assigned PoWR-PG sites such as the above-mentioned Tel Ẓaf and Tel Ali 1b, the range for the PoWR-PG would shrink to some 300-200 years only. In the south, the post-Qatifian (but pre Ghassulian) Besorian is dated to 6600-6500 cal BP (Gilead 2007, 2009, 2011). This is still within the PoWR-PG range. It may thus be said that albeit significant variability, when put together, the PoWR-PG entities show a short range spanning some 300 years between C. 6800-6500 cal BP.

Following the available evidence concerning absolute dating, the PN may be portrayed as follows: The PN followed the Pre-Pottery Neolithic C (PPNC), the latest part of the Pre-Pottery Neolithic (PPN) period. The PPNC ended between 8500-8400 cal BP. The Yarmukian seems to have lasted from 400 to 600 years – starting 8500/400 and ending C. 8000-7800 years cal BP. The Lodian, immediately following the Yarmukian, may have lasted some 200-300 years. Dating the Lodian is quite challenging as available dates are scarce and problematic (see Gopher 2012, Chapter 41). I would expect the Lodian to have begun sometime after the Yarmukian, around 7900 or 7800 cal BP, and end within a couple of centuries around 7700-7600 cal BP. While there is no satisfying foundation for this estimate at the present stage of research, we may rely on Yarmukian and Wadi Rabah 14C dates which delimit the Lodian 96

A. Gopher: Unresolved Pottery Neolithic Chrono-Stratigraphic and Chrono-Cultural issues There is a general consensus concerning the beginning of the Chalcolithic Ghassulian, and the issue has been discussed in numerous papers since the late 1980’s (e.g., Burton and Levy 2001; Gilead 2004, 2006, 2009; Joffe and Dessel 1995; Levy 1995; Levy and Burton 2006) as well as by Rowan and Golden (2009). While 14C records in the north are few, the central and southern parts of the southern Levant provide a large dataset of dates, all of which indicate that the Chalcolithic Ghassulian emerged around 6500 cal BP. This means that the PN ended around 6500 years cal BP.

always predate the Chalcolithic Ghassulian wherever these two appear together (Fig. 1). Examples of sites in which PN is found over the PPN include Yarmukian over PPN cases such as Shaar haGolan (Barzilai and Garfinkel 2006; Garfinkel 1993; Garfinkel and Ben-Shlomo 2009; Garfinkel and Miller 2002a), Ain Ghazal (Rollefson et al. 1992), Munḥata (Gopher 1989b; Perrot 1968), ha-Bashan St. (Kaplan 1970, 1972a, 1972b), Mishmar ha-Emek (Barzilai and Getzov 2008), Naḥal Beẓet II and Ard el-Samra (Getzov et al. 2009a); Lodian over PPN cases such as Ha-Gosherim (Getzov 2008a; Khalaily 1999), Tel Teo (Eisenberg et al. 2001), Jericho (Kenyon 1981; Kenyon and Holland 1983), Abu Ghosh (Khalaily and Marder 2003), Yiftaḥel (Braun 1997; Khalaily et al. 2008; Milevski et al. 2008), Wadi Shueib (Simmons et al. 2001); and Wadi Rabah over PPN such as Tel Ali (Garfinkel 1992a; Prausnitz 1970) and Naḥal Beẓet I (Gopher 1989c, 1993).

Relative chronology: stratigraphy and a comment on typology Beyond absolute chronology, the elementary building stones of the archaeological story of the PN and its cultural entities, like any archaeological story, are the stratigraphy of the sites and the typology of their finds. Only when these are clarified is it possible to invest in the study of change over time and variation in space and attempt meaningful reconstructions of the dynamic developments in human societies. Stratigraphic positioning of the PN

Examples of PN strata predating the Chalcolithic Ghassulian include cases such as Wadi Rabah (Kaplan 1958b), Abu Ḥamid (Dollfus and Kafafi 1993; Lovell et al. 2007), En Esur (Yannai 2006), Munḥata (Gopher 1989b), and Ḥorbat Uza (Getzov 2009a).

As stated in the introduction, the position of the PN as a whole is quite clear. It always postdates the PPN wherever the two appear together, and PN strata

The stratigraphic position of the different cultural entities within the PN emerges quite consistently from multilayered PN sites in the southern Levant although

Figure 1. Schematic timeline and geography of PN cultures.

97

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead ‘full’ PN stratigraphies are rare. Naḥal Zehora II is one of the few examples where the PN as presented in this paper shows most of its cultural entities in stratigraphic sequence, starting with the Yarmukian, followed by the Lodian and Wadi Rabah and concluding with the minor presence of an undefined PoWR-PG entity. This stratigraphic order is further supported by an analysis of the finds, which reveals regularities in the change of material culture elements over time (Gopher 2012, Chapters 40, 41). In other instances showing more than one PN cultural entity in stratigraphy, things may become unclear or display irregularities. However, no case displays a reversed stratigraphy for any of the PN entities (for detailed discussions on this issue refer to Gopher 1995; Gopher and Gophna 1993; Kafafi 1987, 1998; for contesting opinions see Garfinkel 1992a; Obeidat 1995). Fig. 1 summarizes the chrono-stratigraphy of PN entities as presented here, depicting the major cultural entities plotted over time and schematically over space throughout the region.

Wadi Rabah, always appear above Yarmukian or Lodian layers where present and under either PoWR-PG or Chalcolithic Ghassulian layers. Examples of sites in which Wadi Rabah is found over the Yarmukian or Lodian include Jericho (Kenyon 1981; Kenyon and Holland 1983), Tel Teo (Eisenberg et al. 2001), Teluliyot Batashi (Kaplan 1958a), Tell Abu Zureiq (Anati et al. 1973), Naḥal Zehora II (see Gopher 2012, Chapter 4), En Esur (Yannai 2006), Ḥorbat Uza (Getzov 2009a) and Ha-Gosherim (Getzov 2008a, laterally). Examples of sites in which Wadi Rabah is found below PoWR-PG or the Chalcolithic Ghassulian include sites such as Wadi Rabah (Kaplan 1958b), Teluliyot Batashi (Kaplan 1958a), Munḥata (Gopher 1989b), Tel Teo (Eisenberg et al. 2001), Teleilat Ghassul (Bourke 2007 and references therein), Ḥorbat Uza (Getzov 2009a), Ha-Gosherim (Getzov 2008a), En Esur (Yannai 2006) and Abu Ḥamid (Dollfus and Kafafi 1993; Lovell et al. 2007 and references therein). The stratigraphic positioning of PoWR-PG entities is always above Wadi Rabah strata and below the Chalcolithic Ghassulian where present. Examples of sites in which the PoWR-PG is found over Wadi Rabah strata include Tel Ali (Garfinkel 1992b, 1999b), Naḥal Zehora II (see Gopher 2012 Chapter 4), Ein el Jarba (Meyerhof 1982), Ḥorbat Uza (Getzov 2009a) and Teluliyot Batashi (Kaplan 1958a). Examples of sites in which the PoWR-PG is found below the Chalcolithic Ghassulian include Ein el Jarba (Meyerhof 1982), En Esur (Yannai 2006), Ḥorbat Uza (Getzov 2009a), Teleilat Ghassul (e.g., Bourke 2007 and references therein), and Abu Ḥamid (Lovell et al. 2007) unless we consider the pre-Chalcolithic Ghassulian strata of phase II (or even only part of them) as belonging to the Wadi Rabah culture. The Nazur culture defined by Yannai (Yannai 2006; and see also Golan 2006) is known from single uni-component sites in central-western Israel. Consequently, it provides no stratigraphic clues. In the southern parts of Israel, the Qatifian and its successor, the Besorian, always precede the Chalcolithic Ghassulian (Gilead 1990, 2007, 2009; Gilead and Alon 1988; Goren 1990, 1991).

The Yarmukian is the oldest of the PN cultures. Stratigraphically, it appears below any other PN entity if present. There is no case in which the Yarmukian appears stratigraphically subsequent to any other PN cultural entity. Examples of sites in which the Yarmukian is found under other PN entities include under Lodian or Wadi Rabah in sites such as Munḥata (Garfinkel 1992a, 1992b; Gopher 1989b; Gopher and Orrelle 1995a; Perrot 1968); under Lodian at Naḥal Zehora II (Gopher 2012), Teluliyot Batashi (Kaplan 1958a); under Wadi Rabah in sites such as Naḥal Beẓet II, Ard el-Samra (Getzov et al. 2009a), Ḥorbat Uza (Getzov et al. 2009a, 2009b). Wherever the Lodian could be identified in stratigraphy, it was always found either directly over a PPN layer (see above), or over a Yarmukian layer, and always beneath Wadi Rabah layers. Examples of sites in which the Lodian is found over the Yarmukian are Munḥata if we consider the possibility that Perrot’s ‘Munḥata Phase’ represents the Lodian (Perrot 1968), Naḥal Zehora II (see Gopher 2012, Chapter 41). Examples of sites in which the Lodian is found below Wadi Rabah strata include Jericho (Kenyon 1981; Kenyon and Holland 1983), Teluliyot Batashi (Kaplan 1958a), Ha-Gosherim (Getzov 2008a, laterally), Tel Teo (Eisenberg et al. 2001), Naḥal Zehora II (see Gopher 2012), Lod (Gopher and Blockman 2004, laterally), En Esur (Yannai 2006), Yiftaḥel (laterally; Getzov 2009b; Khalaily et al. 2008), possibly also at Tell Abu Zureiq (Anati et al. 1973) if one accepts an on-site Lodian presence (see Gopher and Gophna 1993) and perhaps also at the site of Ghrubba (Kafafi 1987, 1998; Obeidat 1995; and see also notes in Kafafi 2011).

Examples of sites in which the Qatifian or the Besorian are found over earlier PN strata include Teluliyot Batashi (Kaplan 1958a), Ain Waida (laterally; Kuijt and Chesson 2002), and Teleilat Ghassul (Bourke 2007). These two southern entities may provide us with a clue as to the inter-cultural relations of southern and northern pottery-bearing entities predating the Chalcolithic Ghassulian. A relative chronology by typology based on seriation(s) should involve a multitude of sites throughout the region. The sites of Naḥal Zehora may serve as a minicosmos with respect to seriations if their assemblages are considered good and representative of the PN as

Notwithstanding the current debate regarding an accurate definition of the Wadi Rabah culture, strata of the Wadi Rabah culture sensu lato, including late 98

A. Gopher: Unresolved Pottery Neolithic Chrono-Stratigraphic and Chrono-Cultural issues a whole. The finds of the Naḥal Zehora sites provide numerous seriations of material culture elements in pottery (Gopher and Eyal 2012, Chapter 17), flint (Gopher and Barkai 2012, Chapter 26), groundstone tools (Gopher 2012, Chapter 24), and imagery items (Gopher and Eyal 2012, Chapter 29). The repetitive trends of change found within these assemblages allow for a clear reconstruction of relative chronology of the PN (Gopher 2012, Chapter 40). With hardly any exception, the finds reveal differences between the three major PN cultural entities – the Yarmukian, Lodian and Wadi Rabah, which also accord in their order with the abovementioned stratigraphy. The PoWR-PG entities do not emerge as clearly from the Naḥal Zehora finds due to their marginal on-site presence. These trends can be detected in other PN sites. In fact, they are verified by our comparative studies both in sites comprising partial PN sequences and in sites reflecting a single PN cultural entity.

explicitly, that the Lodian was earlier in the sequence. Kenyon refrained from taking a clear stand on this issue (Kenyon 1957, 1960, 1970, 1981; Kenyon and Holland 1983) although she mentioned the possibility that this entity was as old as the PPN. With very few exceptions, the debate died out for a while. The suggestion that the Lodian predates the Yarmukian lost its appeal. In the 1990’s Garfinkel (1992a, 1992b) suggested viewing the two as geographically differentiated contemporaneous entities. This view is advocated in several publications (e.g., Bar-Yosef and Garfinkel 2008; Garfinkel 1992a, 2008; Kafafi 2011; Obeidat 1995). The following arguments have been advanced to support Yarmukian and Lodian contemporaneity: –– There is no clear 14C record negating such a suggestion. –– There is no clear stratigraphy between the two entities, let alone a sequential one. –– Pottery typology shows that ‘the basic vessels’ (Garfinkel 1999b:101) of Yarmukian and Lodian are similar, the differences being merely in decoration, that is, ‘incised frame and herringbone in the Yarmukian [...] painted lines and burnish in the Jericho IX [Lodian – my addition]’ (Garfinkel 1999b:101). This makes one wonder why, then, did he separate them? –– Flint assemblages of the two entities are ‘very similar’ (Garfinkel 1999b:101) and CrowfootPayne is quoted as including the Yarmukian and the Jericho PNA (Lodian) assemblages together, assigning all to the Yarmukian (CrowfootPayne 1983). This begs the question, is there a Yarmukian settlement at Jericho? –– Geography is distinctive. While the Lodian (Jericho IX) is spread in the southern parts of Israel, the Yarmukian is located at the center and north of Israel. In other words, the ‘Sites do not overlap geographically except for some small find spots’ (Garfinkel 1999b:101).

Plotting generalized (average) seriation curves, a gradual and accumulative cultural change is revealed throughout the PN. In other words, each of the cultural assemblages (read entities) shows some major, modal types of finds (not all of which necessarily have similar peaks) alongside tails of older types in decreasing frequencies and rising frequencies of new types. A comprehensive discussion on PN typology as reflected at the Naḥal Zehora sites is found in Gopher (2012, Chapter 40, Sections 40.5 and 40.6) while the social story told by PN material culture is further elaborated in Gopher (2012, Chapter 41). Similarly, PN trends of change (which may also be considered seriations of sorts) may be found also in other aspects. These enhance our picture of the PN and the characteristics of its cultural entities. Economy, for example, shows a continuous change in faunal remains that may be seen, by extension, as a frequency seriation of sorts. Architecture and burial customs, on the other hand, cannot be simply seriated quantitatively. Change in these arenas does not emerge as continuous. Nevertheless, some cultural characteristics may still be traced in these areas, as well.

Thus, according to these arguments, the Yarmukian and Lodian are contemporaneous but geographically distinctive. The differences in typology, as, for example, that of the pottery assemblages, are minor. They reflect regionality rather than chronology. Garfinkel (1999b:101-102) also suggested another entity, the Nizzanim (ware) entity on finds from Givat ha-Parsa and Nizzanim, on the southern coastal plain of Israel (see also Bar-Yosef and Garfinkel 2008:169-170). Surprisingly, the Nizzanim ware is presented both by Garfinkel (1999b) and by Bar-Yosef and Garfinkel (2008) in a similar hierarchical level as the Yarmukian and Lodian. This contention may raise some questions. In my view, the Nizzanim ware is merely a Lodian variant, however, Garfinkel’s singling out this variant is also a statement by which Jericho IX (Mainly as depicted by finds from Jericho but also from Lod, Teluliyot Batashi,

The question of the Lodian The debate concerning the Lodian revolves around the question of its chrono-cultural relationship with the Yarmukian. On this, there is a long history of disagreement. While the Lodian (under the taxon ‘Jericho IX’) has always been considered a separate entity, independent of the Yarmukian, it has not always been considered subsequent to it. For example, Stekelis (1966, 1972) perceived the Yarmukian as the earliest pottery-bearing entity, while Kaplan, at some point in his work (Kaplan 1958a:24 on Teluliyot Batashi, Kaplan 1959b:20, 21, and Kaplan 1977 on Lod) argued, not always 99

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Ghrubba and Khirbet ed-Dharih) is assigned the status of the ‘normative’ Lodian.

IX, or Lodian in my terms; see also Kirkbride 1971). In 1993, when presenting a renewed summary for The New Encyclopaedia of Archaeological Excavations in the Holy Land’, layer 2B was presented as a single unit of the Yarmukian (Perrot 1993). Why this happened is not clear. However, I would venture a guess that this change occurred due to Garfinkel’s studies (1992a, 1992b), which concluded that Munḥata included no Lodian phase (Munḥata phase).

At about the same time, in the early 1990’s, the other, traditional (perhaps conservative) view considering the Yarmukian and Lodian as two sequential entities, was reiterated by Gopher and Gophna (1993) and later by Gopher (1995) and by Gopher and Blockman (2004). This was supported by stratigraphic arguments, geographical arguments, typological considerations and the few available 14C dates. The current debate thus revolves around the question whether the Yarmukian and the Lodian are sequential or contemporaneous entities. In order to resolve the debate, the following questions – presented here in a somewhat different manner than above – should be addressed and answered:

In terms of absolute dates, while the 14C data for the Lodian is far from being satisfactory, Yarmukian dates are clearly shown to be earlier than the Lodian ones (see above and Gopher 2012, Chapter 41, Appendix 41.A). Geographically, the Yarmukian and Lodian appear concurrently in several regions, such as the Jordan Valley, the Jezreel Valley, Israel’s Coastal Plain and the Shephela (for distribution maps of the Yarmukian and the Lodian see Gopher 2012, Figs. 41.a, 41.2b, pp. 1548-49). While in itself this may comprise evidence favorable to the notion of contemporaneity, in the case of the densely populated PN Mediterranean landscapes, the co-existence of two independent and distinct contemporaneous entities seems unlikely. In addition, internal geographical subdivisions are suggested within the Lodian, for example, in the case of the southern coastal plain sites (such as ha-Parsa and Nizzanim – the Nizzanim ware) and in the case of the recently suggested Korenian in the north (Getzov 2008a, 2008b). The ‘southern orientation’ of the Lodian advocated by Garfinkel (1999b; Bar-Yosef and Garfinkel 2008) and also in my past publications (Gopher 1995; Gopher and Gophna 1993) can no longer hold following the discovery of Lodian layers in northern sites such as HaGosherim, Tel Teo, and Yiftaḥel or sites in central Israel such as En Esur and Naḥal Zehora II. Nevertheless, in my view, a potentially additional Lodian variant in the area of the Dead Sea may be suggested. These further support my view of the Lodian as an independent PN entity that post-dated the Yarmukian.

–– Is there or is there not a stratigraphy to show the sequence of these two entities? –– Is the absolute chronology of the two entities similar, showing an identical or overlapping time range? –– Do the two entities show geographical distinction, or, in the case of a geographic overlap – does the stratigraphy of the two entities exhibit any kind of interfingering throughout the distribution, or at least in its border zones? –– Are there similarities in material culture of all sorts, and to what degree? How could such similarities or dissimilarities be measured or given a quantitative value within the cultural entities in which they appear? While I cannot answer these questions in full, I can make the following comments. Lodian stratigraphy was indeed a major source of debate. However, as I have already argued above, wherever the Lodian appears with the Yarmukian, its stratigraphic position is always subsequent to the Yarmukian. Naḥal Zehora II is simply further evidence to this substantiated chrono-stratigraphy. Even in cases where the presence of the two entities, mainly the Lodian, is not clear or accepted by all, no reversed stratigraphy arguing a Yarmukian layer above the Lodian has been reported or proposed. Consequently, neither was any evidence ever presented that may indicate interfingering. Curiously, contemporaneity is argued even in cases where the site’s Yarmukian and Lodian stratigraphy is presented as sequential, as in the case of Ghrubba (see, for example, Obeidat 1995; and see a recent discussion by Kafafi 2011). The Ghrubba case is particularly interesting visà-vis the Lodian question. When Perrot (1968) wrote his preliminary summary of the Munḥata excavations, he subdivided layer 2B into two: a lower, Yarmukian phase and an upper phase which he then named ‘MunḥataGhrubba’ (also known as the Munḥata phase, Jericho

With respect to material culture, seriations carried out for different aspects of pottery, lithics, imagery items, and other finds further support a chronological sequence between these two entities. As a result, the Lodian is established as subsequent to the Yarmukian. Alongside some innovations, Lodian pottery assemblages show certain Yarmukian cultural ‘leftover’ traits, more characteristic of the late Yarmukian than of the early Yarmukian alongside some innovations in the pottery. These include the technological change in pottery-making, mainly the selection of raw material and forming techniques (see Nativ et al. 2012, Chapter 16; Gopher and Eyal 2012, Chapter 17), new vessel shapes especially jars (see Gopher and Eyal 2012, Chapters 9, and 10), new Surface Treatment components such as colored burnished elements, incised elements, and the 100

A. Gopher: Unresolved Pottery Neolithic Chrono-Stratigraphic and Chrono-Cultural issues demise of the herringbone motif (see Gopher and Eyal 2012, Chapters 10, 11, 12).

Are the Wadi Rabah and Powr-PG – Neolithic or Chalcolithic?

Differences among Yarmukian and Lodian practices are also notable in the lithic industry. In the flint assemblage the domination of flakes and the production of arrowheads and especially sickle blades on flakes are notable and more abundant than in the predating Yarmukian (see for example at Lod and Ha-Gosherim V). However, this is a short-lived phenomenon (not present in the postdating Wadi Rabah sickle blades), which seems to be a restricted Lodian trait. At the same time, while Type B sickle blades were already introduced in the late Yarmukian (e.g., Naḥal Qanah Cave and see Gopher and Barkai 2012, Chapter 26), the change from the dominating Type A sickle blades in the Yarmukian to dominating Type B sickle blades in the Lodian is clear (see also Gopher et al. 2001b). Nevertheless, quoting Crowfoot-Payne (1983), Garfinkel considers the Yarmukian and Lodian (Jericho IX, PNA) industries to be very similar. In my view, the typological case presented by these lithic industries, as well as the respective pottery industries, is quite clear. The differences in major aspects of the chaîne opératoire of both pottery vessels and lithic production are indeed sufficient for a distinction. More cultural differences between the Yarmukian and Lodian are manifest in burial customs (jar burials begin only in the Lodian; see Gopher 1996; Gopher and Orrelle 1995b), architectural features (see Gopher 2012, Chapter 6), and imagery items (see Gopher and Eyal 2012, Chapter 29; Gopher 2012, Chapter 40).

The question as to whether late PN entities are Neolithic or Chalcolithic, so to speak, is important in its own right. After all, it involves the end of one of the most dynamic and innovative eras in human history. Its significance is further enhanced when and if transformed into a ‘classical’ question of ‘the transition’ from Neolithic to Chalcolithic. This is an archaeological issue that still requires much thought. According to the manner in which I perceive the definition of culture, this is, in effect, a question of socioeconomics and social (community) worldviews, and how these are expressed in the material record of the entities involved. I will reiterate that the Neolithic spirit, although waning towards its end, effectively continued until the beginning of the Chalcolithic Ghassulian. At that time a new society with a new ethos had emerged. The PN is thus viewed as part of a long-term historical process characterized by significant changes that fully separate it from markers of the earlier hunters-gatherers ethos while bringing the Neolithic Revolution to its full manifestation as reflected in the emergence of Near Eastern agrarian village landscapes. PN developments, it may, therefore, be said, also opened the road to and served as the infrastructure for new developments that occurred first, during the Chalcolithic Ghassulian. That period shows major changes in many realms of life, and later in the emergence of larger-scale urban sociopolitical systems. The history of research concerning this issue starts with statements by Kaplan in the 1950’s (e.g., 1959a: 27). He defined the Wadi Rabah culture, and he assigned it to the Chalcolithic period (his Middle Chalcolithic). Garstang (see Kaplan 1959a: 27) actually suggested a similar terminology much earlier, assigning his Jericho VIII to an early Chalcolithic culture. The term early Chalcolithic was later adopted by Kaplan to describe his Wadi Rabah culture. Kaplan’s following discussion should have argued that the Wadi Rabah culture is closer to the later Chalcolithic (Ghassulian) than to the predating Neolithic (PN) cultures. Instead, he particularly emphasized those points that portrayed the differences and major deviances in the Ghassulian such as geographic distribution of sites, settlement patterns, and a variety of material culture (pottery, lithics) elements. Perhaps due to scarce research of the PN at the time, his view was not immediately adopted. Consequently, the Wadi Rabah was considered by many as a Neolithic entity. His suggestion became accepted only later. However, it is now commonly employed by many PN and Chalcolithic researchers. Notwithstanding, various textbooks and written summaries still present the Wadi Rabah culture as part of the Neolithic period. This may also have to do with the fact that traditionally the Chalcolithic period

Finally, economic aspects of the Yarmukian and Lodian cultures not only show normal Gaussian curves of ontogenetic ‘life’ distributions for a wide variety of elements but also place the Lodian – as established above – between the Yarmukian and the Wadi Rabah cultures (see Davies 2012, Chapters 31; Gopher 2012, Chapter 40). In general, I accept the notion of spatially- or geographically-differentiated contemporaneous entities for both the PN and other cases. Distinct contemporaneous entities coexisted in the desert and the Mediterranean zone. Similarly, about two decades ago, we suggested some coexisting Wadi Rabah variants (Gopher and Gophna 1993). The contemporaneity of the southern Qatifian and the late Wadi Rabah as well as of the Besorian and PoWRPG entities in central and northern Israel have already been mentioned. So also was the contemporaneity of several PoWR-PG entities within the short time span of 200-300 years. Certain geographically distinct Lodian variants also seem to have coexisted (see above). However, the case of the Yarmukian and Lodian does not meet the minimum criteria necessary to support a contemporaneity claim. 101

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead was excavated and taught by Bronze and Iron Age archaeologists. These scholars were not involved in debates concerning the issue. The issue of transition was then complicated when Meyerhof (1982), following his small-scale excavation at the site of Ein el Jarba, suggested what he called a Proto-Ghassulian entity. He placed this entity between the Wadi Rabah culture and what we would call the Chalcolithic Ghassulian.

and the abundance of obsidians. Seeking ‘hybrids indicative of a transitional phase’ (Garfinkel 2009:328) from the PN (Yarmukian and Lodian) to Wadi Rabah but failing to find any, he concludes that the Wadi Rabah entity is significantly different from its precursors. Thus, Garfinkel argues that the Wadi Rabah entity may be perceived as ‘a revolution in material culture of the southern and central Levant’ (Garfinkel 2009:327).

Garfinkel, both in the 1990’s (Garfinkel 1992a, 1992b, 1999b) and recently (e.g., Garfinkel 2009) revisited Kaplan’s suggestion, and he offered a three-partite division of the Chalcolithic period – namely, Early, Middle, and Late Chalcolithic. Garfinkel’s Early Chalcolithic includes what I refer to as ‘normative’ Wadi Rabah as well as certain ‘variants’ (Gopher and Gophna 1993) which I would assign to either early or late Wadi Rabah. Garfinkel’s list of sites includes Wadi Rabah itself, Munḥata 2A, Neveh Yam, Teluliyot Batashi, Ein el Jarba, Ha-Gosherim IV, Tell Abu Zureiq, Ha-Bashan St., Tel Ali 1c, Lod (old excavation by Kaplan), Tel Dan, Naḥal Beẓet I, Ḥorbat Uza 19, En Esur VI, Naḥal Zehora II and I and Jericho VIII. Garfinkel dates this entity to 7750-7250 Cal BP. Garfinkel’s Middle Chalcolithic, dated to 7250-6550 cal BP, includes what I see as late manifestations of the Wadi Rabah culture (sensu lato, e.g., Gopher 1995; Gopher and Gophna 1993) as well as post-Wadi Rabah entities. Because of its temporal range, this Middle Chalcolithic may be seen as a major period or entity in the sequence of the late prehistory of the region. Garfinkel (1999b) lists many Middle Chalcolithic sites under what he calls the Beth Shean ware such as Beth Shean XVIII, Tel Ali 1b, Tel Ẓaf, Tell esh-Shunah, Tell Abu Ḥabil, Tell esSaidiyeh et Taḥta, Tell el-Mafjar, Tell el-Farah North, Kabri, Teleilat Ghassul, Kataret es-Samra, Tel Qiri, Ein el Jarba, Tel Dan, Kefar Galim, Tel Teo 8-10, Ḥorbat Uza 1816, Hayonim Terrace, Ṭabaqaṭ al Bûma, Abu Ḥamid, En Esur Vb-c, and Kefar Giladi. Garfinkel’s Late Chalcolithic refers to the classical Chalcolithic Ghassulian and the northern Chalcolithic Golanian cultures [6550-5550 cal BP].

This seems to serve as the basis for his periodization, in which the assignment of post-Wadi Rabah entities to what Garfinkel calls the Middle Chalcolithic period that preceded the Chalcolithic Ghassulian (his Late Chalcolithic) comes naturally and with no questions. The issue of PoWR-PG gained further momentum in the 1990’s and 2000’s. Following quite intensive fieldwork in a few sites revealing PoWR-PG finds, a round table discussion on the issue of post-Wadi Rabah yet preGhassulian entities took place in Tel Aviv in 2002, under the auspices of the Israel Antiquities Authority. Generally speaking, the notion was accepted that independent entities existed subsequent to the Wadi Rabah culture and prior to the Chalcolithic Ghassulian. It has been agreed upon that these entities show specific characteristics in the different geographical regions of the southern Levant. Yannai (2002, 2006) suggested calling the regional variant of this entity found in central and western Israel the Nazurian culture, following his excavation and study of the finds of Nazur 4 (see also Golan 2006). Getzov (2009a), following his analysis of the stratigraphy and pottery assemblages of Ḥorbat Uza, suggested a similar chronology mainly with respect to strata 18, 17 (my late Wadi Rabah), and 16 (my PoWR-PG) and more. Discussions held in Madrid in the spring of 2006 (see Lovell and Rowan 2011 and specifically Rowan and Lovell 2011) and the ensuing issue of Paléorient [33(1), 2007] a year later, also devoted to these discussions, revealed terminological disorder, disagreement concerning the cultural assignment of relevant assemblages, and varying interpretations of the 14C dates. Two additional recent papers by Gilead and by Garfinkel in a book of essays in honor of O. Bar-Yosef (Shea and Lieberman 2009) summarize two views on the issue, one concentrating on southern entities and one on northern ones. An additional thematic issue of Paléorient [36(1), 2010] was devoted to the social developments in the sixth and fifth millennia calBC in the Levant, including Turkey and Iran. However, the current state of this debate remains quite confusing and in many ways even frustrating.

While referring to the question whether the Wadi Rabah should be considered as Neolithic or Chalcolithic as merely a semantic question, Garfinkel (Garfinkel 1999b:6) consistently refers to the Wadi Rabah culture as a Chalcolithic rather than as a Neolithic entity. This discretion is based mainly on the typology of the Wadi Rabah pottery assemblages (e.g., Garfinkel 1999b; and see also Garfinkel 1992a; Garfinkel et al. 2007; Garfinkel and Matskevitch 2002; Getzov 2009a; Getzov et al. 2009b; Yannai 2006, 2008), but also on other aspects of material culture characterizing the Wadi Rabah culture. In his 2009 paper, Garfinkel lists additional Wadi Rabah characteristics, which set apart this entity from earlier Neolithic (Yarmukian and Lodian) entities. These characteristics include the lithic assemblages, the lack of Yarmukian ‘art objects’, the presence of sling stones,

The view on the PoWR-PG presented in Fig.1 is based on the limited scope of the PoWR-PG entities (sometimes only a single site per entity or sub-entity, or at best a small group of sites), their regional variation, and a far 102

A. Gopher: Unresolved Pottery Neolithic Chrono-Stratigraphic and Chrono-Cultural issues from sufficient 14C record. Nevertheless, the PoWR-PG can be dated to 6800/6700-6500 cal BP based on the dates from Tel Ali 1b and Tel Ẓaf [see Gopher 2012, Chapter 41]. The dates on olive stones from Naḥal Zehora II can be added here, despite their limitations (see Carmi and Boaretto. 2012; Gopher 2012, Chapter 40)]. Thus, we may allow for a rather short ‘transition’ from what I consider as the Neolithic, including the Wadi Rabah culture, to the Chalcolithic Ghassulian. This short time frame must be addressed before any attempt is made to understand the origins of the innovative Chalcolithic Ghassulian (e.g., Rowan and Golden 2009) and the history of the southern Levant after both the PN and the Chalcolithic periods. A comparable and similarly short cultural transition (100-200 years) to the Chalcolithic Ghassulian is manifested by the Besorian in the south of the southern Levant (Gilead 2009, 2011). Gilead also assumes an equivalent development in the central and northern parts of Israel.

distribution curves reflecting trends of change in material culture elements (e.g., typology) that show tails of both old decreasing elements and new rising elements. The ‘modal’ cultural elements do not necessarily constitute a high percentage of the finds. This is the case with respect to the Wadi Rabah entity including its later manifestations1 as it evolved into the small subregional PoWR-PG entities that exhibit a significantly different character and a new spirit when compared to the preceding Wadi Rabah (at best, the PoWR-PG assemblages yield only a little of the Wadi Rabah material culture elements). These, in turn, gave way to a new quite encompassing cultural entity, the Ghassulian (and its contemporaneous entities in the north), which again displays a distinctly unique and clearly defined character portrayed by a distinct (averaged) normal Gaussian distribution curve of its own. It requires the accumulation of a critical mass of change before a new entity can be defined, even within a single period (for example, see comments on the Natufian within the Epipaleolithic period in Gilead 2009:335). While change is easily expressed in the material record, a prerequisite for the definition of a new cultural entity is that the record also exhibits a significant and clearly detected change in socioeconomics and worldviews (ideology). The differences between the Chalcolithic Ghassulian and the Wadi Rabah and PoWR-PG are significant and sufficient for the definition of the Ghasulian as an independent, new culture marking the beginning of a new period. These include the following:

By contrast, the period that Garfinkel considers as transitional is a rather long one, spanning some 700 years (Garfinkel 2009). In his view, it is the period, which he calls ‘Middle Chalcolithic’ (Garfinkel 1999b) that manifests the transition from his early Chalcolithic (Wadi Rabah) to the Chalcolithic Ghassulian. This is a long and varied transition. It is longer than any of the PN cultural components, in which are included cases that in my view pertain to the Wadi Rabah, late Wadi Rabah, or PoWR-PG entities. Consequently, I would like to say a word on this transition. I am more comfortable with views presented by Gilead (2009, 2011), and to a certain extent also by Rosen (2011), on periodization, cultural definitions and cultural terminology, as well as their understanding of the nature of transitions than I am with Garfinkel’s (2009) statements (see also Rowan and Lovell 2011). However, I still see transitions in prehistory in a different, perhaps extreme, manner.

–– With respect to agriculture and subsistence economy, the Chalcolithic Ghassulian saw the full establishment of agriculture, including crop farming and animal husbandry with virtually no hunting, and the simultaneous establishment of pastoralists in marginal areas. During this time, the Secondary Product Revolution (SPR; Sherratt 1981, 1983) was further rooted extending to the establishment of the olive oil industry, new domesticated plant species, and transportation animals. Floodwater farming and check dams were another significant innovation of the Chalcolithic Ghassulian (Levy 1995:230).

In my view, the notion of a transitional culture ingrains a testimonium paupertatis for archaeology and archaeologists alike, both theoretically and methodologically. For me, transitions are a permanent human condition; transition is a state of mind – the sound of the dynamics of life and history, which we utterly fail to compartmentalize into our static nomenclatures and rigid subdivisions. This is why, in my view, a major normative (modal) appearance is expected for any archaeological entity/culture, which would also include some variants (geographical, chronological, economic or other). An archaeological culture is a dynamic, relatively long-lasting entity, whose variability should not be dismissed (or used unsystematically to define new entities) as long as the material culture evidence testifies to the presence of that culture (in as much as a systematic polithetic assignment allows). It is thus only to be expected, in my view, that the (quantitative) analyses of finds commonly show normal Gaussian

Late Wadi Rabah assemblages are prominently characterized by Wadi Rabah elements, albeit in frequencies that are lower – or in some cases higher (e.g., flat loop strap handles or plastic additives as Surface Treatment) – than the ‘normative’ Wadi Rabah levels. In pottery these elements include DFBW, bow rim jars, carinated vessels, cut or flattened rims, characteristic incised elements and patterns, and other distinctive elements; in flint they include typical rectangular backed and truncated sickle blades and few typical arrowheads; and they are also manifested in other material culture assemblages. It is the frequencies of these elements and their extent compared to the earlier Wadi Rabah culture and the subsequent Chalcolithic Ghassulian that mark these sites such as Naḥal Zehora I, as late in the Wadi Rabah sequence. 1 

103

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead –– Chalcolithic Ghassulian pottery shows many innovative aspects, including a new selection of raw materials (paste and temper), a new firing technique, and many typological innovations such as cooking pots, spoons, the cornet, the churn, ossuaries for the dead, and additional changes detected in other typological groups, as well. Chalcolithic Ghassulian Surface Treatment is not as rich as it is in PN assemblages, and the pottery vessels as a whole becomes more standardized. –– Lithic industries of the Chalcolithic Ghassulian show continuity in aspects such as blade production for sickle blades and the production of bladelets (known in Wadi Rabah and PoWRPG assemblages). However, in other aspects they show distinct characteristics. Arrowheads are practically absent from Chalcolithic Ghassulian assemblages, and sickle blades show some minor change. The bifacial tool assemblages are dominated by the versatile woodworking adze, and fan scrapers become an important tool type. In addition, there emerges a series of outstandingly unique perforated flint discs, some of which exhibit conspicuous spikes. Should the claim be confirmed that the Canaanean blade technology also emerged during the Chalcolithic Ghassulian (even if only at its very late stages) (Bar and Winter 2010; Rowan and Levy 1994), this would be another major and most important innovation of the Chalcolithic Ghassulian lithic industry (but see Milevski et al. 2011). –– Imagery items of the Chalcolithic Ghassulian portray a whole new world of symbols manifested in new media, including wall paintings, metal artefacts, ivory artefacts and stone (violin) figurines [Some items found in PoWR-PG sites (e.g., ḤorBat Uza Stratum 16, Getzov 2009b:101) may possibly be precursors of the classical, well known violin figurines of the Chalcolithic Ghassulian]. A whole new world of imagery is also expressed in Chalcolithic Ghassulian Surface Treatment, especially the plastic or sculpted elements found on ossuaries. Among the northern Golanian entities paralleling the Ghassulian, stone (basalt) anthropomorphous figures are another innovative medium also exhibiting new content. –– A new and significant industry of metal emerged in the Chalcolithic Ghassulian. This is a rich and complex industry of high quality items, including many non-functional (non-working) items that testify to a completely new approach to materials, new techniques (lost wax), and a new system of reference. –– Burial customs of the Chalcolithic Ghassulian have become crystallized compared to earlier times, and separate designated cemeteries

emerged outside residential sites. New burial features of the Chalcolithic Ghassulian include structures, caves (both natural and artificial), clay or stone containers for the dead, ossuaries, and more. Onsite burial during this period had become marginal and seems to have been limited mostly to young children as infants and children were not buried in ossuaries. –– Chalcolithic Ghassulian architecture has become well established with familiar rectangular houses and compounds but also with previously unknown elements such as the underground systems of the Beer Sheva area. In an overall view, then, it is quite clear that the Chalcolithic Ghassulian displays distinct characteristics in all realms compared to the Wadi Rabah and PoWR-PG material world, reflecting its full divorce from the little that was left from the older Paleolithic and Neolithic worldviews and conduct. In other words, the PoWR-PG along with the other PN entities mentioned above, are all Neolithic entities, the last cultures predating the Ghassulian of the Chalcolithic period. Thus, following my earlier contentions, the Neolithic ended upon the full establishment of the Neolithic Revolution and once evidence of the hunter-gatherer ethos completely disappeared from among cultural remains. A new era followed the PN, in which there unfolded a new world characterized by new socioeconomic systems, a material culture that presented innovative industries, a new worldview and a new symbolic language. According to the view presented here, the Chalcolithic period thus began with the Ghassulian. References Anati, E., Avnimelech, M., Hass, N. and Meyerhof, E. 1973. ‘Hazorea I’ (Archivi Vol. 5). Capo di Ponte, Italy, Edizioni de Centro. Bar-Yosef, O. and Garfinkel, Y. 2008. The Prehistory of Israel, Human Cultures Before Writing. Jerusalem, Ariel (Hebrew). Bar, S. and Winter, H. 2010. Canaanean Flint Blades in Chalcolithic Context and the Possible Onset of the Transition to the Early Bronze Age: A Case Study from Fazael 2. Tel Aviv 37:33-47. Barzilai, O. and Garfinkel, Y. 2006. Bidirectional Blade Technology after the PPNB: New Evidence from Sha’ar Hagolan, Israel. Neo-Lithics 1 (6):27-31. Barzilai, O. and Getzov, N. 2008. Mishmar Ha’emeq: A Neolithic Site in the Jezreel Valley. Neo-Lithics 2 (8):12-17. Bourke, S. J. 2007. The Late Neolithic/Early Chalcolithic Transition at Teleilat Ghassul: Context, Chronology and Culture. Paléorient 33(1):15-33. 104

A. Gopher: Unresolved Pottery Neolithic Chrono-Stratigraphic and Chrono-Cultural issues Braun, E. 1997. Yiftah’el, Salvage and Rescue Excavations at a Prehistoric Village in Lower Galilee, Israel. Jerusalem, Israel Antiquities Authority. Burton, T. and Levy, E. 2001. The Chalcolithic Radiocarbon Record and its Use in Southern Levantine Archaeology. Radiocarbon 43(3):1223-1246. Carmi, I. and Boaretto, E. 2012. Radiocarbon dating of Nahal Zehora II. In Gopher, A. Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora: Volumes I-III. Chapter 39,. 1474-1480. Emery and Clair Yass Publications in Archaeology, Tell Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29. Crowfoot-Payne, J. 1983. The flint Industries of Jericho. In K. M. Kenyon and T. A. Holland (eds), Excavations at Jericho V: The Pottery Phases of the Tell and Other Finds: 622-759. Oxford: The British School of Archaeology in Jerusalem and Oxford University Press.. Davies, S. J. M. 2012. Animal bones at the Nahal Zehora sites. In A. Gopher, Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora: Volumes I-III, Chapter 31, 1258-1320. Emery and Clair Yass Publications in Archaeology, Tell Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29. Dollfus, G. and Kafafi, Z. 1993. Recent Research at Abu Hamid. Annual of the Department of Antiquities of Jordan 37: 241-263. Amman: Department of Antiquities. Eisenberg, E., Gopher, A. and Greenberg, R. (eds) 2001. Tel Teo, a Neolithic, Chalcolithic and Early Bronze Age site in the Hula Valley. Israel Antiquities Authority Reports 13. Jerusalem, Israel Antiquities Authority. Garfinkel, Y. 1992a. The Material Culture in the Central Jordan Valley in the Pottery Neolithic and Early Chalcolithic Periods. Unpublished PhD. dissertation, The Hebrew University of Jerusalem. Garfinkel, Y. 1992b. The Pottery Assemblage of the Sha’ar haGolan and Rabah Stages of Munhata (Israel). Les Cahiers du Centre de Recherche Français de Jérusalem 6. Paris: Association Paléorient. Garfinkel, Y. 1993. The Yarmukian culture in Israel. Paléorient 19(1):115-134. Garfinkel, Y. 1999b. Neolithic and Chalcolithic pottery of the Southern Levant. Qedem, Monographs of the Institute of Archaeology 39. Jerusalem, Institute of Archaeology, the Hebrew University of Jerusalem. Garfinkel, Y. 2008. The Chronology of the Eighth Millennium BP. In Y. Garfinkel and D. Dag, (eds), Neolithic Ashkelon: 15-21. Qedem, Monographs of the Institute of Archaeology 47. Jerusalem, Institute of Archaeology, the Hebrew University of Jerusalem. Garfinkel, Y. 2009. The Transition from Neolithic to Chalcolithic in Southern Levant: The Material Culture Sequence. In J. J. Shea and D. E. Lieberman (eds), Transitions in Prehistory, Essays in Honor of Ofer

Bar-Yosef: 325-333. American School of Prehistoric Research Monograph Series. Oxford and Oaksville, Oxbow Books. Garfinkel, Y. and Ben-Shlomo, D. 2009. Sha’ar Hagolan, The Rise of Urban Concepts in the Ancient Near East. Qedem Reports 9. Jerusalem, Institute of Archaeology, the Hebrew University of Jerusalem. Garfinkel, Y., Ben-Shlomo, D., Freikman, M. and Vered, A. 2007. Tel Tsaf: The 2004-2006 Excavation Seasons. Israel Exploration Journal 57 (1):1-33. Garfinkel, Y. and Matskevitch, Z. 2002. Abu Zureiq, a Wadi Raba Site in the Jezreel Valley: Final Report of the 1962 Excavations. Israel Exploration Journal 52 (2):129-166. Garfinkel, Y. and Miller, M. A. 2002a. The Archaeology of Sha’ar Hagolan. In Y Garfinkel and M. A. Miller, M.A. (eds), Sha’ar ha-Golan, Vol. I: Neolithic Art in Context: 10-34. Oxford: Oxbow Books. Getzov, N. 2008a. Ha-Gosherim. In E. Stern, (ed.), The New Encyclopedia of Archaeological Excavations in the Holy Land: 1759-1761. Jerusalem, Israel Exploration Society and Carta. Getzov, N. 2008b. Sites of the Pottery Neolithic Period in Northern Eretz Israel (abstract) (Lecture delivered at the Annual Conference of the Israel Prehistoric Society, Jerusalem, December 2008, pp. 10-11). Getzov, N. 2009a. Horbat Uza in the Neolithic and Chalcolithic Periods. In N. Getzov, R. LiebermanWander, H. Smithline. and D. Syon, (eds), Horbat Uza: The 1991 Excavations: 102-107. Israel Antiquities Authority Reports 41, Vol. I: The Early Periods). Jerusalem, Israel Antiquities Authority. Getzov, N. 2009b. Miscellaneous finds form the Neolithic and Chalcolithic periods. In N. Getzov, R. LiebermanWander, H. Smithline. and D. Syon, (eds), Horbat Uza: The 1991 Excavations: 98-102. Israel Antiquities Authority Reports 41, Vol. I: The Early Periods. Jerusalem, Israel Antiquities Authority. Getzov, N., Barzilai, O., Le Dosseur, G., Eirikh-Rose, A., Ktalav, I., Marder, O. et al. 2009a. Nahal Betzet II and Ard El Samra: Two Late Prehistoric Sites and Settlement Patterns in the Akko Plain. Mitekufat Haeven, Journal of the Israel Prehistoric Society 39:81158. Getzov, N., Lieberman-Wander, R., Smithline, H. and Syon, D. (eds) 2009b. Horbat Uza: The 1991 Excavations. Israel Antiquities Authority Reports 41, Vol. I: The Early Periods. Jerusalem, Israel Antiquities Authority. Gilead, I. 1990. The Neolithic-Chalcolithic Transition and the Qatifian of the Northern Negev and Sinai. Levant XXII:47-63. Gilead, I. 2004. The Besorian and Other Pre-Ghassulian Entities. Proceedings of the Meeting of the Society for American Archaeology. Gilead, I. 2006. Fifth Millennium Culture History: Ghassulian and other Chalcolithic Entities in the Southern Levant. Lecture delivered at the 5th International 105

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Conference on the Archaeology of the Ancient Near East, Madrid, March 2006. Gilead, I. 2007. The Besorian – A pre-Ghassulian cultural entity. Paléorient 33(1):33-50. Gilead, I. 2009. The Neolithic-Chalcolithic transition in the southern Levant: Late Sixth-Fifth Millennium Culture History. In J. J. Shea, and E. E. Lieberman (eds), Transitions in Prehistory, Essays in Honor of Ofer Bar-Yosef: 335-355. American School of Prehistoric Research Monograph Series. Oxford and Oaksville, Oxbow Books. Gilead, I. 2011. Chalcolithic Culture History: Ghassulian and Other Entities in the Southern Levant. In J. L. Lovell and Y. M. Rowan (eds), Culture, Chronology and the Chalcolithic, Theory and Transition: 12-24. Levant Supplementary Series 9. London, Oxbow Books. Gilead, I. and Alon, D. 1988. Excavations of Protohistoric Sites in the Nahal Besor and the Late Neolithic of the Northern Negev. Mitekufat Haeven, Journal of the Israel Prehistoric Society 21:109-137. Golan, S. 2006. The Nazur Culture, A Lithic perspective: A Case Study from the Site Nazur 4 – The End of the Pottery Neolithic or the Beginning of the Chalcolithic. Unpublished MA thesis, Tel Aviv University. Gopher, A. 1989b. The Flint Assemblages of Munhata (Israel), Final Report. Les Cahiers du Centre de Recherche Français de Jérusalem 4. Paris, Association Paléorient. Gopher, A. 1989c. Horvat Galil and Nahal Betzet I: Two Neolithic Sites in the Upper Galilee. Mitekufat Haeven, Journal of the Israel Prehistoric Society 22:82-92. Gopher, A. 1993. Sixth-Fifth Millennia B.C. Settlements in the Coastal Plain, Israel. Paléorient 19(1):55-63. Gopher, A. 1995. Early Pottery-Bearing Groups in Israel – The Pottery Neolithic period. In T. E. Levy, (ed.), The Archaeology of Society in the Holy Land: 205-225. London, Leicester University Press. Gopher, A. 1996. Infant Burials in the Neolithic Period in the Southern Levant – Israel: A Social View. In M. Otte (ed.), Nature et culture, actes du colloque international 13-17 décembre 1993: 913-918. Etudes et Recherches Archéologiques de l’Université de Liège 68). Liege, Service de Préhistoire, Univiversité de Liège. Gopher, A. 2012. Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora. Emery and Clair Yass Publications in Archaeology, Tell Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29, Volumes I-III. Gopher, A. 2012. Nahal Zehora II. In A. Gopher, Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora: Volumes I-III, 64-107. Emery and Clair Yass Publications in Archaeology, Tell Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29.

Gopher, A. 2012. Architecture and the PN Nahal Zehora Sites. In A. Gopher, Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora: Volumes I-III, 252-291. Emery and Clair Yass Publications in Archaeology, Tell Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29. Gopher, A. 2012. Ground stone tools in the Nahal Zehora sites. In A. Gopher, Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora: Volumes I-III, 1035-1100. Emery and Clair Yass Publications in Archaeology, Tell Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29. Gopher, A. 2012. Village communities in the Menashe Hills: The PN at the Nahal Zehora sites. In A. Gopher, Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora: Volumes I-III, 1484-1524. Emery and Clair Yass Publications in Archaeology, Tell Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29. Gopher, A. 2012. The pottery Neolithic in the southern Levant – A second Neolithic Revolution. In A. Gopher, A. Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora: Volumes I-III, 1525-1611. Emery and Clair Yass Publications in Archaeology, Tell Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29. Gopher, A. and Eyal, R. 2012. In A. Gopher, Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora: Volumes I-III; Chapter 9, 338-358; Chapter 10,359-523; Chapter 11, 524-554; Chapter 12, 555-580; Chapter 17, 697-744; Chapter 26, 1110-1134; Chapter 29, 1170-1244. Emery and Clair Yass Publications in Archaeology, Tell Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29. Gopher, A., Barkai, R. and Assaf, A. 2001b. Trends in Sickle Blades Production in the Neolithic of the Hula Valley, Israel. In I. Caneva, C. Lemorini, D. Zampetti and P. Biagi (eds), Beyond Tools: Redefining the PPN Lithic Assemblages of the Levant: 411-425. Berlin: Ex Oriente. Gopher, A. and Blockman, N. 2004. Excavations at Lod (Nevé Yaraq) and the Lodian Culture of the Pottery Neolithic Period. ‘Atiqot 47:1-50. Gopher, A. and Gophna, R. 1993. Cultures of the Eighth and Seventh Millennia BP in the Southern Levant: A Review for the 1990’s. Journal of World Prehistory 7:297-353. Gopher, A. and Orrelle, E. 1995a. The Groundstone Assemblages of Munhata, a Neolithic Site in the Jordan 106

A. Gopher: Unresolved Pottery Neolithic Chrono-Stratigraphic and Chrono-Cultural issues Valley – Israel, a Report. Les Cahiers des Missions Archaéologiques Françaises en Israël 7. Paris: Association Paléorient. Gopher, A. and Orrelle, E. 1995b. New Data on Burials from the Pottery Neolithic Period (Sixth-Fifth Millennium BCE) in Israel. In S. Campbell and A. Green (eds), The Archaeology of Death in the Ancient Near East: 24-28. (Oxbow Monograph 51). Oxford, Oxbow Books. Goren, Y. 1990. The ‘Qatifian Culture’ in Southern Israel and Transjordan: Additional Aspects for its Definition. Mitekufat Haeven, Journal of the Israel Prehistoric Society 23:100*-112*. Goren, Y. 1991. The Beginnings of Pottery Production in Israel: Technology and Typology of Proto-Historic Ceramic Assemblages in Eretz-Israel (6th-4th Mllennia B.C.E.). Unpublished PhD dissertation, The Hebrew University of Jerusalem. (Hebrew with English summary). Joffe, A. H. and Dessel, J. P. 1995. Redefining Chronology and Terminology for the Chalcolithic of the Southern Levant. Current Anthropology 36 (3):507518. Kafafi, Z. 1987. The Pottery Neolithic in Jordan in Connection with other Near Eastern Regions. In A. Hadidi (ed.), Studies in the History and Archaeology of Jordan Vol. III: 33-39. Amman, Department of Antiquities/ London, Routledge and Kegan Paul. Kafafi, Z. 1998. The Late Neolithic in Jordan. In D. D. Henry (ed.), The Prehistoric Archaeology of Jordan: 127138. British Archaeological Reports International Series 705. Oxford. Archaeopress. Kafafi, Z. 2011. Ghrubba: Ware or Culture? In J. L. Lovell and Y. M. Rowan (eds), Culture, Chronology and the Chalcolithic, Theory and Transition: 25-35. Levant Supplementary Series 9. London, Oxbow Books. Kaplan, J. 1958a. Excavations at Teluliot Batashi, Vale of Soreq. Eretz Israel 5:9-24 (Hebrew with English summary). Kaplan, J. 1958b. Excavations at Wadi Rabah. Israel Exploration Journal 8:149-160. Kaplan, J. 1959a. The Archaeology and History of Tel AvivJaffa. Tel Aviv: Massada (Hebrew). Kaplan, J. 1959b. The Neolithic Pottery of Palestine. Bulletin of the American Schools of Oriental Research 156:15-22. Kaplan, J. 1970. The Excavation of Habashan Street. In B. Mazar (ed.), Encyclopedia of Archaeological Excavations in the Holy Land (Vol. 2). Jerusalem, Israel Exploration Society and Massada: 563 (Hebrew). Kaplan, J. 1972a. The Archaeology and History of TelAviv-Jaffa. Biblical Archaeologist 35:66-95. Kaplan, J. 1972b. Twenty Years Since the Discovery of the Chalcolithic Wadi Rabah Culture. The Ha’Aretz Museum Annual 14:9-13 (Hebrew). Kaplan, J. 1977. Neolithic and Chalcolithic Remains at Lod. Eretz Israel 13:57-75 (Hebrew with English summary).

Kenyon, K. M. 1957. Digging up Jericho. London, Ernest Benn. Kenyon, K. M. 1960. Archaeology in the Holy Land (1st ed.). London, Ernest Benn Kenyon, K. M. 1970. Archaeology in the Holy Land (3rd ed.). London, London, Ernest Benn Kenyon, K. M. 1981. Excavations at Jericho III: The Architecture and Stratigraphy of the Tell. Oxford, British School of Archaeology in Jerusalem. Kenyon, K. M. and Holland, T. A. (eds) 1983. Excavations at Jericho V: The Pottery Phases of the Tell and Oher Finds. Oxford, British School of Archaeology in Jerusalem. Khalaily, H. and Marder, O. 2003. The Neolithic Site of Abu Gosh: The 1995 Excavations. Israel Antiquities Authority Reports 19. Jerusalem, Israel Antiquities Authority. Khalaily, H., Milevski, I., Getzov, N., Hershkovitz, I., Barzilai, O., Yarosevich, A. et al. 2008. Recent Excavations at the Neolithic Site of Yiftahel (Khalet Khalladyiah), Lower Galilee. Neo-Lithics 2(8):3-11. Khalaily, H. and Nagorsky, A. 2013. Tel Hanan – a Wadi Rabah Site East of Haifa. ‘Atiqot 73:1-17. Khalaily, M. 1999. The flint assemblage of Layer V at Hagoshrim: A Neolithic Assemblage of the Sixth Millennium B.C. in the Hula Basin. Unpublished MA thesis, Hebrew University of Jerusalem (Hebrew). Kirkbride, D. 1971. A Commentary on the Pottery Neolithic of Palestine. Harvard Theological Review 64:281-289. Kuijt, I. and Chesson, M. S. 2002. Excavations at ‘Ain Waida’, Jordan, New Insights into Pottery Neolithic Lifeways in the Southern Levant. Paléorient 28(2):111-124. Levy, T. E. 1995. Cult, Metallurgy and Rank societies – Chalcolithic Period (ca. 4500-3500 BCE). In T. E. Levy (ed.), The Archaeology of Society in the Holy Land. London, Leicester University Press: 226-243. Levy, T. E. and Burton, M. 2006. Appendix 2: Radiocarbon Dating of Gilat. In T. E. Levy (ed.), Archaeology, Anthropology and Cult: The Sanctuary at Gilat, Israel. London, Routledge: 863-866. Lovell, J. L., Dollfus, G. and Kafafi, Z. 2007. The Ceramics of the Late Neolithic and Chalcolithic: Abu Hamid and the Burnished Tradition. Paléorient 33(1):51-76. Lovell, J. L. and Rowan, Y. M. 2011. Culture, Chronology and the Chalcolithic, Theory and Transition. Levant Supplementary Series 9. London, Oxbow Books. Meyerhof, E. 1982. Ein el Jarba (preliminary report). Mitekufat Haeven, Journal of the Israel Prehistoric Society 17:79-85 (Hebrew). Milevski, I., Fabian, P. and Marder, O. 2011. Canaanean Blades in Chalcolithic Contexts of the Southern Levant? In J. L. Lovell and Y. M. and Rowan (eds), Culture, Chronology and the Chalcolithic, Theory and Transition: 149-159. Levant Supplementary Series 9. London, Oxbow Books. Milevski, I., Khalaily, H., Getzov, N. and Hershkovitz, I. 2008. The Plastered Skulls and other Pre-Pottery 107

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Neolithic B Finds from Yiftahel, Lower Galilee (Israel). Paléorient 34(2):37-46. Nativ, A., Gopher, A. and Goren, Y. 2012. Pottery Production at Nahal Zehora II. In A. Gopher, Village Communities of the Pottery Neolithic Period in the Menashe Hills, Israel. Archaeological Investigations at the Sites of Nahal Zehora: Volumes I-III, Chapter 16. 657-696. Emery and Clair Yass Publications in Archaeology, Tel Aviv University, Sonia and Marco Nadler Institute of Archaeology Monograph Series number 29. Obeidat, D. 1995. Die Neolithische Keramik aus Abu Thawwab, Jordanien. Studies in Early Near Eastern Production, Subsistence and Environment 2. Berlin: Ex Oriente. Perrot, J. 1968. La préhistoire Palestinienne. Dictionnaire de la Bible, Vol. 8: Supplément, 285-446. Paris, Letouzey et Ané. Perrot, J. 1993. Beer Sheva, the Chalcolithic Sites. In E. Stern (ed.), The New Encyclopedia of Archaeological Excavations in the Holy Land, Vol. 1: 161-163. Jerusalem, Carta. Prausnitz, M. 1970. From Hhunter to Farmer to Trader. Jerusalem, Sivan. Rollefson, G. O. and Köhler-Rollefson, I. 1992. Exploitation Patterns in The Early Neolithic Levant: Cultural Impact On The Environment. Population and Environment 3 (4):243-254. Rosen, S. A. 2011. Desert Chronologies and Periodization Systems. In J. L. Lovell. and Y. M. Rowan (eds), Culture, Chronology and the Chalcolithic, Theory And Transition: 71-83. Levant Supplementary Series, 9. London, Oxbow Books.. Rowan, Y. M. and Golden, J. 2009. The Chalcolithic Period of the Southern Levant: A Synthetic Review. Journal of World Prehistory 22(1):1-95. Rowan, Y. M. and Levy, T. E. 1994. Proto-Canaanean Blades of the Chalcolithic Period. Levant 26:167-175. Rowan, Y. M. and Lovell, J. L. 2011. Introduction: Culture, Chronology and the Chalcolithic. In J. L. Lovell

and Y. M. Rowan (eds), Culture, Chronology and the Chalcolithic, Theory And Transition: 149-150. Levant Supplementary Series, 9. London, Oxbow Books. Shea, J. J. and Lieberman, D. E. (eds) 2009. Transitions in Prehistory. Essays in honor of Ofer Bar-Yosef. American School of Prehistoric Research Monograph Series. Oxford and Oaksville, Oxbow Books. Sherratt, A. G. 1981. Plough and Pasturalism: Aspects of the Second Products Revolution. In I. Hodder, G. Isaac and N. Hammond (eds), Pattern of the Past. Studies in Honour of David Clarke: 261-305. Cambridge, Cambridge University Press. Sherratt, A. G. 1983. The Secondary Exploitation of Animals in the Old World. World Archaeology 15:90103. Simmons, A. H., Rollefson, G. O., Kafafi, Z., Mandel, R.O., al-Nahar, M., Cooper, J. et al. 2001. Wadi Shu’eib, A Large Neolithic Community in Central Jordan: Final Report of Test Investigations. Bulletin of the American Schools of Oriental Research 321:1-39. Stekelis, M. 1966. The Yarmukian Culture. Sha’ar HaGolan, Yarmukian Museum of Shaar HaGolan (Hebrew). Stekelis, M. 1972. The Yarmukian culture of the Neolithic period. Jerusalem, Magnes Press. Yannai, E. 2002. The Northern Sharon in the Chalcolithic Period and the Beginning of the Early Bronze Age in Light of the Excavations at ‘Ein Asawir: 65-86. In E. C. M. van den Brink. and E. Yannai (eds), Quest of ancient settlements and landscapes. Archaeological Studies in Honour of Ram Gophna. Tel Aviv, Tel Aviv University/ Ramot Publishing. Yannai, E. 2006. ‘En Esur (‘Ein Asawir) I, Excavations at a protohistoric site in the coastal plain of Israel. Israel Antiquities Authority Reports 31. Jerusalem, Israel Antiquities Authority. Yannai, E. 2008. Esur, Tel. In E. Stern (ed.), The New Encyclopedia of Archaeological Excavations in the Holy Land: 1732-1734. Jerusalem, Israel Exploration Society and Carta.

108

Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant, ca. 4500-3900 BC Milena Gošić The present paper discusses chronological aspects of changes in ritual practices of the Ghassulian culture. Both chronology and ritual behavior of the Ghassulian culture are among topics, to which Prof. Isaac Gilead made important contributions. In fact, Prof. Gilead was among the first scholars to show that several ritual practices existed simultaneously in the same cultural context (Gilead 2002). Since 2002, 14C dates from sites relating to these various Ghassulian ritual practices have been published (Bourke et al. 2004; Carmi and Boaretto 2004; Fabian et al. 2015: 23; Levy and Burton 2006; Shalem et al. 2013: 413-415), allowing a better understanding of the chronological aspects of the finds related to ritual behavior. The aim of present paper is to establish the chronological relationship between these various rituals of the Ghassulian culture, in order to understand how Ghassulian ritual behavior changed over the second half of the 5th millennium BC.

Teleilat Ghassul (Seaton 2008), En Gedi (Goren 2014; Ussishkin 2014), Peqi‘in (Shalem et al. 2013), and Nahal Qanah (Gopher and Tsuk 1996). The second category of studies offers a generalized view of the Ghassulian ritual. Typical of these studies are the works of Elliot (1977), Epstein (1978), Gilead (2002), Levy (2006b) and Ilan and Rowan (2012; Rowan and Ilan 2007; 2013). These publications present a general overview of Ghassulian ritual practices and regard the different Ghassulian ritual assemblages as contemporary, either implicitly or explicitly. Levy (2006b: 833, emphasis added) states: ‘Based on available dates from these two neighboring geographic zones, the most that can be said is that that both Gilat and Shiqmim (and most of the Chalcolithic cultures of the southern Levant) were all contemporary’. Gilead (2011) already demonstrated that the Ghassulian culture can be divided into an early phase, from 4500-4300 BC, and a late one, which ends around 4000/3900 BC. The aim of the present paper is to analyze how ritual behavior changed during these phases and what might have influenced those changes.

The Ghassulian culture is named after the site of Teleilat Ghassul, located on the northern side of the Dead Sea. It was first excavated by Mallon in 1920s (Mallon et al. 1934; Neuville 1930). Ghassulian sites are dispersed from the Upper Galilee in the north to the Negev desert in the south and from the Mediterranean in the west to the Jordan River basin and the Dead Sea basin in the east. The Ghassulian culture lasted roughly from 4500 to 3900 BC (Gilead 2011: 13-14; Rowan and Golden 2009). The Ghassulian sites and assemblages have been extensively studied since the late 1920’s, and they are relatively well known (e.g. Bourke 2001; Elliott 1978; Gilead 2011; Gonen 1992; Levy 1995; Rowan and Golden 2009). Sites of the Ghassulian culture feature rectangular architecture, made of stone and mudbrick, as well as underground rooms. Common finds include flint sickle blades and fan scrapers, pottery vessel types such as v-shaped bowls, fenestrated bowls, hole-mouth jars and churns, first and second burials and the first metallurgy of the southern Levant.

The Early Ghassulian Rituals Of the early Ghassulian phase, as defined by Gilead (2011), Gilat and Teleilat Ghassul are most extensively discussed when it comes to ritual practices, to name only the monographs devoted to their respective sanctuaries (Levy 2006a; Seaton 2008). Considering that Ghassulian absolute chronology of both Gilat and Teleilat Ghassul is well published and analyzed, below is a brief presentation of dates and their current interpretation. On Fig. 1 we can see the available dates from the Ghassulian sites of Teleilat Ghassul and Gilat (Bourke et al. 2001; Bourke and Lovell 2004; Levy and Burton 2006; Weinstein 1984), calibrated by OxCal v4.2.4 (Bronk Ramsey et al. 2013; Reimer et al. 2013). The eight upper dates come from Gilat (Levy and Burton 2006). Three of those dates (OxA-3555, OxA-3556 and Beta-131730) represent Besorian occupation of the site (Gilead 2009: 347), and three (OxA-4001, RT-860A and Beta-131729) reflect early Ghassulian culture. Two remaining dates (RT-2058 and RT-860B), which correspond to the mid and late fourth millennium BC, are considered to be outliers that might represent post-depositional disturbances (Levy and Burton

Numerous studies of the Ghassulian ritual practices have been conducted so far. Those studies can roughly be divided into two main categories. The first category deals with ritual practices of specific sites, for example Gilat (Alon and Levy 1990; Levy 2006a), 109

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Early Ghassulian ritual architecture

2006: 866). As mentioned above, Levy (2006b: 833) claims that Gilat was contemporary to all of the northern Negev sites, including sites along Nahal Beer Sheva, Nahal Besor and Nahal Grar. The archaeological assemblage of Gilat supports the radiometric data that places the early levels centuries earlier than the dates of the Beer Sheva sites (Gilead 2011: 20).

All alleged Ghassulian temples, including those at Gilat, Teleilat Ghassul, and En Gedi, belong to an early phase of Ghassulian culture. Before going into more details on the sites and their ritual assemblages, there is a terminological issue crucial for discussion concerning ritual behavior. I have mentioned above the sanctuaries of Gilat and Teleilat Ghassul, as well as the temple at En Gedi. As Gilead (2002: 106) clearly demonstrated, the terms temple, sanctuary and shrine have been used interchangeably in the scientific literature concerning the Ghassulian culture, without explicit definition. The architecture of Ancient Israel (Kempinski and Reich 1992) in its glossary (Reich and Katzenstein 1992: 321) defines temples, sanctuaries and shrines under the same entry as: ‘The dwelling of the god. A public building to house the god, in which the god’s statue was erected and his cult and rites performed.’ The term used in the present research is temple defined by Gilead (2002: 106) as ‘an enclosed sacred space, a complex with a building or buildings set apart from the ordinary. The nature of the architecture suggests it was for public use although it may have restricted sections. The artifacts in a temple suggest that domestic activities were neither the only nor the principal function of the structure’.

Remaining dates presented in Fig. 1 come from Teleilat Ghassul. Like Gilat, Teleilat Ghassul was occupied prior to the Ghassulian culture (dates SUA-732, SUA734, SUA-736, SUA-738/1, SUA-739, OZD024, OZD025, OZD026, OZF418, OZF421) (Bourke et al. 2001; Lovell 2001). The Ghassulian culture at Teleilat Ghassul starts at about 4500 BC (Gilead 2011: 14) and ends at around 3900/3800 BC (Bourke and Lovell 2004: 322). This span implies that Teleilat Ghassul covers the full range of the Ghassulian culture. However, Gilead (2011: 20) suggests that, considering a majority of the dates available for the Ghassulian sequence at the site cluster at ca. 4400-4300 BC, the main Ghassulian occupation is contemporary with Gilat and thus, belongs to the earlier phase of the culture. Also, dates coming from a collapsed wooden beam, which held the roof of Sanctuary A of Teleilat Ghassul temple (Bourke et al. 2004: 317), SUA-511a-c, date the temple to the early Ghassulian, while the later dates of the same Area E are not related to the temple (Bourke et al. 2004: 319).

Only En Gedi and Area E structure at Teleilat Ghassul should be considered as temples (cf. Gilead 2002). The existence of a temple at Gilat, suggested by in the 1970s (Alon 1977) and maintained by Levy (2006a), albeit with alteration in terminology as in more recent work he refers to it as a sanctuary, has been repeatedly contested, as architectural remains do not allow such categorization (Gilead 2002: 107-109; Gošić 2014: 138-140). Nevertheless, Gilat attracted the attention of participants in rituals from different regions. Petrographic studies (Goren 1995; Goren 2006) suggest that Gilat was a regional center of northern Negev, which attracted imports, such as obsidian from Anatolia (Yellin et al. 1996), greenstones from Feinan (Golden 2009: 81), basalt from the Golan (Rowan et al. 2006) and a variety of materials from which stone figurines were produced (Commenge et al. 2006) and its ritual significance is elaborated on further below.

Thus, it is evident that both sites were settled before the Ghassulian, which starts at around 4500 BC. The main occupation of both sites ends at around 4250 BC. Gilat is abandoned after that, and the two considerably later dates are an anomaly. Teleilat Ghassul remained settled until the end of the 5th millennium BC, but on much smaller scale. It is noteworthy that both sites were settled prior to the Ghassulian culture. Indeed, this is another feature that sets them apart from the sites of the Beer Sheva cluster. Although not radiometrically dated, it has been suggested (Gilead and Gošić 2014: 234-235; Gošić 2014) that En Gedi is also of early Ghassulian phase. This suggestion is based on the frequency of cornets, which constitute 40% of the pottery assemblage at En Gedi (Gilead 2002: 206; Goren 1995: Fig. 4), and on a complete absence of any remains of metallurgical artifacts or activities. Those artifacts and activities are the main characteristics of the premetallic, early phase of the Chalcolithic1 culture, according to Golden (2009: 46-67). Such dating of the En Gedi temple is essential for the ongoing debate regarding its relationship with the Nahal Mishmar hoard discussed below.

En Gedi En Gedi, located in the Judean Desert overlooking the Dead Sea, is unique in Ghassulian archaeology, as the only site devoted solely to ritual. It is widely considered to be a temple (e.g. Amiran 1989; Gilead 2002; Ilan and Rowan 2012; Rowan and Ilan 2007; Rowan and Ilan 2013; Ussishkin 1971; Ussishkin 2014). It is isolated from any settlement, and it consists of two broad buildings and a courtyard featuring a circular installation. The temple is enclosed by a wall that features two gates, and it is a

Golden (2009, 47) refers to the early phase of the Ghassulian culture as the Ghassulian Chalcolithic, as opposed to later, metallic phase which he refers to as Beer Sheva culture. 1 

110

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant

Figure 1. 14C dates from Gilat and Teleilat Ghassul (Bourke et al. 2001; Bourke et al. 2004; Levy and Burton 2006; Weinstein 1984). Calibrated by OxCal v4.2.4 (Bronk Ramsey et al. 2013, Reimer et al. 2013) and shown here with 2σ probability range.

111

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead good example of temple architecture. Numerous refuse pits have been found in the temple courtyard and in its vicinity (Ussishkin 1980: 4).

Dead Sea, presents a stark contrast to En Gedi, as it is a vast settlement spread across twelve small tells. It was extensively excavated throughout the twentieth century (Hennessy 1982; Koeppel 1938; Koeppel et al. 1940; Mallon et al. 1934; North 1961; Seaton 2008: 17).

Cornets dominate the En Gedi assemblage, with fenestrated bowls also present. Pottery typical of domestic contexts such as storage jars and cooking pots, was found in very small quantities. Notably, flint was virtually absent at the site (Ussishkin 1980: 18028). Other finds include a figurine of a bull laden with two churn-like vessels and two figurines probably representing a serpent, an alabaster vessel paralleled only in pre-dynastic Egypt, and two pendants (Ussishkin 1980: 21-25). Several horns of female and male ibexes were found in one of the building and in refuse pits in the courtyard, alongside many nonmarine shells and small bone (Ussishkin 1980: 2829).

The one feature that allows comparison between En Gedi and Teleilat Ghassul is the temple found in the Area E (Fig. 2). Seaton (2008: 19-20) defines two phases of the temple, with fairly similar outline, with the exception of the paved pathway between the circular structure in the middle of the courtyard and one of the buildings. As in En Gedi, the temple features two buildings and a circular installation in the courtyard. In addition, it was encompassed by a wall, which was partially excavated. Finds at the temple also compare well to those of En Gedi, due especially to the high frequency of cornets. 49% of all cornets from the site come from the area E temple, as well as 70% of the rare forms (goblets, fenestrated stands, and anthropomorphic figurines (Seaton 2008: 53). It remains unclear why the adjacent area is referred to as industrial. This area is comprised of an open space with numerous refuse pits, which are also found in the temple courtyard and which also feature a large number of cornets (Seaton 2008: 50, 57). Considering that such pits were present at En Gedi as well, it is probable that they were related to ritual activity performed at the temple.

There is no consensus regarding the focus of ritual practices performed at En Gedi. Since I regard En Gedi as early, and metallurgy as a late Ghassulian feature, there is no possibility that the objects from the Nahal Mishmar hoard were part of the ritual (Gilead and Gošić 2014: 236), as has been suggested before (Goren 2008; Ussishkin 1971; Ussishkin 2014). Mazar (2000) suggested that the focus of the ritual was a sacred tree located in the circular installation found in the courtyard, a claim which has been recently endorsed by Chanteau (2014). According to him (Chanteau 2014), a naturally occurring tree, located in the center of the temple at equal distance from the main architectural features, held a central place in the world of the worshipers, and the whole temple was constructed around it. He claims (Chanteau 2014: 513) that the more humid climate of the period ‘could easily permit the growth of the trees in this area’. Yet he continues, ‘Subsequently, the isolation of this tree could have incited the itinerant population to sacralize it’. However, he does not specify either what the ‘subsequently’ means in this context or how it relates to the chronology of the temple’s construction and utilization. As for the increased humidity during the period, it seems that the difference between Chalcolithic times and the present period was not significant (Bar-Matthews and Ayalon 1997: 166; BarMatthews et al. 2003: 3196). This means that trees could not easily grow at the temple, especially because it is located on a cliff above the springs. Consequently, no spring water could reach the area naturally. It is likely, however, that the circular construction was of importance and, as is demonstrated below, it has a parallel in the Area E temple at Teleilat Ghassul.

Prior to the discovery of the area E temple, discussion on ritual activity conducted at Teleilat Ghassul was related to locations where famous murals were discovered. Koeppel, Mallon and Neuville discovered the famous mural ‘the Star’ (Mallon et al. 1934: Frontpiece), and also ‘the Notables’ and ‘the Bird’, while later on North uncovered ‘the Geometric Painting’ and ‘the Tiger’. The last two representations, which are also the only ones to survive today, were discovered by Hennessy in 1968 and 1977, the first is a portion of a larger badly preserved mural, while the second is ‘the Procession’ (Cameron 1981: 3). Additional work has been done recently on the murals excavated by Hennessy during the Sydney University campaign, especially on the ‘Procession’ mural (Drabsch 2015b; Drabsch and Bourke 2014). The first reconstruction of the image, originally discovered fragmented, featured three masked standing figures, while the new one features an additional five figures (Drabsch and Bourke 2014: 1085). However, the new reconstruction seems to be somewhat unconvincing, as the eight figures it depicts are all quite fragmentary (Drabsch and Bourke 2014: Fig. 5). The new reconstruction is very different from the original reconstruction that shows three almost complete figures.

Teleilat Ghassul The eponymous site of the Ghassulian culture, Teleilat Ghassul, located not far from the northern edge of the 112

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant

Figure 2. Plan of the temple in Area E at Teleilat Ghassul, ‘Classic Courtyard’ phase. Courtesy of Peta Seaton (reproduced from Seaton 2008, Plate 8).

It has been already established by the original excavators (Mallon et al. 1934: 137) that ‘the Star’ was found in a domestic context. Mallon et al. (1934: 140) also suggested that ‘the Star’ represented a single composition, because all the elements were painted on the same layer of plaster. However, the fact that they were painted on a same layer of plaster implies that walls were not re-plastered between the painting of different elements. It does not necessarily indicate that the elements were painted together, in one event, as a single composition (Gilead 2002: 117). Later study has confirmed (Drabsch and Bourke 2014: 1094) that murals discovered by Hennessy have been repainted on up to twenty successive layers of plaster. This fact further suggests that the production of the murals was a repetitive ritual activity. According to Drabsch (2015b: 55), the houses in which the murals were located were of domestic character. However, they have, over a long time of continual use that involved ritual activity related to murals, gained special significance. In her recent volume on Ghassul paintings, Drabsch (2015a)

draws a strong connection between several of the wall paintings and a temple. For example, she argues that the purpose of the star in ‘the Star’ painting was to designate that the structure shown in the bottom left corner of the same composition is a temple devoted to a star, or to a deity symbolized by a star (Drabsch 2015a: 32). Another depiction of a temple in Ghassul paintings can be seen in ‘the Procession’ where, according to Drabsch (2015a: 161), figures of ritual participants, possibly initiates, are shown next to a temple. Finally, it appears that ritual practices at Teleilat Ghassul had a dual character. This is not an unusual norm. It is common for a centralized cult, which tends to be exhibited in temples, to be only one side of community’s religious behavior. Thus, for example, ceremonies against evil spirits or other forms of magic are frequently practiced outside of temples, in houses of healers, on burial sites of important individuals, and so on (Gilead 2002: 115-116). It is suggested (Gilead 113

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead 2002: 120) that Ghassul murals should be understood in this context, as evidence of performance of local healers/magicians. There is also a possibility that rituals conducted in the temple and those depicted in the wall paintings are closely related. Thus Drabsch (2015a: 161-162) suggests that wall paintings show initiations of priests and other rituals happening in the temple. Consequently, although the structures are habitations, it is possible that they were houses belonging to persons of priesthood lineage (Drabsch 2015a: 172).

number, these burials exceed those found at other Ghassulian sites. For comparison, only 26 burials of mostly children and infants, were found during the first two excavation campaigns at Teleilat Ghassul (Koeppel et al. 1940; Mallon et al. 1934: 50). However, several additional burials were reported in later campaigns (Bourke 2001: 148-149; Elliott 1977: 22; Hennessy 1969: 7; Seaton 2008: 34). While there is no detailed study of Teleilat Ghassul burials, the descriptions available state they were found within houses. Another difference between Teleilat Ghassul and Gilat burials is that while most of burials from Teleilat Ghassul are infants and children, over half of all the Gilat burials were over 18 years old at the time of death (Smith et al. 2006: Fig. 8.3).

Seaton (2008: 127) sees evidence for the existence of different practices at the site simultaneously as uncomplimentary. She interprets them as indicating different religions and beliefs practiced in the site, by different groups. However, there is no indication that a multi-cultural or multi-faith population lived at Teleilat Ghassul. After all, as explained above, different practices commonly flourish side by side in the same cultural context. Also, architectural motifs featured in the murals might represent the temple itself (Drabsch and Bourke 2014: 1085), meaning that temple and murals were various segments of coherent ritual practices.

At Gilat, primary burials were found in shallow pits or fills related to open areas, without particularly designated cemetery area (Smith et al. 2006: 337338). A multiple P burial of nine (Levy et al. 2006: 105) excavated in proximity to the alleged sanctuary is the only one assumed to have had a free-standing mortuary monument (Smith et al. 2006: 337-338). They also suggest that the ritual revolved around the burial of important individual(s), possibly a spiritual or religious figure (Smith et al. 2006: 339).

Mortuary ritual of the early Ghassulian culture Rappaport (1999: 24) defines ritual as ‘the performance of more or less invariant sequences of formal acts and utterances not entirely encoded by the performers’. Considering that ritual defined in this manner is not restricted to religious action, he (Rappaport 1999: 27) adds that it also generates ‘the concept of the sacred and the sanctification of conventional order … theories of the occult, the evocation of numinous experience, the awareness of the divine, the grasp of the holy, and the construction of orders of meaning transcending the semantic’. Such a definition of ritual behavior, which is broad and not restricted by formal setting of behavior in question, makes it evident that temples are by no means the only ground for ritual performance, within Ghassulian culture and otherwise. This is precisely why finds that appear to be, and certainly were, of ritual importance, such as the Gilat woman or the murals from Teleilat Ghassul described above, do not necessarily mean that the location of their discovery, even if found in situ, was a temple. I describe bellow the most prominent finds from Gilat that are considered to be of ritual significance, with emphasis placed on the burial customs, as they change considerably in the later phase of the culture. The change makes them crucial for understanding the transition between the early and late Ghassulian phases.

Other evidence of early Ghassulian ritual behavior

Gilat

Among other ritually significant early Ghassulian finds are figurines that are essential for understanding the plurality of Ghassulian ritual behavior. Figurines found at Gilat alone already reflect this diversity.

It is common to attribute certain pottery types, such as cornets, miniature churns, fenestrated bowls, and v-shaped bowls to ceremonial or ritual actions (Commenge 2006; Helms 1987: 52; Seaton 2008: 43; Ussishkin 1980). Cornets are regarded as ceremonial vessels due to their high frequency in the En Gedi temple (Ussishkin 1980: 38) and Gilat (Amiran 1976: 119) assemblage. Churns, v-shaped and fenestrated bowls found in caves (Gal et al. 2011; Gal et al. 1997; Gopher and Tsuk 1996; Porath 2006; Yannai and Porath 2006) are associated with second burials as also at En Gedi (Ussishkin 1980). Additionally, v-shaped bowls are associated with two burials at Gilat (Smith et al. 2006: 337). Although it seems reasonable that these items were used in rituals, they were not restricted to ritual. In fact, they were also used in domestic contexts (Eldar and Baumgarten 1985: 138; Gilead 1989: 383; Perrot 1955: 82; Rosen and Eldar 1993: 20). Thus, the appearance of any of those vessels should not be taken as a necessary indication of a ritual context. In addition, the practice of using the same artifacts in ritual and domestic contexts is known from ethnographic examples as well (Walker and Lucero 2000: 133).

The ritual significance of Gilat is to be found in the 91 burials found at the site (Smith et al. 2006: 327). In 114

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant They include an exceptionally large number of violinshaped figurines (53) of either various types of stone or bone, and zoomorphic and anthropomorphic types in pottery such as the Gilat Woman, the Gilat Ram and 12 additional zoomorphic figurines (Commenge et al. 2006). Violin-shaped figurines, some of which are breasted, have been interpreted as female representations (Elliott 1977: 6; Ilan and Rowan 2012: 106; Levy and Alon 1987b: Fig. 6.12.4; Rowan and Ilan 2007: 253). Others are more cautious and less conclusive in their interpretations of the nature of the figurines (Commenge et al. 2006: 791-792; Joffe et al. 2001: 16). The latter scholars suggest that the presence of breasts/nipples could suggest that only those with such protuberances were female while the remaining non-adorned examples were intended to portray males. Features of the ‘Gilat woman’ pottery receptacle, which is represented holding a churn on her head, a cornet under her arm and sitting on a cylindrical object resembling a copper ‘crown’, are unequivocally female. Indeed, they feature tumid pubic area, emphasized incisions and painted red dots, contrasted by its relatively reduced presentation of breasts (Commenge et al. 2006: 742-746). Another ceramic vessel, ‘The Gilat ram’, is also clearly sexed. It is a representation of a horned ram with emphasized male genitalia and three cornets on its back (Amiran 1976). The shape of its body is similar to a churn. It was probably produced in the same manner, its horned head and tail in place of churn handles, as in miniature churns, which occasionally are horned (Commenge 2006: 492, Plate 10.31.3, 4, 10, 12).

of 238 gazelle feet have been recovered from a single building on the site, Building 1. Articulated human burials were excavated from the earlier phase of the building (Price et al. 2016: 19) – an already elaborated feature of the early Ghassulian ritual behavior. While it appears reasonable to presume gazelle feet to be remains of a ritual practice (Price et al. 2016: 19-20), the exact practice is difficult to reconstruct, considering it is a unique find. Also, most zoomorphic figurines are of domestic animals, while gazelle appears more prominently in the iconographic repertoire only in the later phase of the culture, as the examples discussed below demonstrate. The above described material brings us to several points regarding early Ghassulian behavior. First, it is obvious that a variety of ritual practices was present in early Ghassulian. Gilat was a regional pilgrimage site, focused on mortuary ritual. Materials such as obsidian from Anatolia (Yellin et al. 1996), greenstones from Feinan (Golden 2009: 81), basalt from the Golan (Rowan et al. 2006) and the variety of materials in which stone figurines were produced (Commenge et al. 2006), indicate that Gilat was part of an interregional network. There is no indication that individuals, who were buried there, belonged to different social strata. Moreover, even if they did, this is not reflected in the burial customs. It is suggested that burial custom is an indirect reflection of a human society, which means that, instead of seeing afterlife as an apparent reflection of a social status (where higher social status means a ‘rich’ burial), it is a reflection of its attitude towards death. As such, the form of burial reflects a complex relationship with the structure of a society that cannot be defined by a simple rule (Hodder and Hutson 2003: 3). There is no evidence of the special selection for burial or the ceremonial treatment of the deceased. Consequently, it appears that that the Ghassulian society was not ranked. At the very least, obvious material demonstration of rank was not required in a mortuary context. However, the mass burial possibly associated with a standing monument might have been the focus of mortuary pilgrimage (Smith et al. 2006: 339). It can be said that the lack of ranking among the graves in its vicinity and their haphazard organization implies equal access to the mortuary site and burial.

Zoomorphic pottery vessels are also found at En Gedi and Teleilat Ghassul. Two animal shaped vessels have been found associated with Sanctuary B of the Area E temple. One is a spouted quadruped, probably a cow. The other object, also a quadruped, is possibly a dog that may have been attached to the lid of a vessel (Seaton 2008: 78-79, Plate 96a-b). Other zoomorphic figurines are found in varying contexts around Teleilat Ghassul, including domestic structures and temple area. Violin-shaped figurines are also present. At En Gedi, a figurine of a bull laden with two churn-like vessels, comparable to the Ram from Gilat and decorated in red paint, was found in the ashes near the altar. The body of the bull is solid, but the churn above is hollow, suggesting it was intended to be used as a vessel. Two additional figurines are probably representations of serpents, and they are also decorated with red paint. Other zoomorphic figurines recovered are too fragmentary to be identified (Ussishkin 1980: 20-21). Ritual significance of animals is also evident in faunal assemblages from settlement at Marj Rabba in lower Galilee where a unique find of burned gazelle feet have been found in a Phase III of the settlement (Price et al. 2016: 10) which corresponds to what is considered here as the early phase of the Ghassulian culture. A total

Ghassulian temples existed at En Gedi and Teleilat Ghassul. While it is evident that En Gedi was not permanently settled and rituals performed there were periodical, the dynamics of ritual activity at Teleilat Ghassul are unknown. However, considering the resemblance of the structure to En Gedi and the assemblage of ceremonial artifacts, it is likely that the rituals were of the same kind in terms of organization and symbolism. This notion is strengthened by the relative proximity of the site. It is likely that 115

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead inhabitants of Teleilat Ghassul practiced rituals at En Gedi as well as at other places. Finally, ritual practices at Teleilat Ghassul were not limited to the temple area and interpretation of murals as the work of local healers and or magicians, as is described above.

possible that the custom of second burials originated there. However, more dates from Peqi‘in, as well as other second burials caves, are necessary to sustain such a claim. Since the writing of this paper, dates from two more second burial caves, Qina and Ashelim, have been published, also belonging to the late phase of the Ghassulian culture (Langgut et al. 2016: 981) It is thus now safe to say that the second burial custom is characteristic of the late Ghassulian phase.

It is safe to say that ritual behavior during early Ghassulian phase varied between the sites. However, on the basis of similar iconographic features, we may conclude that it reflects the same wider belief system. The early phase of the Ghassulian culture was a time of ritual variety, including temples, domestic ritual possibly involving a healer or a shaman, intramural burials, and a pilgrimage site focused on mortuary ritual. Yet, except for En Gedi, all other sites of the early phase are primarily settlements, with domestic architecture and accompanying pottery and flint. There are no cemeteries in the earlier phase, and even Gilat is primarily a settlement.

Second burial customs of the late Ghassulian culture Inhumations are found at late Ghassulian settlements, but in much smaller numbers than in the earlier phase (Eldar and Baumgarten 1985: 138; Gilead 1987: 115; Perrot 1955: 173-174, 189; Perrot 1958: 122123; Scheftelowitz and Oren 2004: 28-33). Second burial rites are a major mortuary feature of the late Ghassulian. They are found either in natural karstic caves, such as Peqi‘in, Nahal Qanah and Ben Shemen, in hilly or mountainous regions, or in man-made caves, such as Bnei Berak, Azor, Palmahim Quarry and Hadera (van den Brink 2005: 165-166). Second burials in caves are the most numerous Ghassulian burial types, as more than fifty second burial sites are known so far. In some of them the numbers of burials have reached hundreds, as in case of Peqi‘in where an estimated number of 600-1000 individuals were buried (Winter-Livneh et al. 2012: 424). Ghassulian second burials are most commonly placed in ceramic ossuaries and large vessels (e.g. Fabian 2012; Gal et al. 1997; Gopher and Tsuk 1996; Gophna and Lifshitz 1980; Goren and Fabian 2002; Perrot and Ladiray 1980), including churns, jars, and large v-shaped bowls (Fabian et al. 2015). While stone ossuaries and so-called tubs are slightly less common, they are present (e.g. Goren and Fabian 2002; Gorzalczany et al. 2012). There are several second burial sites which are not located in caves: Kissufim (Goren and Fabian 2002), Palmahim (North) (Gorzalczany 2006; Gorzalczany et al. 2012), and Shiqmim cemeteries (Levy and Alon 1982; Levy and Alon 1987a; Levy et al. 1993b).

The Late Ghassulian Ritual Behavior Much of this changes with the late Ghassulian phase and the rise of the settlements in the Nahal Beer Sheva. Fig. 3 shows dates from Horvat Beter, Bir esSafadi, Shiqmim, and Giva’t ha-Oranim (Burton and Levy 2011: 179; Carmi and Boaretto 2004; Gilead 1994: 2). Few dates from Shiqmim (RT-6490, RT0649B, OxA2523 and OxA-2524) suggest that occupation started earlier at Shiqmim. However, the bulk of its data places it clearly in the later phase. Based on these dates, the late Ghassulian phase is considered to begin around 4300 and to terminate around 4000/3900 BC, as has been already confirmed (Gilead 2011: 14). Another set of late Ghassulian dates comes from burial caves, which represent a new mortuary custom of the phase. These data are discussed below. Horvat Qarqar South (Fabian et al. 2015: 23), The Nahal Qanah (Carmi 1996), Shoham (North) (Carmi and Segal 2005) and Peqi‘in (Segal et al. 1998; Shalem et al. 2013: 413-415) dates shown in Fig. 4 were the only dates from second burial caves available from a second burial caves at the time present paper was written. The dates from Horvat Qarqar South fall neatly into the later phase of the Ghassulian culture, as do the two dates from Shoham (North). The first date from Nahal Qanah (RT-861E) covers the complete span of the Ghassulian. However, the other two (RT-861C and RT-1545) are of the late Ghassulian period. The range of the date RT-861A is too wide to be considered relevant. Concerning the remaining dates coming from Peqi‘in cave in Upper Galilee, it is possible, based on the earlier dates (RT2377 and BA090183), that the second burial in Peqi‘in started towards the end of the early Ghassulian period, although most burials are late Ghassulian. Considering the relative isolation of the Peqi‘in cave in relation to other burial caves of the Ghassulian culture, it is

Perrot and Ladiray (1980: 28) distinguish between three groups of receptacles used for second burials: rectangular and basin-shaped ceramic and stone ossuaries, ossuaries resembling a shape of a house and ossuaries in shape of rounded craters. However, numerous ossuaries have been excavated since their classification, with exceptionally great variety of shapes coming from Peqi‘in (Shalem et al. 2013). These include anthropomorphic ossuaries, both male and female, as well as zoomorphic ones (Fig. 5). It is somewhat difficult to distinguish between anthropomorphic and zoomorphic ossuaries from Peqi‘in, and the zoomorphic ossuaries, like the one on the Fig. 5.C, are categorized by a knob on the back, interpreted as a tail (Shalem et al. 2013: 76). 116

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant

Figure 3. 14C dates from Horvat Beter, Bir es-Safadi, Shiqmim and Giv’at ha-Oranim (Burton and Levy 2011, 179; Gilead 1994, 2) (Carmi and Boaretto 2004). Calibrated by OxCal v4.2.4 (Bronk Ramsey et al. 2013, Reimer et al. 2013) and shown here with 2σ probability range.

117

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 4. 14C dates from Horvat Qarqar South, Nahal Qanah, Shoham (North) and Peqi‘in (Carmi 1998, Carmi and Segal 2005; Segal et al. 1998; Shalem, Gal and Smithline 2013, 413415). Calibrated by OxCal v4.2.4 (Bronk Ramsey et al. 2013, Reimer et al. 2013) and shown here with 2σ probability range.

Zoomorphic motifs, mostly in the form of ibexes or ibex horns are known from other second burial sites, including ibex representation on a crater from Horvat Qarqar South (Fabian et al. 2015: Fig. 10.18).

A peculiar vessel comes also from Palmahim in the shape of a bird (Gophna and Lifshitz 1980: 6). Fenestrated bowls produced in stone were also found in Nahal Qanah (Gopher and Tsuk 1996: 110-112) and Peqi‘in (Gal et al. 2011: 202). However, they are in general less frequent. Bone and shell pendants and beads are frequent (e.g. Bar-Yosef Mayer 2002; Gal et al. 2011: 202; Gopher and Tsuk 1996: 128-129; van den Brink et al. 2004: 145). Beads in general, and steatite beads in particular (Bar-Yosef Mayer et al. 2004) are evidence of exchange over considerable distances, as are stone mace heads (Gal et al. 2011: 202; Gopher and Tsuk 1996: 109, 113) and palettes (Fabian and Goren 2002; van den Brink et al. 2004). Copper artifacts have also been found in many of the caves, including Peqi‘in (Segal and Goren 2013), Nahal Qanah (Shalev 1996), Cave V/49 (Ketef Jericho) (Segal 2002) and Nahal Zeelim (Key 1980: 239).

Pottery vessels found in second burial sites, apart from ossuaries that are already discussed above, include v-shaped bowls, hole-mouth and squatted jars, jars with pronounced neck, churns (large and miniature), v-shaped bowls and fenestrated bowls and stands (Gal et al. 2011: 201-202; Gopher and Tsuk 1996: 91-108; Gophna and Lifshitz 1980: 5; Goren 2002; van den Brink et al. 2004: 143-144; Yannai and Porath 2006). Cornets are less common (e.g. Gal et al. 2011: 202; Perrot 1992; Yannai and Porath 2006). A single triple fenestrated stand (a stand branching to three smaller stands, each featuring a v-shaped bowl on top) was found at Peqi‘in (Gal et al. 2011: 203). Unusual types of vessels, cylinder stands are found at Et-Taiyiba cave, a site that also features variety of fenestrated bowls (Yannai and Porath 2006: 39).

Anthropomorphic and zoomorphic figurines are also found in second burial sites, with violin-shape figurines 118

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant

Figure 5. Selected ossuaries from Peqi‘in: A. female B. male C. zoomorphic. Courtesy of Israel Antiquates Authority and Dina Shalem, Zvi Gal and Howard Smithline.

coming only from Peqi‘in. Realistic anthropomorphic pottery figurines are also encountered at the site, and they were decorative elements of ceramic vessels Gal et al. (2011: 202). A bird figurine was found at Kissufim road (Bar-Yosef Mayer 2002). Although zoomorphic figurines are not otherwise encountered, images of animals are frequent in late Ghassulian iconography that is discussed below.

Although there is a degree in variety among the above-described sites, a general picture of off-site second burials attests to a shared belief system during late Ghassulian phase, at least in regards to burial customs. Second burials in caves imply considerable investment in mortuary rites. Individuals from throughout the southern Levant, from Beer Sheva to the Golan, participated in the mortuary practices at 119

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Peqi‘in (Gal et al. 2011: 204). Consequently, funerary assemblages, including burial remains, needed to be transported. In addition, in some cases, caves had to be dug out.

(not necessary members of particular nomadic tribes, just people traveling between regions), who died away from their settlements. The Nahal Mishmar hoard and Ghassulian metalworking

The only Ghassulian settlement located in close proximity to a cemetery is Shiqmim cemetery, with its three adjacent cemeteries (Levy and Alon 1982; Levy and Alon 1987a; Levy et al. 1994; Levy et al. 1993a; Levy et al. 1991). At both Cemetery 1 (Mezad Aluf) (Levy and Alon 1982: 42-46) and Cemetery 3 (Levy and Alon 1987a), grave circles were used for second burials, as disarticulated skeletons were found within them. Stone cists used, most likely, for excarnation were also found. They are the only Ghassulian excarnation constructions discovered so far. Each one of the cists contained at least one v-shaped bowl (Levy and Alon 1987a: 335, Table 13.2). Finds other than human remains, such as bone and shell pendants, copper beads, and v-shaped bowls, were unevenly distributed between the grave circles (Levy and Alon 1982: 53-54). Numerous cists have also been excavated recently at Palmahim (North) (Gorzalcany 2018). The scarcity of the poorly preserved osteological remains found in the cemetery (Yossi 2018) opens a possibility the site might have been used for excarnation as well.

The Nahal Mishmar Cave of the Treasure is a unique site in the Ghassulian culture, making it difficult to categorize. It is located in the Judean desert, high up a cliff of the Nahal Mishmar streambed. During the 1961 season, a hoard including numerous copper objects was found in Cave 1 (Bar-Adon 1980). However, its exact location and precise dating remain obscure (Gilead and Gošić 2014: 227). Thirteen 14C dates (Bar-Adon 1980: 198; Davidovich 2008: Table 3) are available from the Cave of the treasure (Fig. 6). The considerable time span of over 1500 years covered by these dates and, especially, different dates obtained from the same sample of the mat in which the objects from the hoard are wrapped (dates AA37102b, AA-37206c and AA-37206a) make radiometric dating of the hoard (Gilead and Gošić 2014: 232-233) difficult. However, the copper objects of the hoard clearly place it in the late, metalworking phase of the Ghassulian culture. The hoard, with its 423 copper artifacts, constitutes most of the currently known Ghassulian copper artifacts. Consequently, most studies (Bar-Adon 1980; Beck 1989; Elliott 1977; Epstein 1978; Gates 1992; Tadmor 1989; Tadmor et al. 1995), both of technology and style, have been conducted on the objects from the hoard. Since the original publication of the Nahal Mishmar hoard, the copper artifacts of the Ghassulian culture have been divided into two groups: utilitarian and prestigious (Potaszkin and Bar-Avi 1980: 235). However, the usefulness of the so-called utilitarian artifacts has been questioned repeatedly (Namdar et al. 2004: 81; Tadmor et al. 1995: 97). In addition, copper artifacts of both groups are found not only in the Nahal Mishmar hoard. They are all found also in other archaeological contexts: production sites (Eldar and Baumgarten 1985; Gilead et al. 1991; Perrot 1955; Shalev and Northover 1987), burial caves (Gal et al. 1997; Gopher and Tsuk 1996; Gophna and Lifshitz 1980; Perrot and Ladiray 1980; Shalem et al. 2013). It has been shown (Gošić and Gilead 2015a) more recently, that all of the Ghassulian copper artifacts are related to ritual. In addition, the metallurgical process, from the smelting of the ore to the casting of objects, was a ritual performance. The symbolism of this performance was probably based on the transformational quality of metallurgy (c.f. Gošić and Gilead 2015b).

Late Ghassulian burial customs pose a striking contrast to early Ghassulian and especially Gilat. Burial in the settlement sites are scarce. In adult burials, the remains were mostly disarticulated. Of all the late settlements, Giv‘at ha-Oranim has the largest number of buried individuals, 15 (Scheftelowitz and Oren 2004: 28-33). It is not clear whether the underground chambers used for burials were originally intended for this purpose. Most of the human remains are scattered throughout the site. Only three features clearly show evidence of an intentional burial, and all three are primary (Scheftelowitz and Oren 2004: Table 2.1). Evidence of second burial is much more elaborate, both in relation to the settlement sites and, especially, when it comes to second burial in caves. However, one important thing that late Ghassulian second burials have in common with the inhumations at Gilat (Smith et al. 2006: 340-348) is that neither of the burials reflect evidence for any social hierarchy Existence of primary burials in caves (Haas and Nathan 1973; Khalaily 2002: 129; Schick 1998) in the Judean desert adds to the variety of burial practice – it is not related to second burial and is quite incomparable to cave sites such as Peqi‘in, Nahal Qanah, Azor, and other previously discussed second burial caves. Considering the small number of individuals dispersed between caves in the Judean Desert, it appears that those burials were ad hoc practices and less organized than second burials in caves. For example, it is plausible that those burials belonged to itinerant individuals

Many of the Ghassulian copper artifacts are elaborately decorated with various motifs, including anthropomorphic (e.g. Fig. 7.A), zoomorphic (e.g. Fig. 120

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant

Figure 6. 14C dates from the Cave of the Treasure at Nahal Mishmar (Bar-Adon 1980, 199; Davidovich 2008, Table 3). Calibrated by OxCal v4.2.4 (Bronk Ramsey et al. 2013, Reimer et al. 2013) and shown here with 2σ probability range.

7.B-C), architectural (e.g. Fig. 7.B.), motifs of tools and weapons (e.g. Fig. 7.C-D), as well as motifs that are considered to be either abstract or floral (e.g. Fig. 7.C) (Gošić and Gilead 2015a: 166-168). These motifs are all present on other Ghassulian ritual artifacts, most notably on ossuaries. Some the artifacts made of copper had been made of stone prior to the introduction of metallurgy. Examples of these are mace heads, chisels and axes. Other kinds of artifacts, such as standards, scepters, horns and ‘crowns’ were made only from metal. I have recently suggested (Gošić 2015: 731) that the former group consists of the skeuomorphic objects, the morphology of which and, at least in case of mace heads, ritual significance, were widely recognized in either early or late Ghassulian contexts. Their purpose was to introduce metallurgy to the members of the Ghassulian population that did not participate in it, both in the Beer Sheva settlements and in other areas. The second group

consists of more innovative objects, such as standards, crowns and scepters, decorated in symbols already familiar to the Ghassulians. However, these objects were different and original enough to demonstrate all of the possibilities of the technology and the superior control metalworkers had over the material world. Other ritual-related assemblages at late Ghassulian settlement sites Although it is beyond scope of this paper to list all ritual artifacts of the Ghassulian culture, there are those that should be mentioned because they testify to the continuity, and the discontinuity between the two phases of the culture. Ivory workshop and related artifacts, both male and female figurines, are found at Bir es-Safadi (Perrot 1964). From the point of view of continuity, the particularly interesting object is a bone figurine that was also excavated at Shiqmim 121

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 7. Selected Ghassulian copper artifacts: A. anthropomorphic standard (redrawn after Bar-Adon 1980: 49, no. 21), B. ‘crown’ featuring architectural and zoomorphic motifs (redrawn after Bar-Adon 1980: 28, no. 7), C. standard featuring ibexes and motifs of tools (redrawn after Bar-Adon 1980: 100, no. 153), D. skeuomorphic axe (redrawn after Bar-Adon 1980: 112, no. 163).

(Levy and Golden 1996). The body presentation is reminiscent of a violin-shape figurine. However, its head and round incisions make it similar to bone figurines from Bir es-Safadi and other late sites. The figurine is considered to be of female gender, based on its similarity to violin-shape figurines. A small clay

figurine with pronounced breasts, resembling a violinshape figurine, was also found at Shiqmim (Levy and Alon 1987b: Fig. 6.12.4). Other small finds at the sites include chalk and hematite mace heads, beads and pendants (Levy and Alon 1987b: Figures 6.11.4, 6.12.13; Perrot 1955: 78, 172). 122

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant A unique feature of the late phase of the culture is the painted pebbles discovered at Abu Matar and Shiqmim, organized in different shapes on the floors of the room (Perrot 1955: 167-170) (Levy et al. 1994: : 89). Although the exact purpose of these pebbles is obscure, it appears that they were found in domestic contexts. Probably, they can, like wall paintings, be connected to activity of local magician. While Levy et al. (1994: 89) claim that the building in which such pebbles at Shiqmim are found is of public importance, a general plan of the Shiqmim village (Levy and Alon 1987b: Fig. 6.2) gives an impression of rather uniform architecture throughout the site. It is doubtful that any particular structure stood out as public. It seems that it is interpreted as public more on the basis of finds considered to be of ritual importance, such as bowls smashed in situ and semi-circular stone installation. However, it is clearly visible on the general plan of the site that stone installations in and outside of houses are not unusual features. The public character of the structure is also suggested by the large storage jars (Levy and Golden 1996: 152). Moreover, ceramic vessels of different shapes are uniformly distributed throughout the site (Levy and Menachem 1987: 325). Changes in ritual Ghassulian Culture

practices

throughout

burial on the other, they suggest (Ilan and Rowan 2012) that treatment of the dead was essential to fertility and regeneration. It is noticeable that Nahal Beer Sheva sites tend to be on the margin of discussions with respect to Ghassulian ritual life. Nevertheless, most of the settlement activity during later phase of the Ghassulian culture occurred there, and copper artifacts, that reflect the Ghassulian world of symbols so elaborately were produced there. It seems reasonable that second burial was practiced primarily in this phase of the culture. However, it is difficult to imagine that ritual life revolved solely around funerary rights and was practiced mostly off-site and consisted of shamanistic rituals related to painted pebbles. It appears that something is missing. I would like to propose here another manifestation of ritual: metalworking. However, in order to approach this subject in more detail, it is necessary to understand first the symbolism behind ritualized metallurgy and how the relevant symbols are practically represented. Drabsch (2015a: 19) mentions the advent of metallurgy, obviously a feature of the Beer Sheva sites, as one of the important means of gaining social power in the Ghassulian society. However, she mentions it in same sentence with technology of wall paining. Thus, she implies that these are roughly contemporary and also that there is a comparable level of innovation involved. I do not mean to undermine the significance of the Ghassulian wall painting. However, making such a comparison seems unjustified, as metalworking involves material transformation of ore to metal. Moreover, the people to whom we refer to when we say ‘the Ghassulian culture’ were the first people to witness such a transformation in the Levant.

the

After presenting the overview of both early and late Ghassulian ritual behavior, it is time to consider the greater picture of the Ghassulian ritual. Research on the subject up until now can be roughly divided into three kinds of studies. The first group of studies is devoted to identifying deities (Elliott 1977) and, in some cases, connecting Ghassulian religion to the later religions and deities of the Near East (Epstein 1978; Merhav 1993; Mirocshedji 1993; Shalem 2008). The second group of studies analyzed the ritual performance in terms of social organization and complexity (Levy 1995; Levy 2006a; Levy and Golden 1996). The third group of studies examined domestic ritual multiple ritual behaviors as coexisting in the same social context (Gilead 2002; Ilan and Rowan 2012; Joffe et al. 2001; Rowan and Ilan 2007).

When the Ghassulian ritual relics are studied from the chronological perspective, i.e., not assuming that they were all contemporary, it becomes obvious that important ritual changes occurred throughout the second half the fifth millennium BCE. When Gilat, En Gedi and Teleilat Ghassul were either abandoned or diminished in size and importance, it appears that ritual practices of the late Ghassulian clearly changed. They seem to be less varied, as the difference of ritual assemblages between different sites is less prominent. This observation excludes, of course, the hoard of Nahal Mishmar, which is a unique find. Indeed, it might testify to a specific ritual practice, namely, the burial of ritual objects (Gilead and Gošić 2014: 235).

Rowan and Ilan (2007) also suggested that Ghassulian ritual behavior includes simultaneous existence of different practices. Thus, En Gedi is once again seen as a temple and Gilat is seen as a mortuary pilgrimage site (Rowan and Ilan 2007: 251-253). The same authors (Ilan and Rowan 2012) addressed a similar question once again in a more recent article, this time including Teleilat Ghassul and second burial practice. They concluded that Ghassulian religion was primarily concerned with death and regeneration of life. Consequently, instead of putting the Gilat woman and other fertility symbols on one side and second

Copper objects exhibit the greatest variety of symbols of all the artifacts. The decorative motifs of the Ghassulian copper artifacts resemble decorations of various other early and late Ghassulian ritual artifacts. Ceramic ossuaries offer the greatest variety 123

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead of analogies. They have been found in numerous burial caves, such as Azor and Ben Shemen (Perrot and Ladiray 1980), Nahal Qanah (Gopher and Tsuk 1996), Peqi‘in (Gal et al. 1997), and Palmahim (Gophna and Lifshitz 1980). Common motifs on ossuaries include anthropomorphic (Epstein 1978: 29, Plate 6cd; Gal et al. 1997: 149, Fig. 3; Merhav 1993: 33, Fig. 4.5) and zoomorphic motifs (Merhav 1993: 33, Fig. 4.3; Milevski 2002: 138-140) as well as doorways (Epstein 1978: 30, Plate 6d; Gophna and Lifshitz 1980: 3, Fig. 3; Merhav 1993: 33, Fig. 4.3). Motifs found in copper artifacts are also found on pottery vessels other than ossuaries. Representations of ibexes has been found on a crater from Qarqar (Fabian 2012). Two unique bird-shaped vessels have been found at Palmahim (Gophna and Lifshitz 1980: 4-6). The spread out wings of the birds resemble the so called birdshaped standard (Bar-Adon 1980: 102). Two birds are also found on a pottery vessel discovered in the northern Negev (Amiran 1986) with a basket handle resembling the handles of the Nahal Mishmar copper jars. A ceramic version of the copper jar is also known from the mortuary site of Kissufim Road (Goren and Fabian 2002: Fig. 4.2). Ceramic figurines with the protruding nose have also been found (Gal et al. 1997: 153). ‘The Gilat Woman’ ceramic figurine (Commenge et al. 2006: 742-746; Joffe et al. 2001) sits on an object similar in shape to a copper crown. Ivory figurines (Perrot 1959) discovered in Bir es-Safadi also have the characteristic Ghassulian nose. While the list of iconographical analogies is too long to be presented in detail, it is obvious that Ghassulian copper artifacts exhibit motifs widely present in the repertoire of the Ghassulian ritual, both late and early. The importance of copper artifacts for understanding changes that occurred in the Ghassulian ritual practices is even greater once it becomes evident that metalworking was not only a new technology, but also a new ritual practice.

transformational aspects of these crafts are neither as striking nor as evident as turning stone/earth into metal. The remarkable transformational quality of metallurgy was a completely new phenomenon. Ghassulian metalworkers, for the first time, gained the power to create a new material, namely, metal. To practice metalworking was to perform a material transformation. It appears that this new practiced diminished the significance of temples. The exact way in which transformational quality of metalworking was conceptually connected to second burial and how this relates to the abandonment of Gilat mortuary practices remains unclear. It is possible that it was related to understanding the transformation from life to death. It is clear, however, that introduction of a new material, namely, copper, was accompanied by significant changes in the ritual life of the Ghassulians. Ghassulian ritual practitioners did not simply use new objects in old rituals. The very nature of the ritual practice changed, together with the location where it was conducted. It is possible either that metallurgy was introduced by the descendants of early Ghassulian ritual practitioners or by the first metalworkers, who became ritual specialists of the late Ghassulian communities. Finally, it is necessary to conclude by emphasizing something that the ritual practices of the early and the late Ghassulian culture have in common and that is that neither of the phases can be characterized by a uniform ritual behavior. On the contrary, various practices flourished throughout both, testifying to the complexity of the Ghassulian culture and its symbolic world. Acknowledgments I would like to thank Mayer Gruber, Haim Goldfus, Peter Fabian and Shamir Yona for inviting me to participate in this publication. Peta Seaton provided me with the plans of the Area E temple at Teleilat Ghassul and Dina Shalem, Zvi Gal, Howard Smithline and the Israel Antiquities Authority granted me the permission to use the images of ossuaries from Peqi‘in. I am also thankful to my mentor Isaac Gilead for endless discussions on Ghassulian ritual behavior, which not only contributed significantly to this paper, but were also influential to my understanding of Ghassulian culture in general.

Conclusions From the material presented above, it appears that the most significant changes in the Ghassulian ritual behavior are abandonment of temples at En Gedi and Teleilat Ghassul, abandonment of Gilat and its mortuary ritual practices, and widespread second burial practice and introduction of metalworking. Although not a form of ritual, it is noteworthy that new settlements are established in the Nahal Beer Sheva and that this is where the new metallurgical rituals were performed. In late Ghassulian times, metallurgy became a prominent symbol of creation and transformation. In order to fully understand the social significance of metal objects in the Ghassulian culture, it is important at this point to emphasize the transformational quality of metalworking. Although pottery and lime-plaster were produced earlier, the

References Alon, D. 1977. A Chalcolithic temple at Gilath. Biblical Archaeologist 40: 63-65. Alon, D. and Levy, T. E. 1990. The Gilat sanctuary its centrality and influence in the southern Levant during the late 5th-early 4th millennium B.C.E. Eretz Israel 21: 23-36. Amiran, R. 1976. Note on the Gilat vessel. ‘Atiqot 11: 119-120. 124

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant Amiran, R. 1986. A new type of Chalcolithic ritual vessel and some implications for the Nahal Mishmar hoard. Bulletin of the American Schools of Oriental Research 262: 83-87. Amiran, R. 1989. The Gilat godess and the temples of Gilat, En-Gedi and Ai. In P. D. Miroschedji (ed.) L’urbanisation de la Palestine à l’âge du Bronze ancien: 53-60. Oxford, B.A.R. Bar-Adon, P. 1980. The Cave of the Treasure. Jerusalem, Israel Exploration Society. Bar-Matthews, M. and Ayalon, A. 1997. Late Quaternary Paleoclimate in the Eastern Mediterranean Region from Stable Isotope Analysis of Speleothems at Soreq Cave, Israel. Quaternary Research 47: 155-168. Bar-Matthews, M., Ayalon, A., Gilmour, M., Matthews, A. and Hawkesworth, C. 2003. Sea–land oxygen isotopic relationships from planktonic foraminifera and speleothems in the Eastern Mediterranean region and their implication for paleorainfall during interglacial intervals. Geochimica et Cosmochimica Acta 67: 3181–3199. Bar-Yosef Mayer, D. 2002. The Shell Pendants. In Y. Goren and P. Fabian, Kissufim Road. A Chalcolithic Morturary Site: 59-52. Jerusalem, Israel Antiquities Authority. Bar-Yosef Mayer, D. E., Porat, N., Gal, Z., Shalem, D. and Smithline, H. 2004. Steatite beads at Peqi‘in: long distance trade and pyro-technology during the Chalcolithic of the Levant. Journal of Archaeological Sciences 31: 493-502. Beck, P. 1989. Notes on the Style and Iconography of the Chalcolithic Hoard from Nahal Mishmar. In A. J. Leonard and B. B. Williams (eds), Essays in Ancient Civilizations Presented to J. Helene Kantor: 3954. Chicago, Oriental Institute of the University of Chicago. Bourke, S. J. 2001. The Chalcolithic period. In B. Macdonald, R. Adams and P. Bienkowski (eds), The Archaeology of Jordan: 107-162. Sheffield, Sheffield Academic Press. Bourke, S. J., Lawson, E., Lovell, J. L., Hua, Q., Zoppi, U. and Barbetti, M. 2001. The chronology of the Ghassulian period in the Southern Levant: new 14 C determinations from Teleilat Ghassul, Jordan. Radiocarbon 43: 1217-1222. Bourke, S. J. and Lovell, J. L. 2004. Ghassul, chronology and cultural sequencing. Paléorient 30: 179182. Bourke, S. J., Zoppi, U., Meadows, J., Hua, Q. and Gibbins, S. 2004. The end of the Chalcolithic period in the south Jordan Valley: new 14C determinations from Teleilat Ghassul, Jordan. Radiocarbon 46: 315-323. Bronk Ramsey, C., Scott, M. E. and Van Der Plicht, J. 2013. Calibration for Archaeological and Environmental Terrestrial Samples in the Time

Range 26–50 ka cal BP. Radiocarbon 55: 20212027. Burton, M. M. and Levy, T. E. 2011. The end of the Chalcolithic period (4500-3600) in the northern Negev Desert, Israel. In J. L. Lovell and Y. M. Rowan (eds), Culture, Chronology and the Chalcolithic. Theory and Transition: 178-191. Oxford, Oxbow. Cameron, D. O. 1981. The Ghassulian Wall Paintings. London, Kenyon-Deane Ltd. Carmi, I. 1996. Radiocarbon dates. In A. Gopher and T. Tsuk, The Nahal Qanah Cave, Earliest Gold in the Southern Levant: 205-208.Tel Aviv, Tel Aviv University. Carmi, I. and Boaretto, I. 2004. Determination Of Age Using The 14C Method. In N. Scheftelowitz and R. Oren, Giv’at ha-Oranim. A Chalcoltihic Site. Tel Aviv, Emery and Claire Yass Publications in Archaeology of the Institute of Archaeology, Tel Aviv University. Carmi, I. and Segal, D. 2005. The Radiocarbon Dates from Cave 4. In E. C. M. Van Den Brink and R. Gophna, Shoham (North). Late Chalcolithic Burial Cave in the Lod Valley, Israel: 163. Jerusalem, Israel Antiquities Authority. Chanteau, J. 2014. The Chalcolithic Shrine at EnGedi. Aesthetics – Symbolism – Structure. In P. Bieliński, M. Gawlikowski, R. Koliński, D. Ławecka, A. Sołtysiak and Z. Wygnańska (eds), Proceedings of the 8th International Congress on the Archaeology of the Ancient Near East 30 April – 4 May 2012, University of Warsaw: 509-526. Wiesbaden, Harrassowitz Verlag. Commenge, C. 2006. Gilat’s ceramics: cognitive dimensions of pottery production. In T. E. Levy (ed.) Archaeology Anthropology and Cult. The Sanctuary at Gilat, Israel: 394-506. London, Equinox. Commenge, C., Levy, T. E., Alon, D. and Kansa, E. 2006. Gilat’s Figurines: Exploring the Social and Symbolic Dimesions of Representation. In T. E. Levy (ed.) Archaeology, Anthropology and Cult. The Sanctuary at Gilat, Israel: 739-830. London, Equinox. Davidovich, U. 2008. The Late Chalcolithic Period in the Judean Desert. Settlement Patterns and Characteristics of the Material Culture as a Basis for Environmental and Social Reconstruction. Unpublished MA Thesis, The Hebrew University of Jerusalem. Drabsch, B. 2015a. The Mysterious Wall Paintings of Teleilat Ghassul, Jordan in Context. Oxford, Archaeopress. Drabsch, B. 2015b. The Wall Art of Teleilat Ghassul, Jordan: When, Where, Why, to Whom and by Whom? Expression 8: 50-57. Drabsch, B. and Bourke, S. J. 2014. Ritual, art and society in the Levantine Chalcolithic: the ‘Processional’ wall painting from Teleilat Ghassul. Antiquity 88: 1081-1098. 125

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Eldar, I. and Baumgarten, Y. 1985. Neve Noy, a Chalcolithic site of the Beer Sheba culture. Biblical Archaeologist 48: 134-139. Elliott, C. 1977. The religious beliefs of the Ghassulians c. 4000-3100 B.C. Palestine Exploration Quarterly 109: 3-25. Elliott, C. 1978. The Ghassulian culture in the Levant. Levant x: 37-54. Epstein, C. 1978. Aspects of symbolism in Chalcolithic Palestine In R. Moorey and P. Parr (eds), Archaeology in the Levant. Essays for Kathleen Kenyon: 23-35. Warminster, Aris and Phillips Ltd. Fabian, P. 2012. Horbat Qarqar. Preliminary Report. Excavations and Surveys in Israel [Online], 124. [Accessed 18 September 2012]. Fabian, P. and Goren, Y. 2002. The Stone Artifacts. In Y. Goren and P. Fabian, Kissufim Road.A Chalcolithic Morturary Site: 44-48. Jerusalem, Israel Antiquities Authority. Fabian, P., Scheftelowitz, N. A. and Gilead, I. 2015. Horvat Qarqar South: Report on a Chalcolithic Cemetery near Qiryat Gat, Israel. Israel Exploration Journal 65: 1-30. Gal, Z., Shalem, D. and Smithline, H. 2011. The Peqi‘in Cave: A Chalcolithic Cemetery in Upper Galilee, Israel. Near Eastern Archaeology 74: 196206. Gal, Z., Smithline, H. and Shalem, D. 1997. A Chalcolithic burial cave in Peqi’in, Upper Galilee. Israel Exploration Journal 47: 145-154. Gates, M.-H. 1992. Nomadic pastoralists and the Chalcolithic hoard from Nahal Mishmar. Levant 24: 131-139. Gilead, I. 1987. A new look at Chalcolithic Beer-Sheba. Biblical Archaeologist 50: 110-117. Gilead, I. 1989. Grar: a Chalcolithic site in Nahal Grar, northern Negev, Israel. Journal of Field Archaeology 16: 377-394. Gilead, I. 1994. The history of the Chalcolithic settlement in the Nahal Beer Sheva area: the radiocarbon aspect. Bulletin of the American Schools of Oriental Research 296: 1-13. Gilead, I. 2002. Religio-magic behavior in the Chalcolithic period of Palestine. In S. Ahituv and E. D. Oren (eds), Studies in Archaeology and Related Disciplines, Aaron Kempinski Memorial Volume: 103128. Beer Sheva, Ben-Gurion University of the Negev Press. Gilead, I. 2009. The Neolithic-Chalcolithic Transition in the Southern Levant: Late Sixth-Fifth Millennium Culture History. In J. J. Shea and D. E. Lieberman (eds), Transitions in Prehistory: Essays in Honor of Ofer Bar-Yosef: 339-359. Oxford, Oxbow Books for the American Schools of Prehistoric Research. Gilead, I. 2011. Chalcolithic culture history: the Ghassulian and other entities in the southern

Levant. In J. L. Lovell and Y. M. Rowan (eds), Culture, Chronology and the Chalcolithic. Theory and Transition: 12-24. Oxford and Oakville, The Council for British Research in the Levant and Oxbow Books. Gilead, I. and Gošić, М. 2014. Fifty Years Later: a Critical Review of Context, Chronology and Anthropology of the Cave of the Hoard in Nahal Mishmar. Mitekufat Haeven – Journal of the Israel Prehistoric Society 44: 226-239. Gilead, I., Rosen, S. and Fabian, P. 1991. Excavations at Tell Abu-Matar (the Hatzerim Neighborhood), Beer Sheva. Mitekufat Haeven – Journal of the Israel Prehistoric Society 24: 173-179. Golden, J. M. 2009. Dawn of the Metal Age. Technology and Society during the Levantine Chalcolithic. London and Oakville, Equinox. Gonen, R. 1992. The Chalcolithic Period. In A. BenTor (ed.) The Archaeology of Ancient Israel: 40-80. New Haven, Yale University Press. Gopher, A. and Tsuk, T. 1996. The Nahal Qanah Cave, Earliest Gold in the Southern Levant. Tel Aviv, Tel Aviv University. Gophna, R. and Lifshitz, S. 1980. A Chalcolithic burial cave at Palmahim. ‘Atiqot 14: 1-8. Goren, Y. 1995. Shrines and Ceramics in Chalcolithic Israel: the view through the petrographic microscope. Archaeometry 37: 287-305. Goren, Y. 2002. The Pottery Assemblage. In Y. Goren and P. Fabian, Kissufim Road. A Chalcolithic Morturary Site. Israel Antiquities Authority Reports 16.: 21-41. Jerusalem, Israel Antiquites Authority. Goren, Y. 2006. The technology of the Gilat pottery assemblage: A reassessment. In T. E. Levy (ed.) Archaeology, Anthropology and Cult. The Sanctuary at Gilat, Israel: 369-393. London, Equinox. Goren, Y. 2008. The location of specialized copper production by the lost wax technique in the Chalcolithic southern Levant. Geoarchaeology 23: 374-397. Goren, Y. 2014. Gods, Caves and Scholars: Chalcolithic Cult and Metallurgy in the Judean Desert. Near Eastern Archaeology 77: 260-266. Goren, Y. and Fabian, P. 2002. Kissufim Road. A Chalcolithic Mortuary Site Jerusalem, Israel Antiquities Authority. Gorzalcany, A. 2018. The Chalcolithic Cemetery at Palmahim (North): New Evidence of Chalcolithic Burial Patterns from the Central Coastal Plain. ‘Atiqot 91: 1-94. Gorzalczany, A. 2006. Palmahim. Excavations and Surveys in Israel [Online], 118. [Accessed 10 January 2013]. Gorzalczany, A., Winter-Livneh, R., Dagot, A. and Shustin, V. 2012. Palmahim (North). Preliminary 126

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant Report. Excavations and Surveys in Israel [Online], 124. [Accessed 10 January 2013]. Gošić, M. 2014. Metallurgy, Magic and Social Identities in the Ghassulian Culture of the Southern Levant (ca. 4500-4000). Unpublished Doctoral disertation Doctoral dissertation, Ben-Gurion University of the Negev. Gošić, M. 2015. Skeuomorphism, Boundary Objects and Socialization of the Chalcolithic Metallurgy in the Southern Levant. Issues in Ethnology and Anthropology 10: 717-740. Gošić, М. and Gilead, I. 2015a. Casting the Sacred – Chalcolithic Metallurgy and Ritual in the Southern Levant. In N. Laneri (ed.) Defining the Sacred: Approaches to the Archaeology of Religion in the Near East: 161-175. Oxford, Oxbow. Gošić, М. and Gilead, I. 2015b. Unveiling Hidden Rituals: Ghassulian Metallurgy of the Southern Levant in Light of Ethnohistorical Record. In K. Rosińska-Balik, A. Ochał-Czarnowicz, M. Czarnowicz and J. Dębowska-Ludwin (eds), Copper and Trade in the South–Eastern Mediterranean. Trade routes of the Near East in Antiquity: 25-37. Oxford, Archaeopress. Haas, N. and Nathan, H. 1973. An attempt at a social interpretation of the Chalcolithic burials in the Nahal Mishmar caves. In Y. Aharoni (ed.) Excavations and Studies: 144-153. Tel Aviv, Tel Aviv University. (Hebrew; English summary: pp. XVIIXVIII) Helms, S. W. 1987. Jawa, Tell um Hammad and the EB I/Late Chalcolithic Landscape. Levant 19: 4981. Hennessy, J. B. 1969. Preliminary report on the first season of excavations at Teleilat Ghassul. Levant 1: 1-24. Hennessy, J. B. 1982. Teleilat Ghassul: its place in the Archaeology of Jordan. In A. Hadidi (ed.) Studies in the History and Archaeology of Jordan: 55-58. Amman, Department of Antiquities. Hodder, I. and Hutson, S. 2003. Reading the Past: Current Approaches to Interpretation in Archaeology. Cambridge, Cambridge University Press. Ilan, D. and Rowan, Y. M. 2012. Deconstructing and Recomposing the Narrative of Spiritual Life in the Chalcolithic of the Southern Levant (45003600 B.C.E.). Archeological Papers of the American Anthropological Association [Online], 21. Joffe, A., Dessel, J. P. and Hallote, R. S. 2001. The ‘Gilat Woman’ female iconography, Chalcolithic cult and the end of southern Levantine prehistory. Near Eastern Archaeology 64: 9-23. Kempinski, A. and Reich, R. (eds) 1992. The Architecture of Ancient Israel. Jerusalem: Israel Exploration Society. Key, C. A. 1980. The trace-element composition of the copper and copper alloy artifacts of the Nahal

Mishmar hoard. In P. Bar-Adon, The Cave of the Treasures: 238-243. Jerusalem, Israel Exploration Society. Khalaily, H. 2002. Chalcolithic and Early Bronze Age pottery and other findes from Caves VII/9 and VIII/28. ‘Atiqot 42: 129-131. Koeppel, R. 1938. Tuleilat Ghassul. Quarterly of the Department of Antiquities of Palestine VI: 225226. Koeppel, R., Senes, H., Murphy, J. W. and Mahan, G. S. 1940. Teleilat Ghassul II (1932-1936). Rome, Pontifical Biblical Institute. Langgut, D., Yahalom-Mack, N., Lev-Yadun, S., Kremer, E., Ullman, M. and Davidovich, U. 2016. The earliest Near Eastern wooden spinning implements. Antiquity 90: 973-990. Levy, T. E. 1995. Cult, metallurgy and ranked societies – the Chalcolithic period (ca. 4500-3500 BCE). In T. E. Levy (ed.) The Archaeology of Society in the Holy Land: 226-245. London, Leicester University Press. Levy, T. E. (ed.) 2006a. Archaeology, Anthropology and Cult, the Sanctuary at Gilat, Israel. London: Equinox. Levy, T. E. 2006b. Conclusion: The Evolution of a Levantine Prehistoric Regional Cult Center. In T. E. Levy (ed.) Archaeology, Anthropology and Cult. The Sanctuary at Gilat, Israel: 831-846. London, Equinox. Levy, T. E. and Alon, D. 1982. The Chalcolithic Mortuary Site near Mezad Aluf, Northern Negev Desert: A Preliminary Study. Bulletin of the American Schools of Oriental Research 248: 37-59. Levy, T. E. and Alon, D. 1987a. Excavations in Shiqmim Cemetery 3: Final Report on the 1982 Season. In T. E. Levy (ed.) Shiqmim I. Studies Concerning Chalcolithic Societies in the Northern Negev Desert, Israel (1982-1984): 333-355. Oxford, B.A.R. Levy, T. E. and Alon, D. 1987b. Excavations in the Shiqmim Village. In T. E. Levy (ed.) Shiqmim I. Studies Concerning Chalcolithic Societies in the Northern Negev Desert, Israel (1982-1984): 153-218. Oxford, Archaeopress. Levy, T. E., Alon, D., Goldberg, P., Grigson, C., Smith, P., Buikstra, J., Holl, A., Rowan, Y. M. and Sabari, P. 1994. Protohistoric investigations at the Shiqmim Chalcolithic village and cemetery: interim report on the 1988 season. In W. G. Dever (ed.) Preliminary Excavations Reports Sardis, Paphos, Caesarea Maritima, Shiqmim, Ain Ghazal: 87-106. American Schools of Oriental Research. Levy, T. E., Alon, D., Goldberg, P., Grigson, C., Smith, P., Buikstra, J., Holl, A., Rowen, Y. and Safari, P. 1993a. Protohistoric Investigations at the Shiqmim Chalcolithic Village and Cemetery: Interim Report on the 1988 Season. In W. G. Dever (ed.) Preliminary Excavation Reports: Sardis, 127

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Paphos, Caesarea Maritima, Shiqmim, Ain Ghazal 87106. Levy, T. E., Alon, D., Goldberg, P., Grigson, C., Smith, P., Buikstra, J., Holl, A. F.-C., Rowan, Y. M. and Sabari, P. 1993b. Protohistoric investigations at the Shiqmim Chalcolithic village and cemetery: interim report on the 1988 season. Annual of the American School of Oriental Reserach 51: 87106. Levy, T. E., Alon, D., Goldberg, P., Grigson, C., Smith, P., Holl, A., Buikstra, J., Shalev, S., Rosen, S., Ben Itzhak, S. and Ben Yosef, A. 1991. Protohistoric investigations at the Shiqmim Chalcolithic village and cemetery: interim report on the 1987 season. Bulletin of the American Schools of Oriental Research Supplement 27: 29-46. Levy, T. E., Alon, D., Rowan, Y. M. and Kersel, M. 2006. The Sanctuary Sequence: Excavations at Gilat: 1975-77, 1989, 1990-92. In T. E. Levy (ed.) Archaeology, Anthropology and Cult. The Sanctuary at Gilat, Israel: 95-212. London, Equinox. Levy, T. E. and Burton, M. 2006. Radiocarbon dating of Gilat. In T. E. Levy (ed.) Archaeology, Anthropology and Cult The Sanctuary at Gilat, Israel: 863-866. London, Equinox. Levy, T. E. and Golden, J. M. 1996. Syncretism and Mnemonic Dimesions of Chalcolithic Art. Biblical Archaeologist 59: 150-159. Levy, T. E. and Menachem, N. 1987. The pottery from the Shiqmim village: topological and spatial considerations. In T. E. Levy (ed.) Shiqmim I, studies concerning Chalcolithic societies in the Northern Negev desert, Israel: 313-331. Oxford, B.A.R. Lovell, J. L. 2001. The Late Neolithic and Chalcolithic Periods in the Southern Levant: New Data from the Site of Teleilat Ghassul, Jordan. Oxford, Archaeopress. Mallon, A., Koeppel, R. and Neuville, R. 1934. Teleilat Ghassul I. Rome, Institut Biblique Pontifical. Mazar, A. 2000. A Sacred Tree in the Chalcolithic Shrine at En-Gedi: A Suggestion. Bulletin of the Anglo-Israel Society 18: 31-36. Merhav, R. 1993. Scepters of the divine from the Cave of the Treasure at Nahal Mishmar. In M. Heltzer, A. Segal and D. Kaufman (eds), Studies in the archaeology and history of ancient Israel in honour of Moshe Dothan: 21-42. Haifa, Haifa University Press. (Hebrew) Milevski, I. 2002. A new fertility figurine and new animal motifs from the Chalcolithic in the southern Levant: finds from cave K-1 at Quleh, Israel. Paléorient 28: 133-141. Mirocshedji, P. D. 1993. Cult and religion in the Chalcolithic and Early Bronze Age. In A. Biran and J. Aviram (eds), Biblical archaeology today 1990: 208-220. Jerusalem, Israel Exploration Society, The Israel Academy of Sciences and Humanities.

Namdar, D., Segal, I., Goren, Y. and Shalev, S. 2004. Chalcolithic Copper Artifacts. In N. Scheftelowitz and R. Oren, Giv’at ha-Oranim. A Chalcolithic Site: 70-83.Tel Aviv, Emery and Claire Yass Publications in Archaeology of the Institute of Archaeology, Tel Aviv University. Neuville, R. 1930. Notes de Préhistore Palestinienne. The Journal of the Palestine Oriental Society 10: 193221. North, R. S. J. 1961. Ghassul 1960, excavation report. Rome, Pontifical Biblical Institute. Perrot, J. 1955. The excavations at Tell Abu Matar near Beersheba. Israel Exploration Journal 5: 17-40, 73-84, 167-189. Perrot, J. 1958. Note sur la 6e campagne. Israel Exploration Journal 8: 122-123. Perrot, J. 1959. Les statuettes en ivoire de Beershéva. Syria 34: 6-19. Perrot, J. 1964. Les ivoires de la 7e campagne a Safadi près de Beershéva. Eretz-Israel 7: 92*-93*. Perrot, J. 1992. Umm Qatafa and Umm Qala’a: two ‘Ghassulian’ caves in the Judean Desert. Eretz Israel 23: 100*-111*. Perrot, J. and Ladiray, D. 1980. Tombes à Ossuaires de la Région Côtiere Palestinienne au IVe Millénaire Avant l’ère Chrétiene. Paris, Association Paléorient. Porath, Y. 2006. Chalcolithic burial sites at Ma`abarot and Tel Ifshar. ‘Atiqot 53: 45-63. Potaszkin, R. and Bar-Avi, K. 1980. A material investigation of the metal objects from the Nahal Mishmar treasure. In P. Bar-Adon, The Cave of the Treasure: 235-237. Jerusalem, Israel Exploration Society. Price, M. D., Hill, A. C., Rowan, Y. M. and Kersel, M. 2016. Gazelles, Liminality, and Chalcolithic Ritual: A Case Study from Marj Rabba, Israel. Bulletin of the American Schools of Oriental Research 376: 7-27. Rappaport, R. A. 1999. Ritual and Religion in the Making of Humanity. Cambridge, Cambridge University Press. Reich, R. and Katzenstein, H. 1992. Glossary of Architectural Terms. In A. Kempinski and R. Reich (eds), The Architecture of Ancient Israel: 311-322. Jerusalem, Israel Exploration Society. Reimer, P. J., Bard, E., Bayliss, A., Beck, J. W., Blackwell, P. G., Bronk Ramsey, C., Buck, C. E., Cheng, H., Edwards, R. L., Friedrich, M., Grootes, P. M., Guilderson, T. P., Haflidason, H., Hajdas, I., Hatté, C., Heaton, T. J., Hoffmann, D. L., Hogg, A. G., Hughen, K. A., Kaiser, K. F., Kromer, B., Manning, S. W., Niu, M., Reimer, R. W., Richards, D. A., Scott, E. M., Southon, J. R., Staff, R. A., Turney, C. S. M. and Van Der Plicht, J. 2013. IntCal13 and Marine13 Radiocarbon Age Calibration Curves 0-50,000 Years cal BP. Radiocarbon 55: 1869-1887. 128

M. Gošić: Going Through Customs: Changing Rituals of the Ghassulian Culture of the Southern Levant Rosen, S. A. and Eldar, I. 1993. Horvat Beter revisited: the 1982 salvage excavations. ‘Atiqot 22: 1327. Rowan, Y. M. and Golden, J. M. 2009. The Chalcolithic Period of the Southern Levant: A Synthetic Review. Journal of World Prehistory 22: 1-92. Rowan, Y. M. and Ilan, D. 2007. The Meaning of Ritual Diversity in the Chalcolithic of the Southern Levant. In D. A. Barrowclough and C. Malone (eds), Cult in Context. Reconsidering Ritual in Archaeology: 249-256. Oxford, Oxbow Books. Rowan, Y. M. and Ilan, D. 2013. The Subterranean Landscape of the Southern Levant during the Chalcolithic Period. In H. Moyes (ed.) Sacred darkness: A Gobal Perspective on the Ritual Use of Caves: 87-107. Boulder, University Press of Colorado. Rowan, Y. M., Levy, T. E., Alon, D. and Goren, Y. 2006. Gilat’s Ground Stone Assemblage: Stone Fenestrated Stands, Bowls, Palettes and Related Artifacts. In T. E. Levy (ed.) Archaeology, Anthropology and Cult. The Sanctuary at Gilat, Israel: 575-684. London, Equinox. Scheftelowitz, N. and Oren, R. 2004. Giv’at Ha-Oranim. A Chalcolithic Site. Tel Aviv, Emery and Claire Yass Publications in Archaeology of the Institute of Archaeology. Schick, T. 1998. The Cave of the Warrior. A Fourth Millennium Burial in the Judean Desert. Jerusalem, The Israel Antiquities Authority. Seaton, P. 2008. Chalcolithic Cult and Risk Management at Teleilat Ghassul. The Area E Sanctuary. Oxford, Archaeopress. Segal, D., Carmi, I., Gal, Z., Smithline, H. and Shalem, D. 1998. Dating a Chalcolithic burial cave in Peqi‘in, Upper Galilee, Israel. Radiocarbon 40: 707712. Segal, I. 2002. The Copper Axe at Cave V/49. Atiqot 41: 99-100. Segal, I. and Goren, Y. 2013. A Chemical, Metallograpical, Isotopic and Petrographic Study of the Copper Finds. In D. Shalem, Z. Gal and H. Smithline, Peqi‘in. A Late Chalcolithic Burial Site, Upper Galilee, Israel: 379-385. Kinneret, Kinneret Academic Collage, Institute for Galilean Archaeology and Ostracon. Shalem, D. 2008. Iconography on Ossuaries and Burial Jars from the Late Chalcolithic Period in Israel in the Context of the Ancient Near East. Unpublished Ph.D dissertation, University of Haifa. Shalem, D., Gal, Z. and Smithline, H. 2013. Peqi‘in. A Late Chalcolithic Burial Site, Upper Galilee, Israel. Kinneret, Kinneret Academic Collage, Institute for Galilean Archaeology and Ostracon. Shalev, S. 1996. Metallurgical and Metalographic Studies. Copper objects. In A. Gopher and T. Tsuk, The Nahal Qanah Cave. Earliest Gold in the

Southern Levant. Tel Aviv, Tel Aviv University Press. Shalev, S. and Northover, P. J. 1987. Chalcolithic metalworking from Shiqmim. In T. E. Levy (ed.) Shiqmim I, Studies Concerning Chalcolithic Societies in the Northern Negev Desert, Israel (1982-1984): 357371. Oxford, B.A.R. Smith, P., Zagerson, T., Sabari, P., Golden, J., Levy, T. E. and Dawson, L. 2006. Death and the sanctuary: the human remains from Gilat. In T. E. Levy (ed.) Archaeology, Anthropology and Cult. The Sanctuary at Gilat: 327-366. London, Equinox. Tadmor, M. 1989. The Judean Desert Treasure from Nahal Mishmar: A Chalcolithic Traders’ Hoard? In A. J. Leonard and B. B. Williams (eds), Essays in Ancient Civilization presented to Helen J. Kantor: 249-261. Chicago, The Oriental Institute of the University of Chicago. Tadmor, M., Kedem, D., Begemann, F., Hauptmann, A., Pernicka, E. and Schmitt-Strecker, S. 1995. The Nahal Mishmar Hoard from the Judean Desert: Technology, Composition, and Provenance. ‘Atiqot 27: 96-148. Ussishkin, D. 1971. The ‘Ghassulian’ Temple in Ein Gedi and the Origin of the Hoard from Nahal Mishmar. Biblical Archaeologist 34: 23-39. Ussishkin, D. 1980. The Ghassulian shrine at En-Gedi. Tel Aviv 7: 1-44. Ussishkin, D. 2014. The Chalcolithic Temple in En Gedi: Fifty Years after Its Discovery Near Eastern Archaeology 77: 15-26. Van Den Brink, E. C. M. 2005. Chalcolithic burial caves in coastal and inland Israel. In E. C. M. Van Den Brink and R. Gophna, Shoham (North): Late Chalcolithic Burial Caves in the Lod Valley, Israel: 175-189. Jerusalem, Israel Antiquities Authority. Van Den Brink, E. C. M., Horowitz, L. and Khalaily, H. 2004. A Chalcolithic Dwelling and Burial Cave at Horvat Castra. Israel Exploration Journal 54: 129153. Walker, W. H. and Lucero, L. J. 2000. The depositional history of ritual and power. In M.-A. Dobres and J. E. Robb (eds), Agency in archaeology: 130-147. London, Routledge. Weinstein, J. M. 1984. Radiocarbon Dating in the Southern Levant. Radiocarbon 26: 297366. Winter-Livneh, R., Svoray, T. and Gilead, I. 2012. Secondary burial cemeteries, visibility and land tenure: A view from the southern Levant Chalcolithic period. Journal of Anthropological Archaeology 31: 423–438. Yannai, E. and Porath, Y. 2006. A Chalcolithic burial cave at Et-Taiyiba. ‘Atiqot 53: 1-44. Yellin, J., Levy, T. E. and Rowan, Y. M. 1996. New Evidence on Prehistoric Trade Routes: The 129

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Obsidian Evidence from Gilat, Israel. Journal of Field Archaeology 23: 361-368.

Yossi, N. 2018. The Human Remains from the Chalcolithic Cemetery at Palmaḥim (North). ‘Atiqot 91: 95-96.

130

Animal Offerings from the Nawamis Fields and Coeval Habitation Sites in Southern Sinai Liora Kolska Horwitz tomb fauna from southern Sinai and compares these assemblage to remains recovered from roughly contemporaneous habitation sites in the region.

Introduction Salient architectural features in the landscape of southern Sinai are the nawamis. These are circular or beehive-shaped, above-ground structures built of local field stones, with a lintel above the entrance (Fig. 1). These structures were first documented in the Sinai Peninsula, and excavated in the late 19th century by members of the Ordnance Survey of Sinai, Holland (1870, 1871) and Palmer (1871). Subsequently, several archaeological missions undertook further investigations and excavation. Most prominent among the archaeologists involved were Currelly (1906), who worked with the Petrie expedition, Albright (Albright 1948a,b), Rothenberg (1972), and most recently Goren, Bar-Yosef, Hershkovitz and colleagues (Bar-Yosef et al. 1977, 1983; Goren 1980, 1998; Hershkovitz et al. 1982).

Background Location: With the exception of one group comprising three structures, that is located in Wadi Fuqiyah on the Tih plateau (Rothenberg 1975), all the fields are located in southern Sinai. They form an arc around the high mountains, with Saint Catherine’s monastery at the centre (Fig. 3; Goren 1998). None are located in the high mountains. However, most of the fields are located on hilltops and in close proximity to ancient routes. Especially rich concentrations of nawamis are found in Wadi Solaf, the southern part of Wadi Nasb, and along Nakb Hibran (Goren 1998). Architecture: The nawamis structures are double-walled with the outer wall standing straight while the inner wall makes an arch to form a corbelled roof. They were built of dry-stone, usually slabs of local granite or sandstone that have not been shaped. The structures range in size from 3-6 m in diameter and between 1.802.10 m in height (Bar-Yosef et al. 1977, 1983; Goren 1980, 1998) and are usually covered with a stone slab. The fill within the structures was usually composed of gravel and dust/ash or in some instances stone slabs. There is only one entrance per structure and in all cases this faces west or, as at the ‘Ein Huderah nawamis, westsouth-west. The western orientation of the door of the nawamis has been interpreted as reflecting a symbolic association between death on earth and the sun setting in the west suggesting a relationship with Egyptian beliefs of the afterlife (for details see Bar-Yosef et al. 1983; Goren 1980, 1998; Hershkovitz et al. 1985).

With the exception of Palmer, who proposed that the nawamis served as habitations, most researchers have concluded that they functioned as tombs. Nawamistype structures (also known as ‘high circular tombs’) are known not only from Sinai, but also from desert regions of eastern Jordan, central Saudi Arabia through to Yemen (Rowan et al. 2011; McCorriston et al. 2011), and possibly as far south as Sudan (Murray 1935). A total of 22 nawamis clusters (fields) are known in Sinai (Avner 2002; Bar-Yosef et al. 1998; Goren 1998). They vary in size and number of tombs, from several to more than 200 individual structures (Fig. 2), identified at sites such as Wadi Solaf and Wadi Hebran that extend over an area of some 5 km. As part of the surveys and excavations undertaken by Israeli archaeologists in Sinai from 1971 through 1982, 21 such fields were investigated, and nine fields were excavated (Fig. 2; Bar-Yosef et al. 1977, 1983, 1986; Goren 1980, 1998; Rothenberg 1972a, 1975, 1979).

Human Remains: Nearly all the excavated nawamis contained human remains, though bone preservation was extremely poor (Bar-Yosef et al. 1983:57). They were often multiple burials representing both sexes and all ages (Hershkovitz 1987; Hershkovitz et al. 1982). Secondary burials were the most common but primary flexed burials were also found. As such it is possible that, as suggested by Bar-Yosef et al. (1977), many of the secondary burials actually represent disturbed primaries. This is corroborated by the

Although the architecture and some of the material finds from the excavated nawamis have been published, the animal remains have not (with the exception of a footnote in Bar-Yosef et al. 1977). Fauna recovered from ritual or mortuary contexts can provide useful insights into the cosmology and spiritual beliefs of past communities. Consequently, they can complement findings obtained from other archaeological sources. This paper provides a description of the nawamis 131

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 1. Example of a nawamis from Ein Huderah (Photograph: Uzi Avner).

Figure 2. Photograph showing a section of the nawamis field at Ein Huderah (Photograph: Uzi Avner).

spatial patterning of human bones, most of which were found adjacent to the inner walls indicating that older interments had been pushed aside to make room for newer ones (Bar-Yosef et al. 1977). On the basis of the age and sex of the interred, as well as the presence of a congenital pathology, it has been suggested that the nawamis represent family tombs (Bar-Yosef et al. 1977, 1986; Goren 1980, 1998; Hershkovitz et al. 1982).

especially common in ritual and mortuary sites in the southern Levantine deserts. Pottery is rare in the nawamis but finds include a few sherds of Egyptian Dynasty I vessels and a fragment of a churn. The most abundant finds were ornaments, especially beads, made of carnellian, faience, turquoise, hematite, copper, ostrich eggshell, bone and marine shell, as well as bracelets, bangles and pendants also made of marine shell (Avner 2002; Bar Yosef et al. 1977, 1986; Bar-Yosef Mayer 1999; 2002a, b; Goren 1980; AradAyalon n.d; Hovers 1981; Ilan and Sebbane 1989; Sass 1980).

Material Culture: Remains of cloth were found in some nawamis in addition to a wide range of artifacts, presumably representing grave goods. Items recovered included: bone and wooden points, arrows with their shafts, copper awls, flint artefacts – especially tabular scrapers, transverse arrowheads, groundstone utensils, the latter made of basalt, sandstone and limestone. Avner (2002:77) notes that tabular scrapers are

Chronology and Cultural Connections: On the basis of the lithic artefacts, copper implements, copper and shell beads and shell bracelets recovered, Currelly 132

L. Kolska Horwitz: Animal Offerings from the Nawamis Fields and Coeval Habitation Sites

Figure 3. Map showing location of the nine nawamis fields in southern Sinai excavated by Israeli teams (based on Bar-Yosef Mayer 2002: Fig. 1).

(1906) attributed the nawamis to the Predynastic period of Egypt. While some of the ornaments recovered are unique to the Sinai, the presence of transverse arrowheads, faience beads and ceramics all point to an Egyptian Early Dynastic connection.

Chalcolithic – Early Bronze, while Stager (1992) placed them at the end of the fourth millennium BC i.e. Early Bronze Age I. Following more extensive excavation by the Israeli teams, two of the nawamis were radiocarbon dated (Table 1). These dates placed them in the fourth millennium BC i.e. Late Chalcolithic/Early Bronze I {EBI] (Bar-Yosef et al. 1977, 1983, 1986; Carmi and Segal 1994; Rothenberg 1972a). Subsequently, Avner (2002:

Later researchers were divided in their assessment with Rothenberg (1972) dating the nawamis to the 133

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Nawamis Tombs

Lab No.

C uncal. BP

C cal. BP

Calibrated BC

Reference

Ein Um Ahmed

RT-1856

5815±50

6618±63

4770-4600

Carmi and Segal 1994

Ein Um Ahmed

RT-1857

5575±60

6371±50

4455-4355

Carmi and Segal 1994

Ein Um Ahmed

RT-1851

5130±55

5858±76

3990-3800

Carmi and Segal 1994

Abu Khalil

RT-1853

5200±70

6003±95

4220-3940

Carmi and Segal 1994

14

14

Table 1. Radiocarbon dates for the nawamis. (Dates from Segal and Carmi 1996; calibration from Avner 2002: Table 1, based on OxCal 3.4).

77) suggested that they may have been built earlier i.e., in the 5th millennium BC, though he concurs that the majority of the nawamis finds are Chalcolithic in age with some continuation into the EB I as evidenced by copper awls and tabular scrapers (Avner 2002: 59; BarYosef et al. 1977, 1986; Ilan and Sebbane 1989). These finds also attest to ties with sites in the southern Levant. This conclusion is borne out by the findings of Bar-Yosef Mayer (1999, 2002a, b) who analysed the malacological assemblages recovered from nine clusters of nawamis. She noted that the species of marine shells and types of shell ornaments recovered are the same as those commonly found in sites dating to the Chalcolithic and EB I periods in Israel and Jordan. In terms of chronology, Conus shell apex beads are known only from EBAII/III sites in the southern Levant, perhaps indicating that the nawamis continued to be used up until this period. In conclusion, the tombs appear to contain material items deriving from multiple cultural and chronological sources.

During excavations, intrusive items of glass and plastic were found in some nawamis and may be due to natural collapse, tomb robbery or re-use by the Bedouin as documented by Goren (1998: 62) for the nawamis in Wadi Solaf (Fig. 4). Fauna from the Nawamis Animal remains were recovered from eight of the nine nawamis fields excavated by the Israeli teams: Ein Um Ahmed, Gebel Gunna, Wadi El Abar, Hzeimeh (also termed Upper Wadi Nasb), Wadi Sawawin, Abu Halil and Nakb Hibran (Fig. 2). Bone preservation in all sites was very poor and sample sizes extremely small. Consequently, the faunal remains are described below and listed in Table 2 by nawamis field rather than individual tomb. All bones were identified with reference to the zoological collection of The Hebrew University of Jerusalem. Since it is possible that both domestic goats and ibex are represented in the assemblage, age estimates based on tooth eruption and bone fusion are given from both Silver (1969) – for unimproved domestic goats and Noddle (1974) – for feral goats. The latter was thought to be a better proxy for ibex fusion ages in the absence of bone fusion data for this species. Bones and teeth were measured following standards given in von den Driesch (1976). Ein Um Ahmed: The site is located on a tributary of Wadi El Ein. Based on the field notes, between 13 and 15 nawamis were excavated in this field (Bar-Yosef Mayer 1999).

Figure 4. Collapsed nawamis photographed ca. 1990-1920 by G. Eric and Edith Matson. Open access print, Library of Congress, Prints & Photographs Division, Photograph Collection – Reproduction number: LC-DIG-matpc-01976.

134

A total of nine animal remains were recovered from this nawamis field of which 2 were unidentified fragments (Table 2). One bone (distal scapula) was tentatively identified as belonging to a domestic goat given its small size (measurements in mm: BG 14.4; SD 15.3). However, since it is an immature animal it cannot be excluded that it represents an immature ibex. The species attribution of a further six remains (teeth and a

135

tooth

Shark

7

Total ID

5 tooth frags.

1st ph F 1 tooth frag. 1 upper molar 1 metapodial shaft

1 incisor Meriones sp.

1 femur 1 tibia

Table 2. Faunal remains from the nawamis tombs (given as NISP counts).

1 dist. humerus 1 pelvis 1 femur dist. 5 longbone frags. 1 rib

1 pelvis

1 left & 1 right jaw 1 dist. humerus 2 radius shafts 1 rib frag.

1 jaw Uromastyx

15

0

49

1 jaw Uromastyx

Reptile

1 left humerus prox.-UF 1 patella frag. 1 premaxilla frag. 9 tooth frags. 1 left jaw corpus frag. 5 rib shafts 10 prox. rib frags. 1 cervical vertebra – UF 1 thoracic vertebra – UF 2 vertebral frags. 1 incisor Meriones sp. 1 longbonemisc. rodent

Rodent

3

1 sacrum 1 lumbar vertebra 1 rib

1 rib

Hare Lepus capensis

1 goat horncore

1 upper 2nd molar 3 tooth frags. 1 right scapula dist.-UF 2 right humerus shaft frags.

Goat/Ibex Capra sp.

frag. = fragment; prox. = proximal; dist.= distal; UF = unfused. * number cited in Bar-Yosef et al. 1977: Table 2, minus the three identified horncores listed here.

1 tooth frag.

2 longbone frags.

10 tooth frags.

Gazelle/Caprine Gazella/Capra

Nakb Hibran

1 horncore 1 horncore tip

Gazelle Gazella sp.

7 tooth frags.

1 worked pelvis

Cattle Bos

Abu Halil

Wadi Sawawin

Hzeimeh (Upper Wadi Nasb)

Wadi El Abar

Ein Huderah

Ein Um Ahmed

Species Sites

6 burnt frags. 7. unburnt frags.

13 frags.

35 bones*

2 frags.

Total Non-Id

L. Kolska Horwitz: Animal Offerings from the Nawamis Fields and Coeval Habitation Sites

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead humerus shaft) is even less certain and they represent an undetermined caprine – either sheep, goat or ibex. Irrespective of whether bone fusion and tooth eruption rates are used based on Silver (1969) or Noddle (1974), remains of a goat and an unidentified caprine from this nawamis were aged 12 months or slightly less (based on an unfused distal scapula and an erupted but unworn upper second molar). Since there is no duplication of skeletal elements or differences in age, the estimated Minimum Number of Individual animals (MNI) is probably one goat. The upper molar indicates that a cranium was originally present in the tomb, while the right scapula and humerus may indicate the presence of a joint of the same upper forelimb.

be possible to add to the list of items traded between Sinai and regions to the north, hippopotamus ivory or ornaments made of this material. Such trade may have been facilitated by Canaanites coming to Sinai to mine turquoise, as proposed by Beit-Arieh (2003). However, the option that the ivory from the nawamis originated in the Nile Valley, cannot be negated. This, since there is evidence for the movement of other faunal items – shells and fish – from the Nile to Chalcolithic and Early Bronze Age sites in the Mediterranean region of Canaan (e.g. Bar-Yosef Mayer 1999, 2002a, b; Reese et al. 1986; Van Neer et al. 2004a), while as noted above, there are many cultural objects of Egyptian origin in the nawamis (Bar-Yosef et al. 1977, 1983, 1986).

Ein Huderah: The field is located on the banks of Wadi ‘Abaya, on a Nubian sandstone plateau. A total of 42 nawamis were identified by Bar-Yosef et al. (1977) at this location. Of these, 24 were cleared and included 18 that had been previously disturbed but were undamaged, and another six that had collapsed.

Wadi El Abar: Forty-nine bones were recovered from the nawamis field at Wadi El Abar of which 30 belonged to caprines – domestic goat, sheep or ibex; a further two were positively attributed to Capra sp. – either ibex or domestic goat; and 10 small tooth enamel fragments to small- medium sized ruminants either gazelle/caprines (Table 2).

No faunal remains from these nawamis were available for study by the author. However, Bar-Yosef et al. note (1977: 78, note 48) that 38 bones were found in this field including three animal bones that were found in association with human burials: a horncore of a female dorcas gazelle (Gazella dorcas), and two horntips – one possibly of a gazelle, the other possibly from a goat (Capra sp.). In a subsequent paper, the species inventory for Ein Huderah is given as: goat, gazelle and cattle (Bar-Yosef et al. 1983: 57). The reference to cattle (Bos sp.) probably relates to a modified cattle pelvis which constitutes an implement (Bar-Yosef 1977: 78). A further eight nawamis in this area that were subsequently excavated by A. Goren did not yield animal remains.

Two caprine vertebrae had unfused epiphyses, representing at least one animal aged less than 4-5 years (Hilzheimer 1961). In addition, an unfused proximal Capra humerus gave an age of 23-24 months, if domestic (Silver 1969), or 48 months, i.e. an ibex (Noddle 1974). The MNI estimate for this assemblage was one immature goat aged ca. 1 year, with all other caprine and small to medium-sized remains probably belonging to the same animal given that there was no overlap in age or skeletal elements represented. Notably, the skeletal remains included elements of the cranium and dentition, trunk and limbs suggesting that this may originally have been an almost complete animal.

Gebel Gunna: No animal remains were recovered from this nawamis field. However, a special find noted by BarYosef et al. (1986: 137) was the presence of ivory beads – possibly of hippopotamus tusk – whose ivory ‘could have come from various parts of the Near East and North-East Africa’.

Small quantities of bones of hare (Lepus capensis), rodent – jird (Meriones sp.) and reptile – mastigure (Uromastyx sp.), were also found. They are all species currently occurring in the region and as such may represent recent intrusions and natural mortalities. Indeed, Currelly (1906: 2440) mentioned that although the entrance to the nawamis he excavated was small, ‘it was necessary to place a flat stone inside to prevent small animals from entering’.

If indeed the ivory derives from hippopotamus, then two close sources for this raw material were at hand – the Nile delta and/or the Israeli Mediterranean coast (Horwitz and Tchernov 1990). In fact, the earliest evidence for hippopotamus in the latter region is an astragalus from the Chalcolithic site of Qatif Y2 (Grigson in press). Additional finds of hippopotamus bones which attest to the continuous presence of this taxon along the southern coast have recently been reported from the EB III levels at Tel es-Sakan, Gaza (Miroschedji et al. 2001). A rich corpus of ornaments and objects made of elephant and hippopotamus ivory are known from the Chalcolithic sites in the Beersheva Basin (e.g. Levy 1995; Perrot 1959), such that it may

The precise origin of an unmodified shark tooth found in this tomb has not been determined. It may represent: an accidental inclusion from earlier geological formations in the region; perforated fossil shark teeth are known from Neolithic sites in Fayyum and Maadi, Egypt (Mayor 2000:175), a paleontological relic collected by the local population and placed in the tomb as a grave offering (like the collection of fossils found in the Timna Temple, or the fossil from Heliopolis, Egypt that originated in Sinai – Mayor 2000: 136

L. Kolska Horwitz: Animal Offerings from the Nawamis Fields and Coeval Habitation Sites Fauna from Habitation Sites in Southern Sinai

175), or finally, a scavenged or fished shark tooth from the Red or Mediterranean Sea coasts (e.g. Cione 2006, on presence of shark in a 7th-6th century BC site in northern Sinai).

It is assumed that the people, who used the nawamis were nomadic pastoralists who subsisted from caprine herding supplemented by exchange of goods with populations living in neighbouring regions. However, the archaeozoological information to support this claim is partial and based primarily on the absence of permanent sites since the habitation sites identified to date more closely resemble ephemeral camps. Examination of the nawamis architecture pointed to these sites having been constructed on a seasonal basis, during the fall and spring (Bar-Yosef et al. 1983), again emphasizing the nomadic character of these communities.

Hzeimeh, also termed Upper Wadi Nasb: Only 13 unidentified mammalian bone fragments were recovered from this field. Wadi Sawawin: A total of 15 animal bones were recovered from this nawamis field (Table 2). A single fused first phalanx of a goat/ibex (Capra sp.) was complete and could be measured (in mm: GL 31.5; Bd 9.9; SD 9.2; Dp 10.6). The relatively small size of this bone suggests that it probably belongs to a domestic goat. Based on the timing of the fusion of the proximal epiphysis this animal was aged at least 1 year of age, while if ibex, the age would probably be closer to 2 years (Noddle 1974). The MNI estimate of goats in this nawamis is one. Other species represented are hare, unidentified rodent and mastigure (reptile).

The few habitation sites in southern Sinai that are roughly contemporeous with the nawamis i.e. spanning the Chalcolithic to Early Bronze Age II, and which have yielded faunal remains, may elucidate this issue. Specifically, such data may offer some support for the identification of the goat remains from the nawamis as those of domestic animals. Moreover, bone remains from domestic sites can fill the gaps in our knowledge of the subsistence activities of these communities and highlight whether the species found in the nawamis represent specially selected taxa rather than the full range of animals exploited.

Abu Halil: Of the 26 faunal remains found in this field, half could be identified to taxon (Table 2). Interestingly, of the unidentified fragments recovered, six were burnt making this the only nawamis to have yielded burnt bones. However, Glycymeris shells (originating in the Mediterranean Sea) found at the Ein Huderah nawamis ‘show black stains of apparently burnt organic matter, but no positive identification could be made’ (Bar-Yosef et al. 1977: 79). It is unclear whether the burning relates to a post-depositional event or occurred during the initial deposition of these finds.

Conclusion

Chalcolithic, Serabit el-Khadim (Site 1105): Only one Chalcolithic site in southern Sinai has been excavated to date; the turquoise mining site at Mount Serabit el-Khadim in southern-western Sinai, dated by radiocarbon to 4216-3985 cal. BC (Bet-Arieh 2003b). Fauna were excavated from a structure at the site and comprised small fragments of sheep/goat long bones, as well as a large fragment of a right cattle (Bos taurus) calcaneus (Bet-Arieh 2003b:97). It was proposed that this find did not attest to cattle herding at the site, but that a joint of beef was brought to the site from areas to the north or from the Egyptian delta recalling the worked cattle pelvis from the Ein Huderah nawamis. The presence of a First Dynastic jar of Egyptian origin at the site indicates that it probably continued to be occupied in the late EB I-early EB II.

A narrow range of animal species were recovered from the nawamis fields in southern Sinai, many (rodents, reptiles, birds, hares) probably represent recent intrusive taxa. The most common animal remains that appear to have served as offerings are bones of medium-sized ruminants – sheep/goat (and/or ibex) and gazelle. Remains of goat (Capra sp.) and dorcas gazelle (Gazella dorcas) were both positively identified. The small size of the few measurable bones supports a tentative identification of the goat bones to domestic animals as opposed to ibex. No remains of sheep have been positively identified.

EB I camps: Several habitation sites or camps, that were located in close proximity to the nawamis were investigated by the Israeli expedition who concluded that the two were associated based on the overall resemblance of their material finds (dating to the EBI and possibly continuing into the EBII), though the habitation sites contained an abundance of daily objects and lithic debitage not found in the nawamis (Goren 1980; Hovers 1981). Likewise, Bar-Yosef Mayer (2002) noted that the molluscan assemblages from the Gunna habitation sites (25, 50, 100) were more varied than those of the nearby Gunna nawamis.

Identified faunal remains were mainly fragments of tooth enamel, some so small as to make it difficult to distinguish whether they derive from gazelle, goat or ibex. The pelvis of a hare was an additional find. Nakb Hibran: A total of 11 bones were identified from this field; one tooth fragment of gazelle/caprine, eight bones of hare and a rodent incisor identified as a species of jird (Table 2).

137

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Excavations of structures at three of the camps yielded faunal remains; Moyat Daba’iyeh, also known as ‘Abu Halil Megurim’ (Goren 1980), Gunna 25 (Bar-Yosef et al. 1986: 149; Goren 1980: 254). and Gunna 50 (Bar-Yosef et al. 1986: 129 ff). In the publication of Guna 25, BarYosef et al. (1986) documented the presence of skull and leg remains of two specimens of Capra, a phalanx of a gazelle, a hare bone and remains of recent rodents. The faunal remains from Gunna 50 were briefly described in Bar-Yosef et al. (1986: 130) who noted that they were poorly preserved and very fragmentary. They listed the presence of 11 bones of Capra sp., 4 of Gazella sp., 7 bones of gazelle/goat and 2 fragments of an ass jaw. The domestic status of the ass and goats was not determined however. Re-examination of the fauna from these three sites indicated a wider spectrum of taxa – small mammals, reptiles and birds and a small component of gazelle and caprines (Table 3). Most importantly, there is a clear predominance of remains belonging to Capra sp. (goat/ibex) at all three sites. This factor, combined with the enormous quantities of goat dung reported by Goren (1998) at the habitation sites, offers compelling evidence for these being domestic goats, despite the absence of morphometric data to support this claim (due to the poor bone preservation).

in Beit-Arieh 1977: 143) and others in southern Sinai, were said to be of a small-sized breed, similar to the black goat that inhabits Sinai today (Hakker-Orion in Beit-Arieh 1977). However, no morphometric data were given to substantiate this claim. Wattiya North (Site No. 1042): Remains of goat were identified from a room adjacent to the courtyard (Hakker-Orion in Beit-Arieh 1977: 143). Feiran I (Site No. 1150): Only 15 well preserved bones could be identified to species. Taxonomic identifications made by Hellwing (reported in Beit-Arieh 1982) list the presence of sheep/goat (Ovis/Capra), gazelle (Gazella sp.), cattle (Bos sp.) and dog (Canis familiaris). Sheikh ‘Awad (Site No. 1118): According to Hellwing (in Beit-Arieh 1981), the assemblage was dominated by remains of a small-sized domestic goat (Capra hircus), similar to the Black Bedouin goat found in the region today. Other taxa represented were dorcas gazelle, ibex, hare, hyrax and dog. Guna 100: This site, excavated by Bar-Yosef and colleagues (1986), comprised several oval structures and associated courtyards. One of the courtyards yielded deposits of organic material which were interpreted as dung. Only a few broken molar teeth of Capra sp. were identified (Bar-Yosef et al. 1986: 132).

EBII sites: Remains of domestic goat are clearly present in several EB II sites in southern Sinai that were excavated by Itzhak Beit Arieh (summarised in Table 4): Wattiya North, Site 1014, Sheikh Muhsein, Nabi Salah (Beit-Arieh 1977: 142-143), and Feiran I and Sheikh ‘Awad (Beit-Arieh 1981:117, Beit-Arieh 1982:154-155).

Conclusion

Based on this brief review of the published literature, the first conclusive archaeozoological (morphometric) evidence for domestic goats (Capra hircus) in southern Site 1014: This site was not fully excavated. Bones of Sinai is found in the EB II sites, where it was the most goat that were identified in this site (Hakker-Orion common species exploited. Though based on smaller Sites Moyat Daba’iyeh Gunna 25 Gunna 50 samples, the faunal data Species % % % from EB I and Chalcolithic sites in this region also offer Goat/ibex (Capra sp.) 77 9 32 a compelling case for the Gazelle (Gazella sp.) – 2 4 presence of domestic goat Medium sized ruminant – 29 58 herds – a predominance of Equid (Equus sp.) – – 1 remains of Capra, presence Large mammal – – 2 of goat dung and the small Hare (Lepus capensis) 33 9 3 size of the identified goat Jird sp. (Meriones sp.) – 33 – bones, though only two could be measured. Concerning Mastigure (Uromastyx sp.) – 1 – the presence of cattle in the Bird sp. – 1 – Chalcolithic site of Serabit elSand partridge (Ammoperdix heyi) – 0.5 – Khadim (Site 1105), the Ein Quail (Corturnix coturnix) – 8 – Huderah nawamis and the EB Raptor cf. sparrowhawk – 0.4 – II site of Feiran I, the isolated Passerifomes – 7 – nature of these remains and Total Identified NISP 9 361 187 the fact that at least in one instance the find represented Table 3. Identifications of faunal species an artefact, suggests that from the nawamis habitation sites. 138

L. Kolska Horwitz: Animal Offerings from the Nawamis Fields and Coeval Habitation Sites they represent trade items rather than attest to cattle keeping in southern Sinai.

2.5% of the sample. Given that trunk elements have low bone mineral density values (Lyman 1994: Table 7.6), their prominence in this assemblage should be taken to indicate preferential selection of this body part. They have a lower utility index value than most long bones suggesting selective interment of the least meaty skeletal elements. This is corroborated by the high number of cranial remain, similarly poor in meat (unless the brain is taken into consideration).

Game animals such as gazelle, (ibex in the EB II), hare, birds and reptiles, as well as the presence of ostrich eggshell, point to the continuing importance of hunting in these desert economies, corroborated by the high frequencies of arrowheads (Hovers 1981). Despite the presence of marine shells in all sites (Red Sea as well as Mediterranean Sea), no fish remains were found. This may of course be a function of deleterious preservation or bone collection techniques and lack of fine sieving. Small sized, fragile bones of rodents, reptiles and birds are represented in the sites but many, if not all of these, are probably more recent intrusions.

The presence of many arrowheads (Bar-Yosef et al. 1977; Hovers 1981) in addition to gazelle remains in the nawamis assemblage indicates that hunting was still practiced. Likewise, it has been suggested that the high numbers of tabular scrapers in the tombs may be related to meat cutting, animal skinning, and leather processing. These functions could be associated with animal sacrifice and carcass processing (Avner 2002:5960).

Discussion Bar-Yosef et al. (1977: 78) state that at the Ein Huderah nawamis field, ‘practically no animal bones were found’. The small size of the bone assemblages may be attributed to the fact that in so many instances, the tombs had been disturbed in antiquity or collapsed, although the possibility that to begin with few faunal items were introduced as grave offerings, cannot be excluded.

Examination of the faunal record of Sinaitic sites for these periods (Tables 3 and 4) indicates that goats (probably domestic) rather than gazelle contributed the most animal protein to the diet. Thus, gazelle hunted in ‘desert kites’ probably served only as a dietary supplement, perhaps even a delicacy, reserved for special occasions or rituals such as burial rites. Indeed, horned animals, identified as gazelle, ibex, or wild goat, are a motif commonly found in the iconography of Chalcolithic mortuary or cultic contexts (e.g. Milevski 2002; Miroschedji 1993; Shalem 2015). The presence in the nawamis of goat horns as well as those of gazelles, may represent a desert variant of this symbolic behaviour relating to sexuality and fertility, which would imply belief in renewal in a mortuary context such as the nawamis. Symbolically the horns may point to an embodiment of male gods (Shalem 2015) or an association with milking (Milevski 2002). To the latter interpretation may tentatively be added the presence of a fragment of a churn from the ‘Ein Huderah nawamis (Bar-Yosef et al. 1977:80).

The paucity of remains is borne out by data for the nawamis fields presented here, with only 156 animal bones recovered from all tombs excavated. Of these, 106 remains (67.9%) could be identified to taxon or body size (Table 2) but comprised a high proportion of probable intrusive elements (N=34, hare, birds, rodents, reptiles). The remaining 72 ruminant skeletal elements probably represent food offerings intentionally introduced into the tombs with the deceased. A total of 68% of these offerings (N=49 remains) were identified as Capra sp., and as argued above, may be attributed to domestic goats rather than ibex with some confidence. Their number is undoubtedly higher given that their remains may be represented in the additional 20 skeletal parts currently identified as belonging to either goat or gazelle.

Some 40 ‘desert kites’ are known from southern Sinai, many in close proximity to the archaeological sites discussed in this paper (Nadel et al. in press). Consequently, the gazelle found in the nawamis, may have been trapped using desert kites. After all, these installations are thought to have been used specifically for hunting these taxa (e.g. Holzer et al. 2010). Notably, from Kite 14 in Wadi Marrah, the only kite to have yielded fauna during excavation, remains of both dorcas gazelle and wild ass, were recovered. This kite has been dated by radiocarbon to 3750±45 BP (2280-2040 cal BC) (Segal and Carmi 1996:103), an age falling in the EB II. Another Sinai kite in Jebel Hamra (Sinai 10) has yielded three radiocarbon dates of 3300-2900 cal BC (Nadel et al. in press), indicating that use of the Sinai installations may have begun earlier.

Few of the caprine skeletal elements could be aged. Those which could indicate that in the nawamis there was a preference for immature animals. Most of the animals were aged as less than 1 year old (unfused distal scapula), but bones of an adult animal aged at least 3.5 years (if a domestic form, or slightly older if wild) were also found (a fused distal femur). When skeletal elements are summed (Table 4.7), they show that of the 49 identified goat bones, 55% were teeth (fragments) but only 4% were cranial bones/horns and 2% were from lower jaws; 28.5% were vertebrae and only 8% were ribs. Upper forelimbs (humerus and scapula) constituted only 4% of the sample while lower hind limbs were represented by only one bone— 139

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Sites

Site 1014 1

Wattiya North 1

Species

Feiran I 2

Sheikh ‘Awad 3

Wadi Gunna Wadi Gunna 32 4 100 5

Sheikh Muhsein 1

Nabi Salah 1

NISP

Domestic Goat (Capra hircus)

X

X

9

X*





X*

X*

Ibex (Capra ibex)







X





X

X

goat/ibex (Capra sp.)









X

X

Cattle (Bos taurus)





1







X



dorcas gazelle (Gazella dorcas)





1

X







X

Dog (Canis familiaris)





1

X









Hare (Lepus capensis)







X

X



X



Hyrax (Procavia capensis)







X







X

Rodent (Rodentia)













X



Dove/Pigeon (Columba sp.)













X



Ostrich Eggshell (Struthio camelus)

X

X

X

X

X



X

X

Total NISP





15











X = Taxon present but no numerical data given. * most common taxon in assemblage. 1 = Analysis of the fauna from Site 1014 and Wattiya North was carried out by Dr. D. Hakker-Orion and reported in Beit-Arieh (1977). 2 = Analysis of the Feiran I fauna was carried out by Dr. S. Hellwing and reported in Beit-Arieh (1982). 3 = Analysis of the Sheikh ‘Awad fauna was carried out by Dr. S. Hellwing and reported in Beit-Arieh (1981). 4 = Fauna from this site were briefly described in Horwitz and Tchernov (1989a) who analysed the remains. 5 = Data on Gunna 100 are reported in Bar-Yosef et al. (1986: 132).

Table 4. Species list for the EBII Southern Sinai faunal assemblages.

There may even have been a seasonal component to these integrated activities of kite trapping, nawamis construction, secondary burial, and nomadic cycles. As proposed by Finkelstein (1995:19), the nawamis were built by pastoral nomads. Consequently, these installations may have served as community cemeteries, perhaps even burial grounds for tribal confederacies (e.g. Braemer et al. 2001). Bar-Yosef et al. (1977:87) noted that between tombs, ‘differences in the finds reflect the wealth of each group of burials. This variability may indicate burial in family groups and the various clusters of nawamis at ‘Ein Huderah may hint at social ties between families, such as affiliation on the tribal level.’ This idea is supported by the data from the physical anthropology which also suggested they served as family tombs (Hershkovitz et al. 1982).

occupation of this site (and maybe the others) at these times of year, although occupation at other seasons cannot be disproved. Spring or fall occupation is further supported by the suggestion that most nawamis fields were constructed during these seasons (BarYosef et al. 1983; Hershkovitz et al. 1985). Many of the human interments are secondary burials. Therefore, it is possible that people came to the tombs only at a specific time of year, in the spring or fall, to inter their dead. Perhaps, as a communal act, they also repaired or built the desert kites, and hunted gazelle to proffer as burial offerings. Like the nawamis, they represent a major investment in resources to construct and maintain. This activity would have required communal activity (kites would take at least a month to complete; see Nadel et al. in press). Such building projects are an integral part of the transformation of the desert landscape by Chalcolithic-EB desert peoples. It has been suggested that augmented visual signalling, as manifest in aboveground tombs, which stand out in the landscape, is associated with increased pastoral nomadism following aridification and environmental

Evidence in the archaeozoological record for seasonality is the fact that Gunna 25, a nawamis habitation site, has yielded remains of quail and sparrowhawk. Both of these bird species are found in the Sinai only in spring or fall (Paz 1987). Their attestation points to the 140

L. Kolska Horwitz: Animal Offerings from the Nawamis Fields and Coeval Habitation Sites patchiness (Close and Minichillo 2010). The specific location of the nawamis on hilltops and/or next to water sources and routes indicates that they also served as markers, perhaps for territorial boundaries, and/ or water points. Alternately, their location and form may have been related to local cosmologies and belief systems (Cleuziou and Tosi 1997; Close and Minichillo 2010; Harrower 2008; Madsen 1982; Tilley 1994).

cemeteries are clustered in Upper Egypt. However, some examples are also found in Lower Egypt at Maadi, Wadi Digla and Heliopolis (Flores 2003; Van Neer et al. 2004b). In these Egyptian sites, parts of animals were interred with people. Alternatively, whole animals were buried in pits of their own in the same cemetery as the human remains. A wide range of animal species were used for this purpose, including wild taxa. Given that there are several other features that tie the nawamis to Egypt, it is possible that animal interments with humans can be added to this list.

Coeval analogues to the nawamis faunal offerings are sparse in the archaeological record of the southern Levant (Avner and Horwitz 2017). The best comparanda may be found in the Late Neolithic-Early Chalcolithic burial site at Eilat. It offers the closest parallel to the nawamis in terms of geographic location, period, architectural context, material culture as well as fauna, (Avner 1991). At this mortuary site, skeletal elements of sheep/goat were interred with the deceased. As in the nawamis, remains of hare, rodent, and bird were also identified. However, as at the latter sites these may represent intrusive elements rather than grave offerings (Avner and Horwitz 2017). The presence of grave goods in close association with human remains in the nawamis and at Eilat has been interpreted as further evidence for a belief in an after-life (Avner 2002; Goren 1998).

To conclude, in terms of their architecture, the nawamis in southern Sinai appear to reflect a unique desert mortuary tradition. However, it seems to have been influenced by multiple cultures and periods, from Predynastic-Early Dynastic Egypt to Chalcolithic-EB II traditions from the southern Levant. It is suggested that organised hunting of gazelle using ‘desert kites’, in order to use their skulls (and perhaps skins and meat) for grave offerings, may have been an integral part of seasonal mortuary activities associated with the nawamis. The placing of offerings of whole animals or joints of meat inside the tombs, implies a belief in the after-life and offers a tantalizing insight into the spiritual world of these communities.

From as early as the Mousterian, there are examples of intentional interment of faunal remains with human burials from sites in the southern Levant (for a brief summary see Goring-Morris and Horwitz 2007). In most cases the skeletal elements chosen for interment were those with least meat—horns, jaws, skulls, and terminal long bones. As such, they may symbolize pars pro toto. In other instances, they may represent remains of mortuary feasts (Goring-Morris and Horwitz 2007; Horwitz and Goring-Morris 2004; Grosman and Munro 2016). To date, from the Mediterranean region of the southern Levant there are no published examples from the Chalcolithic or Early Bronze Age I periods of faunal remains that have clearly been intentionally interred together with human remains. With a few exceptions (e.g. Nesher-Ramla; see Horwitz 2012), most Chalcolithic examples—Ben Shemen, Sha’ar Ephraim, Nahal Qanah—are problematic. They appear to represent random associations following disturbance to tombs or carnivore activity (Davis 1980; Smith and Horwitz 1998). Likewise, a partial gazelle skeleton found at the entrance, but not inside, an Early Bronze Age tomb at Bab edh-Dhra (Hesse and Wapnish 1975/78), may also represent an accidental association.

Acknowledgements I am grateful to the Israel Antiquities Authority for facilitating this study, which was carried out as part of their research program on archaeological remains recovered during the Israeli excavations in Sinai; to Avner Goren and Ofer Bar-Yosef, the excavators of the nawamis for permission to study these assemblages; and to Uzi Avner for the photographs used here. References Albright, W. F. 1948a. Exploring in Sinai with the University of California African Expedition. Bulletin of the American Schools of Oriental Research 109: 5-20. Albright, W. F. 1948b. The early alphabetic inscriptions from Sinai and their decipherment. Bulletin of the American Schools of Oriental Research 110: 6-22. Arad-Ayalon, N. n.d. Finds From the Nawamis Sites in Sinai. Unpublished Israel Antiquities Authority Monograph. Jerusalem. Avner, U. 1991. Late Neolithic-Early Chalcolithic Burial Site in Eilat. American Journal of Archaeology 95: 496497. Avner, U. 1993. Mazzebot sites in the Negev and Sinai and their Significance. In A. Biran and J. Aviram (eds), Biblical Archaeology Today, 1990: 166-181. Jerusalem: Israel Exploration Society. Avner, U. 1998. Settlement, Agriculture and Paleoclimate in ‘Uvda Valley, Southern Negev Desert, 6th-3rd millennia B.C. In A. Issar and N. Brown (eds), Water,

In contrast, the Predynastic and Early Dynastic periods in Egypt have yielded a wealth of human-animal interments (e.g. Boessneck et al. 1992; Flores 2003; Ikram 2013). The Egyptian practice of funerary sacrifice of animals is considered Nubian in origin (Van Neer et al. 2004b). As such, it is not surprising that most of these 141

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Environment and Society in Times of Climate Change: 147-202. Amsterdam, Kluwer. Avner, U. 2002. Studies in the Material and Spiritual Culture of the Negev and Sinai Populations During the 6th-3rd millennia B.C. Unpublished PhD thesis, The Hebrew University of Jerusalem. Avner, U. and Horwitz, L.K. 2017. Sacrifices and Offerings from Cult and Mortuary Sites in the Negev and Sinai, 6th-3rd millennia BC. ARAM 29 (1-2): 35-70. Bar-Yosef, O., Belfer-Cohen, A., Goren, A. and Smith, P. 1977. The Nawamis near ‘Ein Huderah (eastern Sinai). Israel Exploration Journal 27 (2-3): 65-88. Bar-Yosef, O., Hershkovitz, I., Arbel, G. and Goren, A. 1983. The Orientation of Nawamis Entrances in Southern Sinai: Expressions of Religious Belief and Seasonality? Tel Aviv 10 (1): 52-60. Bar-Yosef, O., Belfer-Cohen, A., Goren, A., Hershkovitz, I., Ilan, O., Mienis, H. and Sass, B. 1986. Nawamis and Habitation Sites Near Gebel Gunna, Southern Sinai. Israel Exploration Journal 36 (3-4): 121-167. Bar-Yosef Mayer, D. E. 1999. The Role of Shells in the Reconstruction of Socio-Economic Aspects of Neolithic Through Early Bronze Age Societies in Sinai. Unpublished PhD thesis, The Hebrew University of Jerusalem. Bar-Yosef Mayer, D. E. 2002a. Egyptian-Canaanite interaction during the Fourth and Third millennia BCE: The shell connection. In E. C. M. van den Brink and T. E. Levy (eds), Egypt and the Levant. Interrelations from the 4th through the Early 3rd Millennium B.C.E.: 12135. London, Leicester University Press. Bar-Yosef Mayer, D. E. 2002b. The Shells of the Nawamis in Southern Sinai. In H. Buitenhuis, A. M. Choyke, M. Mashkour and A. H. Al-Shiyab (eds), Archaeozoology of the Near East vol. V: 166-180. Groningen: Archaeological Research and Consultancy Publicaties 62. Beit-Arieh, I. 1977. Southern Sinai in the Early Bronze Age Period. Unpublished PhD thesis, Tel Aviv University. (Hebrew). Beit-Arieh, I. 1981. An Early Bronze II Site near Sheikh ‘Awad in Southern Sinai. Tel Aviv 8: 95-127. Beit-Arieh, I. 1982. An Early Bronze II Site near the Feiran Oasis in Southern Sinai. Tel Aviv 9: 146-156. Beit-Arieh, I. 2003a. Archaeology of Sinai. The Ophir Expedition. Emery and Claire Yass Publications in Archaeology. Tel Aviv. Beit-Arieh, I. 2003b. Excavated sites in south central Sinai. VII. A Chalcolithic site. In I. Beit-Arieh (ed.), Archaeology of Sinai. The Ophir Expedition: 78-100. Emery and Claire Yass Publications in Archaeology 21. Tel Aviv, Monograph Series of The Sonia And Marco Nadler Institute Of Archaeology, Tel Aviv University. Boessneck, J., Driesch, A., von den and Eissa, A. 1992. Eine Eselsbestattung der 1. Dynastie in Abusir. Mitteilungen des Deutschen Archäologischen Instituts Kairo 48, 1-10. Braemer, F., Steimer-Herbet, T., Buchet, L., Saliège, J.F. and Guy, H. 2001. Le Bronze ancien du Ramlat

as-Sabatayn (Yémen). Deux nécropoles de la première moitié du IIIe millénaire à la bordure du désert: Jebel Jidran et Jebel Ruwaiq. Paléorient 27, 21-44. Carmi, I. and Segal, D. 1994. Radiocarbon dates from sites excavated in Sinai. Unpublished Report No. 42. Radiocarbon Lab., Weizmann Institute of Science. Cione, A. L. 2006. Fishes from  Tell el-Ghaba. In P. Fuscaldo and S. Lupo (eds), Tell el-Ghaba, a Saite Settlement in North Sinai, Egypt. Argentine Archaeological Mission 1995-2004, Vol II: Studies. Buenos Aires, Consejo Nacional de Investigaciones Científicas y Técnicas. Cleuziou, S. and Tosi, M. 1997. Hommes, climats et environnements de la Peninsule arabique a l’Holocène. Paleorient 23 (2) 121-35121-35. Close, A. E. and Minichillo, T. 2010. Neolithic Tombs in Southwestern Sinai. Journal of Arid Environments 74 (7): 829-841. Currelly, C. T. 1906. Gebel Musa and the Nawamis. In W. F. Flinders Petrie (ed.), Researches in Sinai: 229-244. London: John Murray. Davis, S. J. 1980. Les ossements d’animaux. In: J. Perrot and D. Ladiray (eds), Le cimetiére de Ben Shemen. Tombes a ossuaires de la region côtiére Palestinienne au IVe millenaire avant l’ere Chretienne: 94 [x2]. Memoires et Travaux du Centre de Recherches Préhistoriques Français de Jérusalem Jérusalem, No. 1. Paris, Association Paléorient. Driesch, A. von den 1976. A Guide to a Measurement of Animal Bones from Archaeological Sites. Cambridge: Peabody Museum Bulletin 1, Peabody Museum of Archaeology and Ethnology. Flores, D. V. 2003. Funerary Sacrifice of Animals in the Egyptian Predynastic Period. British Archaeological Reports International Series 1153. Oxford, Archaeopress. Goren, A. 1980. The Nawamis in Southern Sinai. In Z. Meshel and I. Finkelstein (eds), Sinai in Antiquity. Tel Aviv, Hakkibutz Hameuchad Publishing House: 243263. (Hebrew). Goren, A. 1998. The Nawamis in southern Sinai. In S. Ahituv (ed.), Studies in the Archaeology of Nomads in the Negev and Sinai: 59-85. Beer Sheva: Ben Gurion University of the Negev Press (Hebrew). Goring-Morris, A. N. and Horwitz, L. K. 2004. Animals and ritual during the Levantine Pre-Pottery Neolithic B: A Case Study from the Site of Kfar Hahoresh, Israel. Anthropozoologia 39 (1): 165-178. Goring-Morris, A. N. and Horwitz, L. K. 2007. Funerals and Feasts during the Pre-Pottery Neolithic B of the Near East. Antiquity 81: 902-919. Grigson, C. in press. Preliminary Report on the Mammal Bones from Chalcolithic Qatif, Site Y2 (including Y2a), on the Sinai Coastal Plain (Excavations of 1979, 1980 and 1983). In G. Bar-Oz and L. K. Horwitz, (eds), Discovering Noah’s Ark: Zooarchaeology of the Holyland. 142

L. Kolska Horwitz: Animal Offerings from the Nawamis Fields and Coeval Habitation Sites Israel Antiquity Authority Monographs. Jerusalem, Israel Antiquities Authority. Grosman, L. and Munro, N. D. 2016. A Natufian Ritual Event. Current Anthropology 57 (3). Harrower, M. 2008. Hydrology, Ideology, and the Origins of Irrigation in Ancient Southwest Arabia (Yemen). Current Anthropology 49(3): 497-510. Hershkovitz, I. 1987. Trephination: The Earliest Case in the Middle East. Mitekufat Haeven: Journal of the Israel Prehistoric Society 20:128-135. Hershkovitz, I., Kobyliansky, E. and Arensburg, B. 1982. Coxa Vara in a Chalcolithic Population from the Sinai. Current Anthropology 23:320-322. Hershkovitz, I., Arbel, G., Bar-Yosef, O. and Goren, A. 1985. The Relationship between Nawamis Entrance Orientations and Sunset Direction. Tel Aviv 12:204211. Hesse, B. and Wapnish, P. 1975/78. The Gazelle from a Shaft Tomb at Bab edh-Dhra. Bulletin of the American Schools of Oriental Research 12: 4-15. Hilzheimer H. K. 1961. Die Altersbestimmung bei Haustieren, Peltztieren und beim jagdbaren Wild. Berlin: Paul Parey. Holzer A., Avner U., Porat, N. and Horwitz, L. K. 2010. Desert kites in the Negev desert and North East Sinai: Their function, chronology and ecology. Journal of Arid Environments (Special Issue in honour of Aharon Horowitz) 74/7: 806-817. Holland, F. W. 1870. Sinai and Jerusalem. London, Society for Promoting Christian Knowledge. Horwitz, L. K. 2003a. Early Bronze Age Archaeozoological remains. In I. Beit-Arieh (ed.), Archaeology of Sinai. The Ophir Expedition: 242-257. Tel Aviv: Emery and Claire Yass Publications in Archaeology. Tel Aviv, Monograph Series Of The Sonia And Marco Nadler Institute Of Archaeology, Tel Aviv University. Horwitz, L. K. 2012. Faunal remains. In V. W. Avrutis (ed.), Late Chalcolithic and Early Bronze Age I Remains at Nesher-Ramla Quarry. Haifa, Zinman Institute of Archaeology, University of Haifa: 239-250. Horwitz, L. K. and Tchernov, E. 1989. Animal exploitation in the Early Bronze Age of the Southern Levant: An Overview. In P. de Miroschedji (ed.), L’urbanisation de la Palestine a l’age du bronze ancient: 179-296. British Archaeological Reports International Series 527(ii). Oxford, British Archaeological Reports. Horwitz, L. K. and Tchernov, E. 1990. Cultural and Environmental Implications of Hippopotamus Bone Remains in Archaeological Contexts in the Levant. Bulletin of the Schools of Oriental Research 280: 67-76. Hovers, E. 1981. The Assemblages from Nawamis and Habitation Sites of the Fourth Millennium in Sinai. Lithics of Ein Huderah, Ein um Ahmed and Wadi Nasb. Unpublished MA Seminar Paper, The Hebrew University of Jerusalem. (Hebrew). Hume, W. F. 1906. The Topography and Geology of the Sinai (South-Eastern Portion). Cairo, National Printing Department.

Ikram, S. 2013. Man’s Best Friend for Eternity: Dog and Human Burials in Ancient Egypt. Anthropozoologica 48 (2): 299-307. Ilan, O. and Sebbane, M. 1989. Copper Metallurgy, Trade and the Urbanization of Southern Canaan in the Chalcolithic and Early Bronze Age. In P de Miroschedji (ed.), L’urbanisation de la Palestine a l’age du Bronze ancien: 139-162. British Archaeological Reports International Series, 527(i). Oxford, British Archaeological Reports. Levy, T. E. 1995. Cult, Metallurgy and Rank Societies – Chalcolithic period (ca. 4500-3500 BCE). In T. E. Levy (ed.), The Archaeology of Society in the Holy Land: 226244. London, Leicester University Press. Lyman, R. L. 1994. Vertebrate Taphonomy. Cambridge, Cambridge University Press. Madsen, T. 1982. Settlement Systems of Early Agricultural Societies in East Jutland, Denmark: A Regional Study of Change. Journal of Anthropological Archaeology 1:197-236. Mayor, A. 2000. The First Fossil Hunters: Paleontology in Greek and Roman Times. Princeton, Princeton University Press. McCorriston, J., Steimer-Herbet, T., Harrower, M., Williams, K., Saliège, J-F. and Bin ‘Aqil, A. 2011. Gazetteer of Small-scale Monuments in Prehistoric Hadramawt, Yemen: A Radiocarbon Chronology from the Roots of Agriculture in Southern Arabia— Ancient Human Social Dynamics Project Research 1996-2008. Arabian Archaeology and Epigraphy 22: 1-22. Mendelssohn, H. and Yom-Tov, Y. 1999. Fauna Palaestina – Mammalia of Israel. Jerusalem, Israel Academy of Sciences and Humanities. Milevski, I. 2002. A New Fertility Figurine and New Animal Motifs From The Chalcolithic in the Southern Levant: Finds From Cave K-1 at Quleh, Israel. Paléorient 28 (2): 133-142. de Miroschedji, P. 1993. Cult and Religion in the Chalcolithic and Early Bronze Age. In A. Biran and J. Aviram (eds), Biblical Archaeology Today, 1990: 208220. Pre-Congress Symposium: Population, Production and Power. Jerusalem: Israel Exploration Society. de Miroschedji, P., Sadeq, M., Faltings, D., Boulez, V., Nagghar-Moliner, L., Sykes, N. and Tengberg, M. 2001. Les fouilles de Tell es-Sakan (Gaza): Nouvelles données sur les contacts Egypto-Cananéens au IVeIIIe millénaires. Paléorient 27 (2): 75-104. Murray, G. W. 1935. Bee-Hive Graves (Nawamis) in the North-Eastern Sudan and Sinai. Man 35: 17-18. Nadel, D., Malkinson, D., Bar-Oz, G., Nachmias, A., Avner, U., Horwitz, L. K. and Porat, N. in press. Small, Sparse and Effective: The Negev and Sinai Kites. In A. Betts and W. P. Van Pelt (eds), The Gazelle’s Dream. Game Drives of the Old and New Worlds. University of Sydney Press (USCAP). Noddle, B. 1974. Ages of epiphyseal closure in feral and domestic goats and ages of dental eruption. Journal of Archaeological Science 1: 195-204. 143

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Palmer, E. H. 1871. The Desert of the Exodus. Cambridge, Deighton, Bell and Co. Paz, U. 1987. The Birds of Israel. Tel Aviv, Ministry of Defence Publishing House. Perrot, J. 1959. Statuettes en ivoire objects en ivoire et en os provenant des gisements préhistoriques de la region de Beersheba. Syria 36: 8-19. Reese, D. S., Mienis, H. K. and Woodward, F. R. 1986. On the Trade of Shells and Fish from the Nile River. Bulletin of the American Schools of Oriental Research 264: 79-84. Rothenberg, B. 1972. Sinai explorations 1967-1972. Museum Haaretz Bulletin: 37-45. Rothenberg, B. 1975. Two excavations in Sinai. Museum Haaretz Bulletin 17/18: 36-42. (Hebrew). Rothenberg, B. 1979. Sinai: Pharaohs, Miners, Pilgrims. Berne, Kümmerly and Fry. Rowan, Y. M., Rollefson, G. O. and Kersel, M. M. 2011. Maitland’s ‘Mesa’ Reassessed: A Late Prehistoric Cemetery in the Eastern Badia, Jordan. Antiquity (Gallery) http://www.antiquity.ac.uk/projgall/rowan327/ Segal, D. and Carmi, I. 1996. Rehovot Radiocarbon Date List V. Atiqot 29: 79-106. Shalem, D. 2015. Motifs on the Nahal Mishmar Hoard and the Ossuaries: Comparative Observations and Interpretations. Journal of the Israel Prehistoric Society 45: 217-237.

Silver, I. A. 1969. The Ageing of Domestic Animals. In D. R. Brothwell and E. S. Higgs (eds), Science in Archaeology. London: Thames and Hudson. Pp. 283302. Smith, P. and Horwitz, L. K. 1998. Human and Animal Remains from the Burial Cave at Sha’ar Ephraim Central. Tel Aviv 25 (1): 110-115. Stager, L. E. 1992. The Periodization of Palestine from Neolithic through Early Bronze Times. In R. W. Ehrich (ed.), Chronologies in Old World Archaeology, vol. I, 22-41. Chicago, University of Chicago Press. Tilley, C. 1994. A Phenomenology of Landscape: Places, Paths, and Monuments. Oxford, Berg. Van Neer, W., Linseele, V. and Friedman, R. 2004a. Animal Burials and Food Offerings at the Elite Cemetery HK6 of Hierakonpolis. In S. Hendricks, R. F. Friedman, K. M. Cialowicz and M. Chlodnicki (eds), Egypt at its Origins: 67-130. Leuven, Peeters. Van Neer, W., Lernau, O., Friedman, R., Mumford, G., Poblome, J. and Waelkens, M. 2004b. Fish Remains from Archaeological Sites as Indicators of Former Trade Connections in the Eastern Mediterranean. Paléorient 30: 101-148. Wilson, C. W. 1871. The Recovery of Jerusalem: A Narrative of Exploration and Discovery in the City and the Holy Land. London, R. Bentley.

144

An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila Hamoudi Khalaily, Anastasia Shapiro and Ofer Marder Introduction

and 1.5 kilometers southeast of Tel Nagila (Fig. 1.) The topography of the surrounding region is dominated by very low loessial hills (Rosen 1991: 192) with an average elevation of 180 m ASL. Ephemeral streams have incised the topography with shallow streams running toward Nahal Shiqma. Intensive modern farming has caused changes in the natural landscape and obscured the natural vegetation. The archaeological remains were recovered on a moderate slope on the eastern bank of Nahal Shiqma. At this point, the valley becomes narrow and steep and its meander truncates deeply the alluvial deposits exposing part of the geological sequence in this region. The quaternary alluvial deposition (Rosen 1986) is overlaid by the Ahuzam formation conglomerate, and Pliocene marine deposits of the Pleshet formation below them (Bochbinder 1969; Sneh and Bochbinder 1984). Flint pebbles and cobbles—derived from the Ahuzam formation, which considered to be a reworking of the marine source of the Pleshet formation, as well as Senonian—originated from uphill, where the main source of the lithic industry was located. The site and it surroundings suffered additional disturbance when the national petroleum pipeline was dug nearby, causing severe damage to the archaeological remains.

The Pottery Neolithic period in the inland Northern Negev is still poorly known, due to the lack of comprehensive survey and to the dearth of deep exposure in the few major tell excavations. In contrast, the southern Coastal Plain was intensively studied during the late 1960s, when several Pottery Neolithic sites were discovered and later excavated, such as Nizzanim (Yeivin and Olami 1979), Ziqim (Noy 1977; Garfinkel et al. 2002) and Qatif Y-3 (Epstein 1984). In recent years, the publication of three comprehensive survey maps (Ziqim [Berman, Stark and Barda 2004]; Nizzanim East and west [Berman and Barda 2005) has provided vital information about the Pottery Neolithic. Evidence for the occupation of inland sites in the Early Pottery Neolithic (Hencforth EPN)-Jericho IX culture has been provided by the recent publication of Tel Lachish (Gophna and Blockman 2004: 873-899: Rosen 1988-1989) and the excavation of Horvat Ptora, east of Kiryat Gat (Milevski and Baumgarten 2008). These new data indicate that the Jericho IX culture appears in the Northern Negev during the early six millennium cal. BCE. In multi-period sites the strata from this period are likely to be buried and damaged by later occupation (Gilead 1990:47).

The Excavations

In the fall of 2007, a salvage excavation was conducted c. 1.5 km southeast of Tel Nagila, as a result of damage caused by a trench cut for the Israeli national gas pipeline.1 The section of this trench revealed several ash pits, while many sherds were found scattered over an area of approximately 5000 m2. Due to the proximity of this site to Tel Nagila and because it adds an important facet to the site’s environs, it was decided to include this report in the present volume.

The salvage excavation aimed to uncover the archaeological remains in an area of 80 x 10 m. A total of 10 squares (5 x 5 m) were located adjacent to the points damaged by the pipeline, running through the area from north to south (Fig. 2). Two squares were located at the southern end, three in the center and another three at the northern part of the area. An additional two squares were located where archaeological remains are visible directly on the surface. One archaeological layer was exposed and subdivided in two levels (Fig. 3).

Description of the Site

The upper level is anthropogenic, consisting of clayey soil (loess), with a brownish-gray color. The thickness of this level is unknown, the topsoil was removed during the flattening of the area, in preparation for digging. What remained, however, was approximately 40 cm deep. At the top of the layer, it was possible to identify remains of at least two walls associated with a thick cobble layer. At the base, outlines of pits cutting into the lowest layer of sterile loess were visible. Some

The site is located in the southern Coastal Plain, approximately 10 kilometers south-west of Kiryat Gat The excavation (permit A-4974/2006; map ref. NIG 177150-600680177300-600850) was conducted by H. Khalaily and O. Marder on behalf of the Israel Antiquities Authority and sponsored by the National Gas Authority. Additional assistance was provided by I. Milevski, G. Mazor and I. Peretz; as well as S. Lander (administration), A. Hajian (surveying and drafting), M. Smilansky (flint drawings) and I. Lidseki (pottery drawings). 1 

145

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 1. Location map.

Figure 2. Plan of excavation areas.

146

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 3. Main section showing the main archaeological deposits.

Figure 4. Archaeological layer overlaid the sterile loess.

The majority contained molded mud bricks in various shapes and sizes, along with brick material, a few pottery fragments, flint and flaked stones such as those in Feature 102 (Fig. 6).

of the pits were rounded and others elliptical. All of the features contained ceramic sherds, stone vessels and flint artifacts. This layer sealed a thick sterile loose deposit, light brown in color and rich in carbonate nodules (Fig. 4).

Four elliptic pits were discovered (Loci 203, 206, 207, 302), two of which (Loc. 301 and 307) were fully excavated, while two (Loci 203 and 302) were only partially dug. They have large, asymmetric apertures and wide, flat bases. They were dug into the sterile loess to a depth of ca. 0.5 m. However, their upper portions were truncated, indicating that they were originally deeper. The pits were filled with loose, dark brown sediment rich in archeological remains, including pottery fragments, flint artifacts and grinding stones (Fig. 7).

Two types of features were exposed: pits and a living surface. The pits vary in shape and size, with circular or elliptical outlines, an average diameter of 1 m. These penetrate the virgin soil to an average depth of 0.5 m. They were scattered over the entire area. Only two pits (Loc. 206, 207) intersected each other, indicating that they were dug in different construction phases. Four circular pits were excavated (Loc. 102, 206, 301 and 307). They have U-shape profiles and straight walls, and are wide and flat at their base (Fig. 5). They are up to 0.5 m deep. The pits were filled with loose grey sediment rich in organic material, yet lacking artifacts.

In the central excavation area (Area B), in Squares X6-7, a living surface was exposed. The living surface extends over 20 m2 and is 0.2 m thick. It is composed 147

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 5. Plan of the excavation squares and the main loci.

of a dense layer of flat, rounded cobbles and pebbles originating from the nearby Nahal Shiqma. It is possible to distinguish two constructional phases (Fig. 8): The uppermost phase is horizontal and composed of flat, well sorted pebbles, yielding large quantities of archaeological material. The lower phase is thinner than the upper one, consisting of dense angular cobbles yielding scarce archaeological remains.

of large field stones and small angular stones between them. Wall 209 displays the same characteristics. However, the upper stones most likely collapsed, making its shape unclear. It is clear that this wall fragment was constructed with the upper phase of the living surface. It is most probable that the wall was part of a large structure and functioned as an enclosing or partition wall. If this interpretation is correct, this is a rare occurrence of architecture at an EPN site in the south.

Two fragments of possible walls were identified in the uppermost phase. The two fragments could, in fact, be one long wall that was cut by the modern trench. Both segments were distinguished by their shape and the large angular stones used to build them. Wall 208 (Fig. 9, 1 m long and 0.4 wide) is constructed of one row

The Lithic Implements The flint assemblage consists of 1249 items (Table 1). Most of the finds are waste products, with tools 148

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 6. Pit 101 in Sq. X-3.

Figure 7. Photo of pit 102 with molded mud-bricks.

149

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 8. Plan of Sq. Y-9 including L 206.

appearing in low frequency. Débitage represents ca. half of the flint assemblage, followed by debris (40%), tools (6.4%) and cores (5.6%). Flakes are the predominant waste product (70.8% of the débitage). Blades and bladelets represent 8.3% of the débitage, reflecting the secondary use of blanks in the production of tools (see below). It appears that tools were manufactured locally, as all production stages were identified in the assemblage. Therefore, the large amount of flake blanks

may reflect a primary industry oriented towards flake production. Approximately, 80% of the assemblage was collected from Area B, from features and living surfaces. 13.5% of the artifacts were found in Area A, mostly deriving from two shallow pits. The rest of the items derived from Area C (6.5% of the assemblage), from a living surface and two bell-shaped pits. Clearly, there is a marked 150

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 9. Two phases of Living surfaces in Area B.

Area A

Area B

Area C

Total

%

Debitage Primary Elements

21

76

16

113

18.9

Flakes

69

324

31

424

70.8

Blades

6

24

3

33

5.5

Bladeles

1

16

17

2.8

C.T.E

2

6

4

12

2.0

Total

99

446

54

599

100

Debris Chunks

8

63

6

77

15.4

Chips

33

379

11

423

74.6

Total

41

442

17

500

100

Debitage

99

446

54

599

48.0

Debris

41

442

17

500

40.0

Cores

14

48

8

70

5.6

Tools

15

54

11

80

6.4

Total

169

990

90

1249

100

distinction in the relative distribution of the waste products and tools among the feature types. The large amount of waste products collected from the cobble surface and related features in Area B, in comparison to the finds from Area A and C, may indicate that most of the knapping activity took place within the domestic quarters. In contrast, the similar type frequencies observed in the various features indicate that these features were not directly connected to the production of flint tools and their content was not influenced by this industry. The presence of flint artifacts (waste and tools) in these features may be interpreted in terms of a

Table 1. Breakdown of Flint Artifacts According to Type and Area of Excavation.

general waste disposal, not the product of a specialized workshop. Various types of flint raw material were used at the site: grey flint (79%), brown flint (12%), translucent chalcedony (5%), and light brown flint (4%). The majority of the flint originated from flint pebbles from the Ahuzam conglomerate, readily available in the Negev wadis. Brown Senonian chalcedony flint probably originated from the hills surrounding the site. 151

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Cores

were used as the primary striking platforms. A series of elongated flakes were produced to remove the cortex and an additional series of small flakes were struck in the second production stage. At times, an additional striking platform was prepared in the process, producing small flakes (Fig. 11:2-3). After two to three series of utilization, the core lost its volume and shape and was discarded. The secondary reduction sequence noted was aimed at blade production (Fig. 10:1, 4). The blade was struck from an elongated pebble with a single platform. The flaking surface is flat and usually located on the narrow side. It is important to note that the core exploitation was limited to this flaking surface and there was no core rejuvenation; it was discarded after two production series. The blade blanks were usually wide and thick, and most were broken in the process of production. Many retained a cortex on the lateral side.

Seventy cores were recovered from the excavated areas, more than half collected in Area B. The majority were originally small pebbles, flat on both sides and exhibiting a thin whitish cortex. The pebbles were tested and knapped at the site, as reflected by the high amount of primary elements within the débitage. The cores were intensively exploited on one side, leaving a large surface covered with cortex. Most of the primary pebbles originated from the Mishash formation, yielding raw material of average quality. The cores were utilized until they lost their shape, making the amorphous core the most common in this assemblage. As mentioned above, the majority of cores were shaped from local grey and brown flint and only two were of translucent flint. Translucent and brown cores were used mainly for the production of bladelets. The most common method of core reduction is simple flake production from one striking platform (Fig. 10). The medium-sized flint pebbles were split in two using the block-on-block technique. The breakage surfaces

Tools The tool assemblage consists of 80 tools, two thirds of which were found in Area B. The assemblage is clearly dominated by ad hoc tools, with fewer formal tools.

Figure 10. Walls 2007 and 208 in Sq. X-6, 7.

152

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 11. Cores. 1-4 single platform cores.

Arrowheads

Flakes were the preferred blanks for tool manufacture while blades were used exclusively for sickle and retouch blades. One of the sickle blades however, was clearly shaped on either a thick flake or even an elongated pebble. It is of interest to note the absence of bifacial tools, other than a single item that may be classified under this category.

The two arrowheads are Ha-Parsa points, with small wings and thin tangs (Fig. 12:1-2). Both exhibit pressure retouch on both sides. One is broken and missing one of the wings. It is wide and the pointed tang forms almost half of it. It seems that the initial blank for this item 153

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead was a flake, although the pressure retouch on both sides precludes identification of the original shape. These two arrowheads are of a group of small arrowheads that are characteristic of Levantine EPN assemblages (Gopher 1994) and are most common in the Jericho IX culture (Khalaily 1999).

Similar sickle implements can be found in EPN assemblages of the Jericho IX-Lodian culture (Garstang, Ben-Dor and Fitzgerald 1936; Gopher and Blockman 2004), although smaller quantities can also be found in Yarmukian assemblages (Gopher 1994; Garfinkel and Miller 2002; Khalaily 2006). These are the dominating sickles at Jericho (Crowfoot-Payne 1983), Hagoshrim (Khalaily 1999), Lod (Gopher and Blockman 2004), Givat ha-Parsa (Olami, Burian and Friedman 1977) and Nizzanim (Yievin and Olami 1979). This type was also found in the Neolithic layers at Lachish (Gophna and Blockman 2004) and Horvat Petora (Milevski and Baumgarten 2008).

Sickle blades Only four sickle blades were identified, comprising 5% of the tool assemblage. All belong to Munhatta Type B (Gopher 1989). They are wide, rectangular blades which display full pressure retouch on both faces. Both extremities are truncated and their back is shaped by bifacial steep retouch, while their working edges have coarse denticulation shaped by pressure retouch and exhibit pronounced sickle gloss (Fig. 12:3-4).

Ad hoc tools Notches and denticulates (e.g Fig. 13:2) clearly dominate the ad hoc tools, consisting of one third of the tool assemblage and almost 40% of the ad hoc tools. Perforators are the next most frequent, consisting of awls (Fig. 13: 1, 4) and borers, including two massive borers made on pebbles with rounded butts. The points were shaped by two longitudinal removals along the edges and abrupt sharpening at the tips (Fig. 13:3). Retouched flakes, retouched blades and burins are present in small quantities. Most of the scrapers are end scrapers, formed on flakes other than two fan scrapers (Fig. 12:5-6) which were shaped on tabular flint. The latter items are made on brown, fine-grained, flat flint that is not available in the vicinity of the site. Usually, the blanks are completely covered by cortex on their dorsal face. Retouch is usually semi-abrupt. The Pottery Due to the fragmented and incomplete nature of the pottery assemblage retrieved from the three areas, none of the vessels could be restored. The assemblage as a whole is very small and includes 1087 sherds. Diagnostic sherds of vessels include 54 rim fragments of open vessels (52 bowls, 2 basins). Closed vessels include 17 rim fragments of medium-sized holemouth jars and five rims of jars with short, everted rims. In addition, there are 30 handles – mostly fragments, 15 fragments of flat bases and 966 body sherds, 100 of which are decorated by red slip. Technologically, the pottery is friable, handmade and locally manufactured. The clay is composed mainly of silt collected from deposits in the nearby wadis. It was well sorted and tempered with small whitish grits mixed with a considerable amount of straw, similar to the Qatifian pottery (Goren 1990). The cores are dark and the surfaces are light grey and brown, washed and sometimes smoothed, and often decorated by red slip. The main pottery types are presented into two groups; open and closed vessels. Open vessels comprise

Figure 12. Cores. 1 Amorphous core; 2-3 two platform cores.

154

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 13. Tools. 1-2 Arrowheads; 3-4 Sickle blades; 5-6 fan-scrapers.

Open Vessels

the majority (c. 79%) of the assemblage and include mainly bowls in a variety of shapes and sizes, while the closed vessels include mainly holemouth jars and high-necked jars.

One cup and four types of bowls were recognized: small shallow bowls, large shallow bowls, hemispherical bowls and deep open bowls with straight walls. 155

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Cup (Fig. 14:1) The only cup found (Fig. 14:1), is a miniature vessel with a diameter of 2.2 cm. Such miniature cups are common in pottery assemblages of this culture, for example at Jericho (Garfinkel 1999:32, type C4) and Lod (Gopher and Blockman 2004:19).

towards the base. Due to their poor preservation, it was impossible to estimate their depth, although the opening in such vessels is usually wider than the depth. This type is commonly red slipped externally or on both the interior and the exterior (e.g. Fig. 14: 2-3). A pair of knob handles is visible on two examples.

Small shallow bowls (Fig. 14:2-4)

Large shallow bowls

Small shallow bowls are common in Jericho IX assemblages. Eighteen such vessels were found in the excavations. Their walls are narrow and straight near the rims, becoming slightly rounded and thick

Only two rims can be classified as shallow bowls with straight walls and simple rims (Fig. 14:14-15). Similar vessels are found in the Yarmukian and Jericho IX cultures (Garfinkel’s Type C6 – 1999:79) and in the

Figure 14. Tools. 1, 4 Awls; 2 denticulation; 3 massive borer; 4 bifacial in preparation.

156

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila Jericho IX/Lodian assemblages of Lod (Gopher and Blockman 2004:17, Fig. 5).

sharply beveled rims. Only one vessel was decorated with a thin band of red slip which is visible on the inner rim and outer surfaces. The rest of the vessels are large, with crude walls and minimal surface treatment. Only one (Fig. 15: 4) was decorated with a pale surface wash and red slip.

Hemispherical bowls (Fig. 13:4-7) Nine vessels were classified as hemispherical bowls. They have rounded walls and are small in diameter (less than 13 cm according to Blockman (1997:12). Their surfaces were well-treated, with almost half being red-slipped. The rims are rounded or sharply beveled. Garfinkel (1999) classifies these bowls as Type C7; they are present in the Lodian assemblages (Gopher and Blockman 2004: 19, Fig. 7).

High-necked Jars Two everted rims of necked jars were found (Fig. 15:67), probably belonging to high-necked jars. The first example (Fig. 15:6) is much taller and the rim slightly everted, while the second is shorter and more everted. While such jars are present in the Jericho IX/Lodian repertoire, they are less common than the dominant type of converging-neck jars (Gopher and Blockman 2004: 10).

Deep open bowls (Fig. 14:8-13) This type is the most frequent in the Nagila East assemblage and consists of 28 vessel fragments. The walls are straight and outwardly inclined. These bowls are larger, with diameters ranging from 12-16 cm and a depth of up to 12 cm. No bases of these vessels could be discerned with certainty however, it is most likely that they were flat, similar to Fig. 15:3. The surfaces are well treated, with red slip common (e.g. Fig. 14:9, 11), although there is no sign of burnishing. This type, labeled Type C1 by Garfinkel (1999:32), is very common in Yarmukian and Jericho IX assemblages. Similar vessels appear in large quantities in the Lodian assemblages, where some are defined as kraters (Gopher and Blockman 2004: Fig. 8:5-14).

Handles Thirty handles were recovered: 20 knob handles (67%), seven loop handles and three lug handles. The majority of the loop handles were found broken. Only one is complete (Fig. 16:4). Judging from its size, it is possible to assume that this loop handle was attached to the shoulder of a jar. Other handles (Fig. 16:5) have a round section and they are decorated with red slip, possibly belonging to a decorated jar. The knob handles (e.g. Fig. 16:1-3) likely belonged to decorated jars, attached either on the shoulders or near the rims. The wider knobs may also belong to decorated deep bowls and pots. One handle was definitively a lug handle (Fig. 16:1), possibly attached to the interior or exterior of a basin.

Basins Basins are rare in this assemblage, with only two rim fragments retrieved.

Bases

Closed Vessels

Fifteen bases were counted, 14 of which were flat thick bases, most likely of jars. Some are just flat slabs that are crudely connected to the walls of the vessel with no further modification (Fig. 16:8). There are also flat bases that protruded slightly outward (Fig. 16:7, 9). One base (Fig. 16:6) is a high cylindrical ring base, decorated by red slip and signs of burnishing, which likely continued on the upper part of the vessel. The base belongs to either a chalice or an amphoriskos.

Two types of vessels belonging to this group were defined: holemouth jars (89% of the rims) and highnecked jars (comprising the rest). Holemouth jars (Fig. 15:1-5) Holemouth jars are the most common type in EPN assemblages. It is the dominant form in both the Yarmukian and the Lodian assemblages (idem, 57; Gopher and Blockman 2004: Fig. 9:1-20). At Nagila East, the vessels are handmade, straw-tempered, with uneven surfaces and usually with simple rims. Very few sherds show any surface treatment, with rare examples displaying wash and red slip. As opposed to the Yarmukian holemouths, which are open, the vessels from Nagila East are closed although several are wide enough to classify them as ‘open holemouth jars’.

Surface Treatment In general, the most common surface treatment is red slip, with 10% of the body sherds showing signs of it. It is likely that slip appeared on even more sherds, but was not preserved. Only two sherds show signs of burnishing (see above). Incised decoration is very rare, with only a single sherd bearing an incised line near the handle (e.g. Fig. 16:5). Other motifs, such as plastic decoration, are totally absent.

Five of the thirteen holemouth jars at Nagila are medium-sized vessels (Fig. 15:1-3) with thick walls and 157

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 15. Bowls.

petrographic (polarizing) microscope at magnifications of x20 to x200. The thin-sections were described with reference to charts and tables in Whitbread 1986 and Orton et al. (1993). The observed data was compared to the geologic and lithologic settings of the excavated site and its vicinity (Picard and Golani 1987; Bentor 1966; Neev et al. 1987; Blake and Goldschmidt 1947;

The petrographic analysis Four sherds were chosen for petrographic analysis. Prior to the petrographic examination, the samples were examined under the stereoscopic microscope at magnifications of x20 to x40. Then, four thinsections were prepared and examined under the Nikon 158

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 16. Jars.

Ravikovitch 1969) and to the results of previous petrographic analysis of 6th millennium BC pottery from other sites, such as: Teluliot Batashi, Lod (Goren 1991), Munhata (Goren 1992), Neve Yaraq (Goren 2004) and Tel Yosef (Goren forthcoming).

Sample Data All the samples (Table 2.4) have a light brown fabric with a grey to light grayish-brown core. The vessels show traces of organic temper (such as grass), which 159

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Core types

No

%

Total

%

One Striking platform

16

22.9

Type Arrowheads

2

2.5

Two Striking platforms

11

15.7

Sickle blades

4

5.0

Amorphous

43

61.4

Bifacial

1

1.2

Total

70

100

Scrapers

9

11.3

Notches and denticulates

26

32.5

Awls and borers

20

25.0

Table 2. Flint Core frequencies.

disappeared during firing, creating voids of various sizes (between 2.0 x 0.5 mm and 0.5 x 0.2 mm). Some of the voids, located inside the sections, are surrounded by grey aureoles, which are darker than the core. Another tempering material, present only in sample 3009, is grog (crushed pottery), whose grains are the same as the matrix, rounded to sub-rounded and varying in size between 2.0 x 1.0 mm and 0.7 x 0.5 mm. When located within the core they can even be distinguished by the naked eye, due to their light color caused by repeated firing.

Burins

4

5.0

Retouched flakes

11

13.7

Retouched blades

3

3.7

Total

80

100

Table 3. Flint Tool frequencies.

beds with Upper Eocene chalk is visible at Ein Nagila (Blake and Goldschmidt 1947). Conclusions Petrographic examination of 6th millennium BC pottery samples from Nagila East shows a general trend in the use of highly calcareous clay, which becomes a light color during firing, characteristic of the previously studied prehistoric sites mentioned above (Goren 1991; Goren 1992; Goren 2004; Goren forthcoming). Temper identified in the examined samples (i.e. organic material and grog) are typical of pottery of this culture (Goren 1991; Goren 1992; Goren 2004; Goren forthcoming).

The matrix is calcareous, slightly silty (1-2 percent) and extremely rich in foraminifers and their debris, together with thin (50%) and small bits of flint. The waste is characterized by primary and unmodified pieces while blades are almost absent (excluding one primary blade; Table 2.5). Only one CTE, an irregular ridge, was found. It is probably a product of striking platform rejuvenation. Cores (n=12) and choppers (n=10) Amorphous, elliptical and oval wadi cobbles were selected for the production of cores and choppers. The cores are large, displaying minimal removal on their debitage plain (average scar number is 5.2). The

Table 5. Ground stone and knapped limestone artifacts according to raw material.

162

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 19. 1-2 Grinding slab; 3-4 Mobil cup-mark.

dominant core type is the striking platform core (n=7; Fig. 20:3), followed by amorphous (n=3) and discoidal cores (n=2; Fig. 20:4). Most of the cores retained 1075% of cortex on their circumference. Both end and side choppers were present, commonly displaying one active edge. It possible that some of these specimens are in fact tested nodules or products of unsuccessful attempts to remove the initial striking platform.

reconstructed. One third of the primary elements are entame flakes, products of the nodules’ initial knapping. Two elongated primary elements are particularly interesting, with cortical lateral edges, displaying a side scar pattern on their dorsal surface. They are probably the product of platform preparation. The dominant scar pattern observed on the primary elements is cortical (10) and simple (n=7), while side (n=4), opposed (n=1) and radial (n=1) were also noted. The average number of scars on the dorsal surface of the primary flakes is one. In contrast, the unmodified flakes display more complex scar patterns and a higher number of scars on the dorsal surface (average 3.2). Multi-directional (n=9) scar patterning is dominaning although unipolar removals are also common (n=8). The remaining artifacts display indeterminate scars pattern (n=2).

Primary elements (n=23) and unmodified flakes (n=19) Most of the primary elements and unmodified flakes were found complete (ca. 79%). Oval (n=10), spherical (n=7) and sub-spherical (n=8) are the dominant shapes, although triangular (n=4) and square (n=4) objects are also recorded. The shape of nine items could not be 163

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead some degree of cortex on their surface, except for one retouched flake. Half of the tools are cortical, displaying more than 75% cortex, while the remaining tools exhibit less than 50% cortex. Similar to the unmodified flakes, the tools display multi-directional scar patterns (n=6) while side (n=4) cortical (n=3), simple (n=2) and indeterminate (n=2) dorsal scar patterns were also common. The number of scars on the dorsal surface is intermediate between primary elements and flakes 2.5 (S.D. ±4.6). An interesting phenomenon is that the striking platform of half of the tools was removed by the retouch or alternatively by preparing the butts to insert the blank into a haft. Scrapers and retouched flakes dominated the calcareous tool kit, denticulates and an awl (Table 2.7). Half of the scrapers were made on flakes, while the rest were fashioned on cobbles and pebbles. Most of them still display cortex on their butt (e.g. Fig. 21:23). The scrapers are varied in type and include simple but massive scrapers (Fig. 21:1-3), a nosed scraper and a side scraper (Fig. 22:1-2). Of particular importance

Figure 20. 1-2 Grooved stones.

Tools (n=17) Approximately 70% of the tools were found complete. Most of the blanks were oval (ca. 40%) or semi-spherical and spherical in shape (ca. 30%). All the tools display Raw material type

Chalk Limestone Sandstone Basalt Beachrock Flint Chert Metamorphic Total

Grinding stones-working slabs

2

9

Handstones

6

4

Pounders

14

1

1

2

4

7

7

Pestles

1

11 1

Hammerstones Abraders

3

Mortars

1

Cups

5

Grooved item Lid

12

1

16

3 2

1

4 5

2

2

1

Perforated item Sub-total groundstones

1

1 1

2

Modified items

35

1 15

1

1

10

2

8

Primary flakes

23

Unmodified flakes

18

CTE

1

Cores

10

Choppers

10

Knapped Tools

17

Sub-total knapped stones

87

Unidentified fragments

5

66 8 23

1

19 1

2

12 10 17

3 6

90 1

12

Recycled abrader & scraper

1

1

Recycled abrder & chopper

1

1

Recycled-pestle & core

2

2

Recycled core & hammerstone

2

2

Recycled-handstone & core

1

1

Sub-Total -others Total

2

12

6

135

21

1

1

Table 6. Cores on limestone artifacts.

164

10

3

1

19

3

175

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 21. Limestone artifacts. 1-2 abraders; 3-4 cores.

is a standardized denticulate, with its proximal part removed and subsequently used as core (Fig. 22:3). In addition, it is worth mentioning an awl with

multidirectional dorsal scar patterns and a bifacially thinned butt. The point of the awl is distally déjeté shaped by regular retouch (Fig. 22:2). 165

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 22. Limestone artifacts. 1-3 massive scrapers.

166

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila

Figure 23. Limestone artifacts. 1-2 side-scrapers; 3 awls; 4 denticulated artifacts.

167

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Tool Type

N

%

Retouched flakes

4

23.5

Endscraper on flakes

3

17.6

Denticulate on a core

1

5.9

Denticulate

3

17.6

Awls

1

5.9

Massive tools

4

23.5

Side scraper

1

5.9

Total

17

100.0

Lodian cultures produced small arrowheads exclusively. This can be explained either by major shifts in hunting strategies, which required using small points, or by a long process of decline in flint assemblage due to the increased use of metal and pottery (Rosen 1997:159164). The sickle blades are wide and rectangular, shaped by a combination of bifacial and pressure retouch that covers most or the entire surface. This shaping technique appears on a few Yarmukian sickles and becomes the main shaping technique of the Jericho IXLodian implements. The presence of small arrowheads together with wide rectangular sickle blades is a chronological marker of the Jericho IX culture of the sixth millennium BCE.

Table 7. Frequencies of knapped limestone tools.

Discussion

The pottery are also homogeneous, and show much resemblance to the Jericho IX cultural assemblages known in other parts of the country. Both technological and typological characteristics define the Nagila East assemblage. In general, the pottery is handmade using highly calcareous clay and temper with much straw. Petrographic analysis indicates that the pottery was probably manufactured on-site using available clay material. Some of the clay originated 12 km away. The surface was treated with a clay wash and many items were decorated with red slip. Due to poor preservation, it was difficult to determine whether these vessels were burnished or not, with only one vessel showing clear signs of burnishing. Other decoration, such as painted bands and incisions, is rare.

The salvage excavation east of Tel Nagila was restricted to a limited area that was under the threat of destruction. The excavation uncovered of a Neolithic settlement of the sixth millennium BCE. Remains were poorly preserved due to later human activities, including intensive plowing. Nevertheless, the remains included a number of shallow pits, living surfaces and possible remnants of one dwelling structure. The material culture is comprised of a homogeneous assemblage of pottery, stone and flint which can be dated to the early Pottery Neolithic period. The flint assemblage is small but homogeneous. The blade: flake ratio is 1:13, indicating this was primarily a flake industry. This observation is further supported by the dominance of flake cores and the main reduction sequence oriented toward flake production. The tool frequency resembles many EPN assemblages and is clearly dominated by ad hoc groups. However, two formal groups: arrowheads and sickle blades are worth noting.

Repertoire of vessels in this assemblage includes deep bowls, holemouth jars with closed and wide openings, and closed jars with everted rims. Bases that protrude slightly outward and with rounded indentations are also present. These types of bases are frequent in Jericho IX/Lodian assemblages. Assemblages of this culture are known from many pottery Neolithic sites in northern and southern Israel: Hagoshrim V (Getzov 2008), Tel Teo, Stratum X (Eisenberg, Gopher and Greenberg 2001) and Jericho IX (Kenyon and Holland 1982) to the north and east; Horvat ‘Uza, Stratum 20 (Getzov 2009) and in the recent excavations at Yiftahel, Area G (Khalaily et al. 2008), in the Galilee, and Lod (Kaplan 1977; Gopher and Blockman 2004), Nizzanim (Yeivin and Olami 1979) and

The few arrowheads are of the Ha-Parsa type, a small arrowhead present in Pottery Neolithic assemblages of the Levant (Gopher 1994). Such arrowheads appear in Yarmukian assemblages, but in lower frequencies than the large Amuq and Byblos points. The Ha-Parsa type became most common in Jericho IX-Lodian assemblages, becoming even the only type present in several assemblages. The people of the Jericho IXAverage length

SD

Average width

SD

Average Thickness

SD

Cores (n=10)

63.2

9.6

75.4

10.2

47

13.3

Choppers (n=8)

81.3

16.8

83.3

23.9

40

15.3

Primary elements & unmodified flakes (n=31)

54.4

21.8

48.5

16.6

16.2

9.4

Tools (n=13)

78.6

27.3

57.2

16.2

28.2

13.5

Type

Table 8. Dimensions of several limestone categories.

168

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila Ptora North (Milevski and Baumgarten 2008), further south.

area of excavation but the chronological differences are excluded.

Though the excavation was rather small, the presence of groundstone implements at Nagila East is pronounced (n=66) and represents a typical Neolithic domestic assemblage (see Garfinkel 2002; Gopher and Orrelle 1995; Rosenberg, forthcoming). Different types of local raw material, such as limestone, flint and sandstone were commonly used. Only one fragment of a basalt mortar was retrieved most probably was imported to the site as a final product. Grinding and pounding implements dominate the assemblage, displaying similar frequencies, which indicate that both activities were common at the site. Since the great majority of the stone implements were found on the living surface, it is safe to assume that most of the tasks related to pounding, grinding and flaking were conducted in this area.

The chipped limestone assemblage is exceptional, being the only non-flint chipped assemblage from PrePottery Neolithic contexts reported. One exception is the chipped basalt industry reported from the PPNC level at Hagoshrim (Level VI; Rosenberg and Getzov 2006). Knapped limestone assemblages are well documented at northern Negev Chalcolithic sites (Gilead 1989; Gilead and Fabian 1995; 2001; Gilead et al. 2004; Rowan 2006; P. Fabian, pers. comm. (Gat Guvrin and Tel Sheva), as well as in a few LPN sites, such as Qatif Y-3 (I. Gilead pers. comm.). The appearance of a Pottery Neolithic (Jericho IX/Lodian) chipped stone industry at Nagila East can indicate an earlier knapped limestone tradition which continues into the Chalcolithic period of the northern Negev. The lack of continuity in the data set and the isolated published assemblages make it difficult to support or reject this hypothesis at this stage of research.

Interesting phenomenon is the appearance of several knapping devices within the stone implements. These devices are limestone abraders and hammerstones. Hammerstones were probably utilized during the decortication and blank removal stages, while abraders were used for fine knapping, e.g. abrading of the core’s striking platform and tool preparation.

Exploration and publication of the Neolithic settlement east of Tel Nagila is an important contribution to the research of the Neolithic period and its settlement pattern. The recent excavations, together with the few finds from the old excavations, and the identification of the Jericho IX cultural horizon, comprise evidence for this being, to date, the southernmost site of the sixth millennium BCE in the southern Levant.

The grooved pebbles found at the site are well known in Yarmukian and Jericho IX/Lodian assemblages, in Nagila East they are in a clear Jericho IX/Lodian context. In the past, scholars (e.g. Gopher and Orelle 1995) related these artifacts to cultic activities on the household level as well as to functional tasks. In this case, the use marks observed in the grooves suggest that these artifacts were utilized in special tasks, such as sharpening. One of the pebbles was found within the living surface in a clear domestic context.

References Barkai, R. 2005. Flint and Stone Axes as Cultural Markers. Socio-Economic Change as Reflected in Holocene Flint Tool Industries of the Southern Levant. Studies in the Early Near Eastern Production, Subsistence, and Environment 11. Berlin, Ex Oriente. Bentor, Y. K. 1966. The Clays of Israel: International Clay Conference, Guidebook to the Excursions. Israel Program for Scientific Translations. Jerusalem. Blake G. S. and Goldschmidt, M. J. 1947. Geology and Water Resources of Palestine. Department of Land Settlement and Water Commissioner. Jerusalem. Berman, A. Stark, H. and Barda, L. 2004. Map of Ziqim (91). Jerusalem, Archaeological Survey of Israel. Berman, A. and Barda, L. 2005. Map of Nizzanim-West (87), Map of Nizzanim East (88). Jerusalem, Archaeological Survey of Israel. Blockman, N. 1997. The Lodian Culture (Jericho IX) Following the Excavation at Newe Yarak, Lod. Unpublished MA thesis, Tel Aviv University (Hebrew). Bochbinder, B. 1969. Geological Map of Hashephela Region, Israel, with Explanatory Notes, 1:20,000 Scale, 5 sheets. Report OD.1.68. Jerusalem, Geological Survey Israel.

The knapped stone assemblage is characterized by various types of scrapers, retouched flakes, denticulates and massive tools such as awls and denticulates, which were knapped from large nonstandardized cores. In general, the blanks are longer, wider, and thicker than the waste (primary elements, flakes – Table 2.8). The tools were shaped at the site on irregular flakes, which retained cortex on their dorsal surfaces. Their striking platform was frequently removed by faceting or thinning the butt. Formal tools, such as bifacials and rounded/circular scrapers are absent. They are common in some Chalcolithic chipped limestone assemblages (e.g. Gilead and Fabian 1995: 6.5:1-2, 6.9:1-2; 2001:75; Rowan 2006: Figures 11.25:6-7, 11.37:2). It is likely that the absence of bifacial tools at Nagila East is due to the small retrieved sample of both flint and limestone artifacts, or to the tasks preformed in the 169

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Crowfoot-Payne, J. 1983. The Flint Industries of Jericho. In K. M. Kenyon and T. A. Holland (eds), Excavations at Jericho, V. Appendix C: 622-759. London, British School of Archaeology in Jerusalem. Eisenberg, E. Gopher, A. and Greenberg, R. 2001. Tel Teo: A Neolithic, Chalcolithic, and Bronze Age I site in the Hula valley. Israel Antiquities Authority Reports 13. Jerusalem, Israel Antiquities Authority. Epstein, C. 1984. A Pottery Neolithic Site near Tel Qatif. Israel Exploration Journal 34:209-219. Garfinkel, Y. 1999. Neolithic and Chalcolithic Pottery of the Southern Levant. (Qedem 39). Jerusalem, Hebrew University of Jerusalem. Garfinkel, Y. 2002. Sha’ar Hagolan. Neolithic Art in the Jordan Valley. Jerusalem. Israel Exploration Society. Garfinkel, Y. 2004. The Goddess of Sha’ar Hagolan: Excavation at a Neolithic Site in Israel. Jerusalem, Israel Exploration society. Garfinkel, Y. and Miller, M. A. 2002. Sha’ar Hagolan 1: Neolithic Art in Context. Oxford, Oxbow. Garfinkel, Y., Dag, D., Horwitz, L-K, Lernau, O. and Mienis, H. K. 2002. Ziqim, A Pottery Neolithic in the Southern Coastal Plain of Israel: A Final Report. Mitekufat Haeven 32:73-145. Garstang, J., Ben-Dor, I. and Fitzgerald, G. M. 1936. Jericho: City and Necropolis (Report for Sixth and Concluding Remarks). Liverpool Annals of Archaeology and Anthropology 23:67-90. Getzov, N. 2008. Ha-Gosherim. In E. Stern (ed.), The New Encyclopedia of Archaeological Excavations in the Holy Land, Supplementary Volume 5, 1759-1761. Jerusalem, Israel Exploration Society. Getzov N., Lieberman-Wander, R. Smithline, H. and Syon, D. 2009. Horbat ‘Uza, The 1991 Excavations, Volume I: The Early Periods (Israel Antiquities Authority Reports 41). Jerusalem, Israel Antiquities Authority. Gilead, I. 1990. The Neolithic-Chalcolithic Transition and the Qatifian of the Northern Negev and Sinai. Levant XXII: 47-63. Gilead, I. 1989. Knapped Limestone Series from the Chalcolithic Period of the Northern Negev, Israel. Mitekufat Haeven- Journal of the Israel Prehistoric Society 22: 98-111. Gilead, I. and Fabian, P. 1995. Knapped Limestone Series from Grar. In I. Gilead, (ed.), Grar- A Chalcolithic Site in the Northern Negev: 281-309. Beer Sheva: Ben- Gurion University of the Negev Press. Gilead, I. and Fabian, P. 2001. Nevatim: A Site of the Chalcolithic Period in the Northern Negev. In A. M. Maier and E. Baruch (eds), Settlement, Civilization and Culture. Proceedings of the Conference in Memory of David Alon: 67-86. Ramat-Gan: Bar-Ilan University Press (Hebrew). Gilead, I., Marder, O., Khalaily, M., Fabian, P., Abadi, Y. and Yisrael, Y. 2004. The Beit Eshel Chalcolithic Flint Workshop in Beer Sheva: A Preliminary Report. Mitekufat Haeven- Journal of the Israel Prehistoric Society 34:245-263.

Gopher, A. 1989. The Flint Assemblages of Munhata (Israel). (Les Cahiers du Centre de Recherche Français de Jerusalem 4). Paris, Association Paléorient. Gopher, A. 1994. Arrowheads of the Neolithic Levant. American Schools of Oriental Research Dissertation Series 10. Winona Lake, Indiana, Eisenbrauns. Gopher, A. and Blockman, N. 2004. Excavations at Lod (Neve Yarak) and the Lodian culture of the pottery Neolithic period. ‘Atiqot 47:1-50. Gopher, A. and Gophna, R. 1993. Cultures of the Eighth and Seventh millennia BP in the southern Levant: A Review for the 1990’s. Journal of World Prehistory 7:3:297-353. Gopher, A. and Orrelle, E. 1995. The Ground Stone Assemblage of Munhata- A Neolithic Site in the Jordan Valley – Israel: A Report. Cahiers du Centre de recherché français de Jerusalem 7. Paris, Association Paléorient. Gophna, R. and Blockman, N. 2004. The Neolithic, Chalcolithic, Early Bronze and Intermediate Bronze Age Pottery. In D. Ussishkin (ed.), The Renewed Archaeological Excavations at Lachish (1973-1994), Vol III, 873-899. Monograph Series of the Sonia and Marco Nadler Institute of Archaeology, no. 22. Tel Aviv, Tel Aviv University. Goren, Y. 1990. The Qatifian Culture in Southern Israel and Transjordan: Additional Aspects for its Definition. Metikufat Haeven. Journal of the Israel Prehistoric Society 23:100-112. Goren, Y. 1991. The Beginnings of Pottery Production in Israel, Technology and Typology of ProtoHistoric Ceramic Assemblages in EretzIsrael (6th4th Millenia B.C.E.). Unpublished PhD dissertation, The Hebrew University of Jerusalem (Hebrew). Goren, Y. 1992. Petrographic Study of the Pottery Assemblage from Munhata. In Y Garfinkel, The Pottery Assemblages of the Sha’ar Hagolan and Rabah Stages of Munhata (Israel): 329-360. Paris, Association Paléorient,  Goren, Y. 2004. Technological Study of the Ceramic Assemblage from Nevè Yaraq, Lod. ‘Atiqot 47:5155. Goren, Y. forthcoming. The Typology and Technology of the Pottery Neolithic Ceramic Assemblage from Tel Yosef (Tell esh Sheikh Hasan). ‘Atiqot. Horowitz, A. 1977. The Quaternary Stratigraphy and Paleogeography of Israel. Paléorient 3:47-100. Kaplan, J. 1958. The Excavations at Teluliot Batashi, Nahal Soreq. Eretz-Israel 5: 9-24 (Hebrew with English summary on pp. 83*-84*). Kaplan, J. 1977. Neolithic and Chalcolithic Remains at Lod. Eretz-Israel 13:57-75 (Hebrew). Kenyon, K. M. and Holland, T. A. 1982. Excavations at Jericho, Vol 4: The pottery type Series and Other Finds. London, British School of Archaeology in Jerusalem. Khalaily, H. 1999. The Flint Assemblage of Layer V at Hagoshrim: A Neolithic Assemblage of the Sixth Millennium B.C. in the Hula Basin. Unpublished 170

H. Khalaily et al.: An Early Pottery Neolithic (Jericho IX) Site East of Tel Nagila MA thesis, The Hebrew University of Jerusalem (Hebrew). Khalaily, H. 2006. Lithic Traditions during the Late Pre-Pottery Neolithic B and the Question of the Pre-Pottery Neolithic C in the Southern Levant. Unpublished PhD Thesis, Ben-Gurion University of the Negev (Hebrew). Khalaily, H., Milevski, I., Getzov, N., Hershkovitz, I., Barzilai, O., Yarosevich, A., Shlomi, V., Najjar, A., Zidan, O., Smithline, H. and Liran, R. 2008. Recent Excavations at the Neolithic Site of Yiftahel (Khalet Khalladyiah), Lower Galilee. Neo-Lithics 2/08: 3-11. Milevski, I. and Baumgarten, Y. 2008. Between Lachish and Tel Erani: Horvat Ptora, a New Late Prehistoric Site in the Southern Levant. In: J. Mª Córdoba, M. Molist, Mª. C. Pérez, I. Rubio, I and S. Martínez (eds), Proceedings of the 5th International Congress on the Archaeology of the Ancient Near East: 1429-1446. Madrid, Universidad Autónoma de Madrid Ediciones. Neev, D., Bakler, N. and Emery, K. O. 1987. Mediterranean Coasts of Israel and Sinai: Holocene Tectonism from Geology, Geophysics and Archaeology. New York, Taylor and Francis. Noy, T. 1977. Neolithic Sites in the Western Coastal Plain. Eretz-Israel 13:18-33 (Hebrew). Olami, Y., Burian, F., and Friedman, E. 1977. Giv’at haParsa: A Neolithic Site in the Coastal Region. EretzIsrael 13:34-47 (Hebrew with English summary). Orton, C., Tyers, P. and Vince, A. 1993. Pottery in Archaeology. Cambridge, Cambridge University Press. Ravikovitch, S. 1969. Soil Map 1:250,000. Israel, North. Survey of Israel. Jerusalem, Geological Survey of Israel. Rosen, A. M. 1986. Quaternary Alluvial Stratigraphy of the Shephela and its Paleoclimatic Implications. Jerusalem, Geological Survey of Israel.

Rosen, A. M. 1991. Early Bronze Age Tel Erani: An Environmental Perspective. Tel-Aviv 18:192-204. Rosen, S. A. 1988-1989. Pottery Neolithic Artifacts from Tel Lachish. Tel Aviv 2/15-16: 193-196. Rosenberg, D. and Getzov, N. 2006. A Basalt Chipping Floor from Level VI (PPNC) at Hagoshrim. Mitekufat Haeven- Journal of the Israel Prehistoric Society 36:117129. Rosenberg, D. in press. The Stone Assemblage of Hagoshrim – Continuity and change in the Neolithic of Northern Israel. In N. Getzov and H. Khalaily (eds), Hagoshrim. Israel Antiquities Authority Reports. Jerusalem, Israel Antiquities Authority. Rowan, Y. 2006. The Chipped Stone Assemblage at Gilat. In T. E. Levy (ed.), Archaeology, Anthropology and CultThe Sanctuary at Gilat, Israel: 507-574. London and Oakville, Equinox Publishing, Ltd. Sneh, A., ed. 2008. Geological Map of Israel 1:500000. Sheet 10-IV (Qiryat Gat). Jerusalem, GSI, digital editing http://www.gsi.gov.il/eng/?CategoryID =253&ArticleID=788&dbsAuthToken (accessed 02.01.2019) Sneh, A. and Bochbinder, B. 1984. Miocene and Pleistocene Surface and their Associated Sediments in the Shephela Region, Israel. Geological Survey of Israel Current Research 1983-84: 60-64. Whitbread, I. K. 1986. The Characterization of Argillaceous Inclusions in Ceramic Thin Sections. Archaeometry 28(1):79-88. Wright, K. I. 1992. The Classification System for Ground Stone Tools from the Prehistoric Levant. Paléorient 17:19-45. Yeivin, E. and Olami, Y. 1979. Nizzanim: A Neolithic Site near Nahal Evtah. Excavations of 1968-1970. Tel Aviv 6:99-135.

171

Clothes Maketh (Hu*)Man: Textile Production in the Southern Levant in the Chalcolithic Period Janet Levy Further evidence for domestic craft specialization, known but discussed to a far lesser degree even in comprehensive cultural overviews, is the textile repertoire attested from the Judean Desert cave sites, including a unique textile measuring 7 m x 2 m, and the dramatic increase in spindle whorl frequency, the tools of production, from the Beer Sheva Valley sites, Gilat and Teleilat Ghassul. Two recent articles co-

Introduction The Ghassulian, the main culture of the Chalcolithic period dated to 4500-3800-3700 BCE cal., is characterized by small rural settlements. These settlements are located along the coastal plain, the length of the Jordan valley and densely concentrated along the wadis at the northern Negev desert fringe as far east as Arad. The economy is based on cultivation of cereals and pulses and raising of livestock, primarily goats and sheep. The salient features of the material culture include bifacial adzes and chisels, a ceramic repertoire of small V-shaped bowls manufactured on the potter’s wheel, ceramics decorated with red paint to include churns and cornets, basalt bowls and censors, planar violin shaped figurines primarily in stone, and frescoes, with figurative and geometric motifs, some measuring 4 m x 1.5 m. Secondary burial in gabled, box-like clay ossuaries, often with plastic and/or painted anthropoid and zoomorphic features, and also in domestic type vessels, in communal off-site cemetery caves and on-site under-floor burial of infants in ceramic vessels, are the predominant mortuary modes. Also encountered is the earliest evidence for formal temples within enclosures with altars, favissae and ablution facilities. Evidence for incipient domestic craft specialization is attested in ivory and bone artifacts, including a subterranean atelier with non-local raw material and also a corpus of copper artifacts of unprecedented magnitude and virtuosity manufactured in open cast and lost wax technique (Rowan and Golden 2009) (Fig. 1). Original quotation is attributed to Mark Twain. In the light of current academic conventions the title has been amended to include the broadspectrum category humans. *

Figure 1. Map of main sites mentioned in the chapter, courtesy of Isaac Gilead.

172

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period authored by Professor Isaac Gilead and myself (2012, 2013) specifically orientated to archaeological textile specialists are largely unknown even to archaeologists of the Chalcolithic period. The purpose of this article is to discuss this aspect of the Ghassulian culture. This study is directed to the non-specialist archaeologist. Consequently, I have reduced esoteric technical details to a minimum. In addition, I emphasize the integral role of textile production in the economy. Moreover, I bring to the reader’s attention the universality of basic techniques of textile production and observable, wide regional, longevity of tools, processes and cultural behavior.

and toe, by yarn stretched between two fixed bars in a horizontal mode, or by pendant weights in a vertical mode. A textile is formed from two sets of thread; the skeleton warp threads, passive and in parallel, are held in tension on the long axis, and the active weft threads, positioned at 90o, which interlaces with each warp thread in succession to form a coherent whole. Warp threads are divided into two sets: odds and evens, each of which is caused to rise in succession. A shed rod, flat and wide, held within the warp raises one set. Thick and strong yarn, heddle leashes attached each in succession, to the second warp set and also to a heddle rod, which lies across the face of the warp, raises the second set. When a set of warp threads is raised, a triangular space is formed (the shed). The size of this space is determined by the length of the heddle leashes or the width of the shed rod, through which the weft passes in a single action. The weft, in its simplest mode, is carried on a stick (stick spool) wound in a figure of eight, to maximize the length of yarn and to minimize the width. The newly inserted yarn is compacted with a smooth stick (sword/beater) or a short comb. The body tensioned loom e.g. the back-strap loom and the horizontal ground loom (six sticks, breast beam, warp beam, shed rod, heddle rod, beater and stick spool) are archaeologically invisible. The warp weighted loom with clusters of sun dried or fired clay or stone weights from storage or destruction levels leaves evidence (Barber 1991:79-118; Broudy 1979; Roth 1918: 1-16: Speir 1970; 87-90). It appears in the Southern Levant in MBII apart from a single occurrence in EBI at Tell Abu Kharaz (Fischer 2008: Fig. 39; Shamir and Baginski 1998: 55).

A textile navigational guide for the perplexed Archaic and obsolete terms for artefacts and processes, incomprehensible to the uninitiated, are the professional parlance of archaeological, textile specialists and craft revivalists. Herein is a minimal toolkit, not only to negotiate those murky archaeological waters, but also to understand the prevalent current manufacturing processes of most of the clothing worn by all persons who read this article. Worldwide, most fabrics are made from spun yarn. Yarn is formed from aggregates of ordered, parallel, flexible fibres, animal or vegetable, twisted together in one direction to form lengths longer and stronger than found in nature (Kissell 1918: 5). The yarn of the Southern Levant has been spun (twisted) on handheld spindles since the Yarmukian [mid-7th millennium BCE] (Garfinkel 1999: 29). This process continued until the mid-19th century CE in urban environments and until the second half of the 20th century CE in rural environments of southwest Asia and the Nile valley (Weir 1970: PL. 3). Its use was also attested in the 20th century throughout rural Europe (Hochberg 1980: 4, 64; Wilson 1896: 968-969). Most utilitarian spindles feature a wooden shaft on which is lodged a whorl of stone, ceramic or wood, a flywheel, which augments the twist inserted by the hand of the spinner. The newly formed yarn is wound around and temporarily stored on the shaft (Liu 1979: 98). Yarn is spun to the left, anti-clockwise (s) or to the right, clockwise (z). The orientation does not offer a technological advantage. The convention is culturally determined. Early yarn of the region is plied. In other words, two spun strands are spun together, generally in the opposite direction to the initial spin, with a similar developmental trajectory in Europe (Rast-Eicher 2012: 380). Plying in early yarn is a compensatory measure for deficiencies in spinning skills.

Textile findings documented in the most basic mode define raw material, direction of either initial or visible twist and if plied in what direction and the number of strands. Thus s2S, signifies two single spun yarns, spun individually in an anti-clockwise direction and subsequently spun together also in an anti-clockwise direction. This is a typical yarn structure of both the Chalcolithic and Early Bronze age of the Southern Levant. Convention dictates that spinning direction is recorded in lower case and plying direction in upper case. Direction of spin is a marker for an ethnos, a learning group, a geographical area or a period. Research conducted amongst hand spinners has determined that there is not any correlation between spin direction and left or right-handedness (Minar 2002: 92). Significantly, there is no correlation between the intrinsic direction of twist of the fibre and the direction of spin selected by the spinner. Direction is determined by the manner in which the skill was acquired. A master-apprentice relationship will result in a faithful replication of hand movements and body postures and resultant yarn character – an unbroken continuum over countless generations The number of threads in warp and weft (thread count) are also documented; e.g. 16 x 12 tpc (threads per centimeter). It is only possible to

Yarn is woven into a textile on a loom, which is a tensioning device.1 Tension is provided by manipulating body parts, e.g. by yarn held taut between forefinger Fabrics, i.e. twinning or netting were manufactured prior to woven textiles but their construction was not amenable to mechanization only interlacing (darning) could be mechanized. 1 

173

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead differentiate between warp and weft if a selvedge (an edge) is present. Convention dictates that warp threads are recorded first. Thread count ratios are markers for raw materials, periods, or culture areas. Warp threads of the Chalcolithic period in the Southern Levant are generally slightly denser than the weft threads. Weave type is also recorded. Plain weave/linen weave/ tabby weave, one over one under, is the most common weave type until the present day, and typical of the Chalcolithic period.

Flax is no longer cultivated for fibre purposes in southwest Asia; a little is still grown in China. This composite reconstruction of traditional methods of cultivation and processing is based primarily on Egyptian tomb graphics and accompanying terse explanatory phrases from c. 2000 BCE and a report by Crowfoot (1931: 32-36). In fact, Crowfoot witnessed in 1926 in Lower Egypt all three stages of flax processing in the last two villages to cultivate flax for textile purposes. Also relevant, despite climatic differences and water regimes are raw materials, tools and products from submerged late Neolithic Swiss lakeside sites, historical texts, and photographic evidence from rural Europe. The latter finds are important because the intrinsic attributes of flax impose, within a certain leeway, modes of processing, sequences and tool morphology. I conducted experimental archaeology in cultivation and processing to evaluate the potential of the Beer Sheva valley as a locus of cultivation and to understand the multiple phases of processing and associate problems (Fig. 2).

Flax – history Flax2 was the only fibre used for textile production in the Southern Levant in the Chalcolithic and Early Bronze Age. The earliest evidence for wool is from an MB II burial at Jericho (Shamir 2005: 23). In Mesopotamia, defined as ‘The Land of Wool’, it was rare before the 3rd millennium (Breniquet 2014: 65). The oldest attested domesticated linen yarn appears as the binding agent for a non-provenanced composite comb of myrtle wood and bitumen attributed to Wadi Murrabba’at dated to 10, 220 BP not cal. (Bonani 1995: 205; Schick 1995:199202; Shimony 1995:204; Werker 1995: 203). The oldest linen fabrics in linking, looping, twining and knotting techniques, but not interlacing the precursor of mechanized weaving, were found in the MPPNB cave of Nahal Hemar (Schick 1988: 34-38). Kvavadze et al. (2009, 2010) report the microscopic presence of twisted wild flax fibres in various colours in an Upper Palaeolthic stratum dated to 30,000 BP at Dzudzana Cave, Georgia. It is contested (Bergfjord et al. 2010: Breniquet 2013: 7).

Flax-cultivation Flax is not user-friendly; it is grown on prime land, and it has high water demands in cultivation and processing. Processing conducted over several weeks is labour intensive. The fibre yield is low, and there are no byproducts. The genus Linum with more than 200 species is found throughout the New and Old Worlds. Linum bienne syn. Linum angustifolium, the species taken into domestication, is indigenous to Crimea, Caucasia, western Europe and the circum Mediterranean basin, albeit rare in the Southern Levant (Abbo et al. 2014: 2, 11). It is a bast fibre, like nettle or hemp, meaning that the fibres are contained within the stalk as opposed to cotton in which the fibres surround the seed (Emery 1966: 3). Domestic flax, Linum usitatissium, is a highly varied plant, an adaptive response to a wide range of climates and methods of cultivation, giving rise to differences in height, colour of flowers, seed size and growth cycles (Helbaek 1959:107).

Archaeobotanical remains attests to wild flax seeds at 12th millennium cal. BCE Tell Mureybit on the Euphrates (Zohary, Hopf and Weiss 2012:103-106 with additional sites therein). Flax seeds were also found at Cayönü, Turkey and Tell Halula, Syria. Both sites also attest to fine twined fabric impressions on bone and plaster (Alfaro Ginar 2012: 42, Figures 3-4; VogelsangEastwood 1993 4-6). Linen fabrics and textiles, initially determined as woolen, were recovered from mortuary contexts from early ceramic layers at Ҁatal Hüyük (Ryder 1965: 175-176; Vogelsang 1987:18). A gradual increase in the size of the flax seeds and the presence of non-splitting seed capsules, both characteristics of domestication, and also the use of linen fabrics suggest that flax was already cultivated in the PPNB (Weiss and Zohary 2011: S251). However, it is suggested that flax was initially brought into cultivation for its oil-bearing seeds (linseed) and not for its fibres (Weiss and Zohary 2011: S250).

Flax in sub-tropical areas is a winter annual sown in rain fed areas after the second rains and in the Nile Valley after the annual inundation (Vogelsang-Eastwood 2000: 38). It is sown densely for tall stalks for fibre purposes. It is sparsely sown to encourage branching for seed growth for oil purposes. The crop requires 450-750 mm precipitation or careful irrigation in areas of perennial water since the plant has a very shallow taproot and is sensitive to drought (Bedigan 1985: 160; Renfrew 1973:124). Wherever flax was grown, apart from in the great alluvia, with annually renewable topsoil, it was subject to rotational cycles of 7-10 years (Baines 1989: 167; Dempsey 1975: 14; Kislev et al. 2011: 2). Flax with stalks of 1.5-2 mm grows to a height of 0.7-1.3 m. It is ready for harvesting for general textile use when

Flax is a word of Germanic origin and denotes the plants and raw fibres. Linen is a word of Latin origin and denotes yarn spun from flax fibres and fabrics or textiles manufactured from spun linen yarn. Most European languages have one word which covers all states. 2 

174

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period

Figure 2. Harvesting flax, author’s garden, near the Beersheva Valley, March 2013.

the stem changes from green to yellow in c. 100 days (Vogelsang-Eastwood 2000: 270; Weindling 1947: 237). Wherever flax is grown for fibre purposes it is harvested by uprooting. The root also contains usable fibre and cutting would expose the fibre ends during drying and retting causing discoloration (Weindling 1947: 237; Dempsey 1975: 20). Flax-processing In the Southern Levant and Egypt the harvested flax is laid out to dry as opposed to upright stooks/ beets in Europe (Crowfoot 1931: 32; Joshua 2:7) and subsequently the seed bolls are stripped by hand or run through the teeth of an upright, fixed comb. This is seed stock, the basis for a porridge like dish or a source of non-edible oil. Linseed oil goes rancid very quickly. Consequently, it is only suitable for immediate consumption. However, it has wide industrial applications, e.g. in leather processing (Bedigan 1985: 161; Kislev et al. 2011: 583). Olive oil production, with comparatively long storage potential, is attested in the Southern Levant from the PN (Galili and Sharvit 1994-5: 122).

Figure 3. Flax stem anatomy A=cuticle, B=epidermis, C=cortex, D=fibre bundles, E=woody core, F=pith, G=lumen after Baines 1989 Fig. 2, courtesy of Yuval Shach.

process of bacterial decomposition.3 The bundles of flax, which become black and slimy, are turned daily. The water, which becomes increasingly acid, is also changed frequently. The cold criterion for completion of the ret (formed from the causative of the verb to rot) is that an individual straw is broken into several pieces and the casings fall away (Baines 1989: 174). However, obviously far subtler criteria, based on experience and familial craft knowledge, determined its completion. The correct degree of retting is critical. Insufficient retting will cause difficulties in the subsequent mechanical separation of the woody casings, and over retting will give rise to weak yarn (Baines 1989: 174; Dempsey 1975: 30; Weindling 1947: 239).4

The flax fibres, located along the length of the stalk between the outer cortex and the inner woody tissue, are embedded in water soluble pectin. This causes the fibres to adhere one to another and to the surrounding tissue (Weindling 1947: 234) (Fig. 3). To release the fibres the bundles of flax can be either spread out in the fields and exposed to dew and rain, inapplicable to the climatic conditions of southwest Asia, or submerged and weighted down in damned pools or artificial tanks, retted, located downwind from habitation for a period of 7-14 days, depending upon the characteristics of the water and the ambient temperature (Baines 1989: 17, Fig. 123; CarringtonSmith 1975: 22) Water retting is a stinking, polluting

The earliest recorded environmental legislation enacted in 16th century England restricted the locations of retting with heavy fines for infringements (Hochberg 1979: 53). 4  Exactitude in retting is critical to such a degree that Jewish texts state that a mourner, who must refrain from all labour, is allowed to remove his flax from the retting locus ( Mishnah, Moed Qatan 2:1; Babylonian Talmud, Moed Qatan 11b-12a). 3 

175

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead The flax is removed and again completely dried to arrest further decay.

as tools and body parts as work surfaces. However, the yarn is coarse and often weak and production is very slow. Such methods are attested worldwide (CrockerJones 1990; 15, 24, Fig. 8; Frödin and Nordenskiöld 1918: 17; Kissell 1918: 19-19). Spindles used in the Old and New Worlds in the manufacture of yarn vary very little in their basic morphology. However, size and weights vary, as does whorl raw material dependent on suitable local material availability and later on status. Ideally, the shaft is straight, slender, tapering to a point at the lower end and strong enough to sustain constant impact (Raven 1987: 19). Ethnographically. wood predominates for both shaft and whorl. Crowfoot (1931; PLS 35.1, 41.12) published three intact Egyptian spindles, two from 3rd-2nd millennium BCE and one from the early 20th century CE used in spinning flax which ranged from 28-18 cm (Crowfoot 1931; PLS. 35.1, 41.1-2). Diagonal grooves are sometimes observed on the extreme upper end of the shafts of archaeologically intact spindles and on ethnographic specimens (Crowfoot 1954; Fig. 273.H; Hochberg 1980: 62). The grooves lead the newly spun yarn to the unspun fibre source or lead it to the spliced rove. The orientation of the groove is a marker for the direction of spin (anti-clockwise S/clockwise Z). Spindles currently used in southwest Asia feature shafts with truncated upper ends and flat surfaces to accommodate metal hooks. The use of a groove or hook, as opposed to the earlier half hitch made by looping the yarn around the thumb, as when adding on stitches in knitting, increases the rate of production by 30-50% (Carrington-Smith 1975: 82 n. 3; Hochberg 1980: 40: Ryder 1968: 79).

To extract the fibres a bundle of flax straw, rendered brittle by retting, of a quantity that can be comfortably held in one hand is lain horizontally on a stone anvil or wooden block and beaten with measured blows along its length (braking) with a smooth heavy piece of wood (Baines 1989: 177, Fig. 124; Crowfoot 1931: PL. 18). Measured is the keyword. The purpose is to crack the brittle casings but to retain as far as possible the integrity of the fibres. The same bundle is subsequently held in a vertical position against a wooden surface and lightly tapped in a downward motion with a flat piece of wood and shaken (scrutching) causing most of the casing fragments to fall away (Baines 1989: Fig. 125; Crowfoot 1931, PL. 20). Finally, the fibres are drawn through an upright set of teeth (hackling) to remove the residual casings and the short broken fibres (tow) and also to align the long fibres ready for splicing or draft spinning (Baines 1989: Fig. 126; Barber 1991: 14, Fig. 1.2; Crowfoot 1931, Pl. 21).5 At the end of this labour intensive process, one is left with hanks (stricks) of flax fibres 45-60 cm in length (line), textile quality and short broken fibres (tow) suitable for cordage (Baines 1989: 14). Figures for fibre yields from the early 20th century CE with a measure of mechanization, improved seed stock and limited use of industrial fertilizer are 5%, of which 2.5% is of textile quality (Weindling 1947: 151). A recent study conducted in Denmark with a detailed breakdown stage by stage records a 96% loss from harvesting to processed fibres ready for spinning (Ejstrud et al. 2011: 80).

The whorl is generally round. Among the many possible permutations are spheroid, cone, bellshaped, lens, cylinder etc. ranging from 0.8-15 cm (Liu 1978: 90) and even as large as 20 cm (Hochberg 1980: 37) with a central, vertical perforation ranging from 3-10 mm in diameter and a cylindrical, tapered or biconical profile (Lui 1978: 97). All the perforations of sherd or stone whorls of the Southern Levant from the Chalcolithic period are biconical. Archaeological, functional whorls are predominantly of stone, sherd or clay. The recently proposed exotic lead whorl is contested (Langgut et al. 2016: 981, Table 2, 985; BenYosef et al. 2017) To wedge the whorl firmly into position to effect efficient spinning the whorl must be securely packed around the shaft (Crewe 1998: 12). Ethnographic reports attest to the use of unspun fibre, wax and resin (Frödin and Nordenskiöld 1918: 26; Liu 1978: 97 citing Snethlage 1930: Fig. 27). Traces of adhesive, e.g. bitumen, have not been attested in the archaeological record of the Southern Levant. However, two recently published basalt whorls with fragmentary wooden shafts, recovered from a Judean Desert cave site, feature linen textile fragments in the perforation to affect a tight fit (Ben-Yosef et al. 2017: Fig. 4).

In sum, flax cultivation and processing is a drawn out, labour intensive process with low yields. Abbo et al. (2014: 14), who recently foraged and processed local wild flax, commented on the high labour investment and low returns vis à vis wild cereals. The crop demands good land and ample water for its cultivation and surplus water that cannot be recycled for its retting. Retted stalks cannot be discarded on agricultural land because they will cause the soil to sour. Therefore, they must be burnt (Hess 1954: 255). The 4th-3rd millennium BCE genesis of sheep wool brought abought the virtual demise of linen clothing. It was retained only for the clothing and coverings of the deities and the elite (Breniquet 2006: 167; McCorriston 1997: 518). Spinning – tools of production Most yarn is spun on a spindle, but is not requisite. Fibres can be twisted into yarn using fingers or palms Fibre processing is an outdoor activity. Both braking and scrutching generate dust which carries bacteria from the retting phase. Flax workers are prone to brown lung disease (Andersson 2012: 27; Baines 1989: 27). After braking and scrutching my hands were covered in whiteheads although the sharp straws had not broken the skin. 5 

176

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period spinning supported on the thigh is attested at 2nd millennium Beni Hassan, Egypt and observed in 20th century Sudan and Transjordan and also amongst the indigenous populations of the American southwest and more remote tribes of tropical South America (Crowfoot 1931: Figs. 2-3; Frödin and Nordenskiöld 1918:92, PL. 43; Kissell 1918: 25; Weir 1970: 9, PL. 2; Wilson 1896: PL. 22). Ethnographically, it is characterized by large spindles with correspondingly large whorls with coarse uneven yarn, which is frequently re-spun (Kent 1961: 8; Kissell 1918: 26). Spinning supported on the thigh is conducted while seated with the upper shaft of the spindle resting on the thigh at an angle of c. 300 and the lower tip on the ground. Ordered and teased out raw fibre is held in the left hand connected to spun yarn wound on the shaft (Fig. 4). The spindle is propelled down the right thigh (but up the right thigh in the Southern Levant) while the left hand simultaneously slowly attenuates the fibres moving away from the spindle until the arm is fully extended. Thereupon, the rotation of the spindle ceases and the new yarn is wound onto the shaft. Each stage, drawing out, twisting, and winding on is carried out in sequence. The length of the roll is short, and production is slow (Anderson 1987: 224; Crowfoot 1931; 18). The technique, but one in which paying out the yarn and twisting are carried out simultaneously, is widely used in the Nile valley and southwest Asia for plying wool from either two balls or from a ball composed of two single spun strands wound up as one (Crowfoot 1931; 13, PL. 15).

Figure 4. Supported on the thigh spinning, Navajo, from Wilson 1896 PL.22.

The whorl is generally lodged just below just below the top of the shaft (high whorl) or just above the bottom (low whorl). Representations of spinning with high whorled spindles are attested in Mesopotamia and Egypt from the 4th-1st millennium BC (Barber 1991: Figs. 2.16, 2.21, 2.23, 2.43: Crowfoot 1931: PL 16; Vogelsang-Eastwood 2000: Fig. 11.4b; Wiggermann 2000: Fig. 4.44) with a current continued tradition in Egypt, Sudan, Ethiopia, Saudi Arabia and the Southern Levant (Crowfoot 1931: PLS 14. 24, 32; Hochberg 1980: 63; Weir 1970: PL. 3) and in isolated locations in Europe. Representations of spinning with low positioned whorls and the use of distaffs (straight sticks around which is lapped the prepared rove) appear in the 1st millennium BCE in the Aegean world where it is still the current tradition (Barber 1991:70, Fig. 2.35; Hochberg 1980:11, 56). It is also historically and currently the predominant whorl position throughout Europe and Russia (Hochberg 1980: 19, 50, 51 55).

Drop/suspended spinning, which is the most evolved of pre-mechanized hand spinning techniques, is attested worldwide (Frödin and Nordenskiöld 1918: 22; Carrington-Smith 1975: 71; Kramer 1982: 44-45, Fig. 2.4) (Fig. 5). It would appear to have at least two independent loci of invention, southwest Asia and the Andes (Frödin and Nordenskiöld 1918: 46; Ryder 1968: 74). It is the primary spinning mode depicted in the early 2nd millennium tomb complex at Beni Hassan (Vogelsang-Eastward 2000: 272) and on 1st millennium BCE vessels from the Aegean and Middle Europe (Barber 1991: 72, Fig. 2.38; Hochberg 1980: 11, 41; Kissell 1918: 31, Fig. 8). Ethnographers (Crowfoot 1931: 12; Weir 1976: 40) observed that in south west Asia and the Nile Valley the fibre mass, predominantly wool but also cotton and flax, connected to the already spun yarn on the shaft, is loosely held in the left hand or is worked as a strand wound around the wrist or thrown across the left shoulder and paid out with the left hand. In Europe distaffs, fibre holders, some simple sticks, and others elaborately carved (Hochberg 1980: 52-53) are de rigueur. Spindles with high whorls are propelled down the right thigh with the palm of the hand, exploiting the force of gravity (but up the right thigh in the Southern Levant) and then rotate in free fall. At the same time,

Spinning–techniques of production Spinning converts non-differentiated agglomerates of fibres of limited length into ordered, twisted, flexible strands of unlimited length with tensile strength. The morphology of the spinning whorls from the Chalcolithic period (Levy and Gilead 2012: 131-134), Egyptian tomb representations and ethnographic observations from the Old and New Worlds suggest that two spinning techniques were practiced in the Southern Levant in the 5th millennium: spinning supported on the thigh and drop spinning. Representations of 177

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead splicing in their manufacture (Barber 1991: 48, Fig. 5.1; Leuzinger and Rast-Eicher 2011: 540; Shamir 2014: 142). A representation of all phases of textile production from the tomb of Daga at Thebes, Egypt, dated to the early 2nd millennium BCE, illustrates the procedure in action (Vogelsang-Eastwood 2000: 271, Fig. 11.3) showing two splicers working to supply the proto-yarn needs of one spinner. Nota Bene: Splicing is conceptually different from draft spinning, the predominant spinning mode throughout the world. The technique is suitable for long bast fibres but not for wool or cotton. To reiterate, the fibres are stuck end to end forming a continuous proto-yarn and wound up into a ball or skein for spinning. In contrast, in draft spinning the raw fibres are paid out a few at a time into the newly forming yarn so that the ends overlap everywhere and cohesion is caused by twisting together (Barber 1991: 46). In the European sphere the raw fibre is generally wound around a distaff (Hochberg 1980: 12, 33, 51). In the Nile Valley the raw fibre is held in the hand (Crowfoot 1931: PLS. 22, 23). When using the splicing technique a distaff is not used. The yarn used in the manufacture of Ghassulian textiles is spliced. Therefore, the use of distaffs in the production of Ghassulian textiles as recently proposed by Langgut et al. 2016: 984, Fig. 7) must be rejected. Ethnographic figures for rates of production for spinning flax with a handheld spindle do not exist. One or two exist for wool. Rates of production per hour were documented in Peru (n=80) extrapolated from timed 10 minute periods. Spinners, spinning since childhood, spinning wool on a low whorled spindle without a hook, spin at the rate of 62 m per hour (mph) (Bird 1968: 14). Spinners from Bukowina, Ukraine spin on a spindle without a hook at the rate of 60-80 mph (unknown n) (Schwarz 1947: 2138, 2165) and Carrington-Smith, a revivalist spinner, (1975: 82 n. 3) spins wool with a hook and a distaff at the rate of 50 mph. A recent Danish study reconstructing Viking life ways gives a figure of an average for flax of 56 mph spun by four skilled spinners on spindles with low position whorls and with a hook (Ejstrud et al. 2011: 62). To reiterate, the use of a metal hook, as recorded in literary sources from the Classical period (Barber 1991: 68) increases the rate of yarn production by ca. 50%. Reduced to the more visual and concrete; it would take a spinner 1 hour to generate a maximum of 23.373 m of plied yarn sufficient to weave a textile of a maximum dimension of 1 m x 1.17 cm, with a low thread count of 10 threads in the warp and 10 threads in the weft, based on rates of 62 mph for single yarn and 95 mph for plied yarn (Bird 1968:14) (calculation Roi Levy 2016), To reiterate, all Chalcolithic and Early Bronze Age textiles are plied.

Figure 5. Drop spinning, high whorl spindle, Iran after Hochberg 1980, 63 courtesy of Yuval Shach.

the spinner draws out fresh fibres from the mass, until the tip touches the ground, and the new yarn is wound onto the shaft. The process demands good motor coordination. Spindles with low whorls are set in motion by a short twist or twirl with forefinger and thumb, in the manner of spinning a top, of the upper end of the shaft and then rotate in free fall. The fibres are attenuated by the hand of the spinner and by gravity resulting in long lengths of strong yarn c. 50 cm (Chang 1995: 359) with dense uniform twist (Kissell 1918; 32). Drawing out/attenuating and twisting is carried out simultaneously thus speeding up the production process (Kissell 1918: 30). Associated with early flax spinning and probably also with nettle is a preparatory process called splicing. This process entails gluing the long fibres together end to end with water or saliva, setting the join with pressure, inserting loose twist after which the prepared fibres, proto-yarn are ready for spinning. Linen textiles from the submerged sites from eastern Switzerland dated to the 4th millennium BCE and the earliest Egyptian textile from 5th millennium Fayum BCE and also the Chalcolithic textiles of the Southern Levant attest to

Drop spinning constitutes a Great Leap Forward. Apart from the improvement in the quality and tensile 178

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period strength of the yarn is the technique’s potential duality. Spinning in this mode can be conducted sitting, standing or walking as opposed to thigh spun yarn, or grasped spinning or supported on the thigh spinning which can only be conducted in a stationary mode (Kissell 1918: 18, Fig. 1, 22, Figs. 2-3, 24, Fig. 4, 29, Figs. 6-8). In societies of both the New and Old Worlds in which herding, but not of cattle, is a component of the economy, spinning is carried out in tandem with herding (Barber 1991: 4; Carrington-Smith 1975: 8 n. 1; Frödin and Nordenskiöld 1918: 34, 47). Drop spinning, a skill acquired in childhood, becomes an automatic skill, and it can be integrated with any activity with low physical energy requirements (Hardy 2008: 275; Hochberg 1980: 50-51).

dictate the loom size, which will remain, pegged out to its full extent until the completion of the textile. Recent local ethnographic observations attest to looms 12 m in length (Weir 1976: 40). Under these technological limitations, weaving can only take place during the summer.

The earliest loom type attested in southwest and the Nile valley is the frameless, horizontal ground loom appearing on a 4th millennium BCE cylinder seal from Susa (Le Breton 1957: 106 Fig. 20.20) and in a detailed textile scene on a red polished bowl from a Predynastic female burial at Badari, Egypt dated to the mid-5th4th millennium BCE (Brunton and Caton-Thompson 1928: 51, 54, PL. XXXVIII.70k; Roy 2011; 23) (Fig. 6). The Badarian loom features two transverse bars held in position by four stakes with warp threads running between, a shed rod, a heddle rod, and a beater. The weft rows extend beyond the left side of the warp indicating the presence of a weft fringe, a characteristic of early Egyptian textiles (Barber 1991: 83 n. 2: Landi and Hall 1979: 143). Two figures, lacking sexual attributes, stand to one side of the loom next to a transverse beam, one at each end. Suspended from the beam are looped, parallel, double weft lengths ready for weaving (Levy and Gilead 2013: 33). Double weft lengths are consistent with the weaving needs of a textile with a weft fringe. Double weft lengths and weft fringes were also observed on the two major textiles from the Cave of the Warrior, Wadi el-Makkukh (Schick 1998). Neither archaeological representations nor ethnographic observations attest to a windup mechanism for the woven cloth on this type of loom. Thus, the size of the intended textile will

Ethnography attests to the continued widespread use of the horizontal ground loom from Morocco to Baluchistan (Harvey 1996: 54-59; Picton and Mack 1989: 54-59). The basic morphology and principles have remained unchanged. However, there are changes in technology, raw materials and manufacturing practices. The cylinder seal from Susa (Le Breton 1957: 20.20) depicts, apart from a loom, also a warping frame. Textile scenes from Egypt depict configurations of warping pegs on walls (Barber 1991: Fig. 3.8). In both cases, the total warp needed for a given textile is wound in circuits and transferred as one unit to the loom. Recently observed, local laying-in of the warp is more energy intensive. Two weavers sit behind the loom bars, and a child runs between them carrying a ball of yarn. Each warp thread is laid in individually and in succession (Crowfoot 1945: 40). The heddle in Egyptian representation moves vertically opening the countershed for the passage of the weft (Roth 1918: 45). Recent horizontal ground looms looms feature a fixed heddle, in which one shed, the countershed, is permanently open with the heddle rod held on rocks. It moves laterally progressing from the beginning of the loom to the end with the weavers in its wake (Weir 1970: 18). In the above cited representations from both Mesopotamia and Egypt the weavers sit on either side of the loom passing the weft from side to side. Local observations attest to weavers, sometimes three sitting on the loom on the woven cloth and weaving from that position passing the stick spool with the weft yarn from one to the other (Schick 1998: PL. 3.12). Recent textiles woven on the horizontal ground loom are rugs or tent awning, which are made of coarse wool or hair as opposed to earlier flax. It is unlikely that spun flax warp yarn of clothing quality could withstand the sustained weight of three adults.

Figure 6. Horizontal ground loom on ceramic bowl Badari, Egypt from Brunton and Caton-Thompson 1928 PL. XLVIII/70k.

The Badarian bowl provides technological details of loom construction only. Both the Mesopotamian cylinder seal, in itself a marker for social hierarchy and economic structures, and the 2nd millennium Egyptian textile production representations from elite tombs provide additional socio-economic information. Streamlined textile production takes place in atelier-like environments with capital for investment in equipment for greater efficiency. In contrast, ethnographic observations attest to initiative and production at the household or extended family level using traditional time consuming methods with equipment demanding minimal capital outlay to supply pragmatic household needs or as gifts for social situations such as marriages.

Weaving – the horizontal ground loom

179

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Spinning equipment of the Southern Levant–spindle whorls

I have examined all the Chalcolithic spindle whorls held by the Israel Antiquities Authority and also those from the Jean Perrot excavations of the Beer Sheva valleys sites of the 1950’s, but excluding the assemblage from Gilat, which was not accessible,

Unequivocal, albeit small assemblages of ceramic and sherd whorls are known from the Yarmukian (Garfinkel 1999: 29). Fragments of spindle shafts or intact spindles are rarely encountered in the archaeological record. A case is recorded from an Early Bronze Age stratum at Tell Abu Kharaz (Fischer 2008: 353, Fig. 315). The function of the whorl, a flywheel, is to enhance the twist energy inserted by the hand of the spinner. The morphology, i.e. height to diameter ratio and weight distribution will determine, to a large degree, the character of the yarn. The ultimate determinants are the spinner, her skill, yarn vision, and motivation (Kania 2013: 26). If when spinning, in the drop-spinning mode, the whorl is too light in relation to the fibres being spun, it will not magnify the initial twist energy inserted. If the whorl is too heavy, it will cause the yarn to break. The fibres will slip past one another before sufficient twist has been inserted. A whorl with a low height to diameter ratio will spin faster, inserting more twists per unit of length than a whorl with the same weight and a high height to diameter ratio. Yarn with a higher number of twists per unit of length is stronger. It is also harder than yarn with a lower number of twists. There is an observed ethnographic correlation between large, heavy whorls and coarse or long fibres and light whorls and short and/or slippery fibres (Barber 1991: 51-54).

Of the 381 spindle whorls examined from 22 sites, 320 were selected for statistical analysis. Those that fell within the 150-10 g range were included. One hundred and fifty grams is considered the upper limit for a spindle whorl (Barber 1991: 52), and 10 g is considered the minimum weight for effective spinning (Carrington-Smith 1975: 80). Excluded were blanks or those with incomplete perforations, those which constituted less than 40% of the original artifact, those with gross dissymmetry, and those that had a perforation with a bent axis. Also excluded were those with perforations at the edges suggesting vessel repair and those with more than one perforation suggesting a sieve fragment. The whorls were examined with a handheld magnifying glass. The diameters, width and perforations, internal and external dimensions, were measured with manual calipers. Weights were taken, and each whorl photographed. Raw materials were recorded, as were striations of manufacture and use attrition. The artefacts were examined for decoration or polish. The whorls were classified into five major material types with a clear correlation between the energy invested in manufacture and attention to finish and raw material type. Recycled body sherds constitute 47% of the assemblage. The whorls are discoid in form, measuring c. 4 cm in diameter and weighing c. 20 g and ethnographically associated with drop spinning (Hochberg 1980: 8, 63). The artefacts are not symmetrical but roundish. Frequently the contours are irregular and only sometimes ground. The perforation is approximately central, reamed with a bi-conical profile with an internal dimension c. 50% less than the external. They are never intentionally decorated. However, they sometimes feature paint from the original vessel. Frequently potential whorls are encountered with incomplete perforations. Large numbers are also broken in half through the perforation. The whorls often evidence poor choice of raw material, e.g. sherds with large, sharp flint temper. It is obvious that little effort was expended in their production. Timed experimentation in whorl manufacture including reaming with a flint discard, not a tool, took c. 30 minutes (Levy 2006: 164-165). The high incidence of partially completed whorls and low investment in their manufacture appears to indicate an ad hoc functional tool with low cultural value. Similar sentiments were voiced by Médard (2007: 368) regarding the 5th millennium whorl assemblage from Egolzwil, Switzerland.

Ideally, spindle whorls should be round for rotational efficiency. Ideally, the perforation should be vertical and central for maximizing the length of rotational spin and inhibiting wobbling. Ideally, whorls should be smooth on all faces so that fibres or newly formed yarn does not snag and break on protrusions. Ideally, the weight should be evenly distributed around a central axis. Ideally, the perforation should be a tapered cone, drilled from one face, so that the whorl sits snuggly on the shaft (Verhecken 2010, 2013). A few whorls of the Chalcolithic period approximate the ideal. Most do not. Most Chalcolithic sites evidence spindle whorls in very limited numbers. Three Ghassulian sites located on the desert fringe, Gilat (n=163), Bir es-Safadi (n=161) and Teleilath Ghassul (n=310) attest to a dramatic increase in their number (Levy et al. 2006: 705-738; Levy 2006: 111-123; Mallon et al. 1934: 72-73, PL. 36) (Figs. 7 and 8).6 Spindle whorl presence/absence/figures from early excavations should be approached with caution. Often they were considered to have had a low cultural value in the eyes of the excavators and were not worthy of collection or exhibiting in museums, particularly sherd whorls. Typical attitudes are ‘I suppose that it was Schliemann who first brought the spindle whorls into prominence – a venial sin in his case, but today there is no excuse for wasting time and money on this monotonous and profitless material’ (ref. Alalakh {Tell Atchana] Woolley, 1955: 271); ‘whorls of pierced pottery discs and imperforated discs… were found in all sites in great abundance (ref. the Besor sites) and will not be referred to again’ (Macdonald 1932:5). 6 

180

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period

Figure 7. Representative spinning whorls from the Southern Levant (1) photography by author.

181

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 8. Representative spinning whorls from the Southern Levant (2) photography by author.

182

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period Ceramic whorls, specifically those created for the task, which were spheroid or biconical in shape constitute 18% of the assemblage. The whorls measure c. 4.5 cm x 3.3 cm and weigh c. 65 g. This whorl type, known from Mesopotamia to western Spain (Alfaro-Ginar 1984: PL. III; Voigt and Meadows 1983: 167-170 with references therein), is ethnographically associated with supported spinning on the thigh (Crowfoot 1954; Fig. 273 A). They are symmetrical, the surfaces smooth and undecorated with a straight central perforation occasionally with an extruded collar from the manufacturing process. The whorls are low fired, often with a black interior, and they frequently feature attrition around both ends of the perforation. The attrition is probably the result of aggressive padding inserted to prevent the whorl from working loose from the shaft during spinning. Low fired fabric would be particularly susceptible to attrition. Chalk whorls, 12% of the assemblage, roundish, either flat or lenticular with a somewhat central perforation and a bi-conical profile measure 5 cm in diameter and weigh 58 g, with similar dimensions and weight to the ceramic whorl. The whorls are frequently ill balanced, and they often feature gross manufacturing striae. Despite the ease with which this soft material can be processed, little effort was expended in their manufacture.

each category, but they are observable, salient trends. Decorated whorls have not been identified in any material type. All Ghassulian textiles are produced from 2ply yarn. Theoretically, plied yarn requires a heavier whorl than single yarn, approximately twice the weight. However, bimodality was not observed (Levy and Gilead 2012: 132). Within the Southern Levant and the Nile Valley, plied yarn is generally produced in the supported on the thigh mode (Weir 1976: 34) sometimes with a heavier spindle, but frequently with the same spindle used for drop spinning single yarn. However, the name of the spindle changes when it is used in a plying mode. The observed attitude towards the sherd and chalk spindle whorls is anomalous. The Ghassulian culture is characterized by aestheticism, as evidenced in the finely crafted basalt bowls or the V-shaped figurines, and attention to detail and finish. Furthermore, intensive use of drilling as a decorative motif also characterizes the culture (Bar-Adon 1980: 16-21). Yet all the whorls feature bi-conical perforations, which even with dense packing would have created difficulties in stabilizing them on the shaft. Spinning bowls Spinning bowls, which are ancillary equipment used in the manufacture of improved quality linen yarn, are attested from 3rd millennium BCE Crete (Carrington Smith 1975: 273; Warren 1972:153, 208-9, 700-701, Fig. 9), late 3rd millennium to mid-1st millennium BCE Egypt (Allen 1997:19, 25), in the Southern Levant from two distinct periods 5th-4th millennium BCE (Levy and Gilead 2012: 133, Fig. 4) and mid-2nd millennium to mid-1st millennium BCE (Dothan 1963: 97-100), and in contemporary Japan for spinning nettle fibres (Barber 1991:73, Fig. 2.39). Representations of these bowls with rove issuing from them also appear in textile scenes from tombs of 2nd millennium BCE Egypt and in threedimensional wooden models of textile manufacture placed in tombs of the same period (Allen 1997: 17). The bowls are of limestone or well fired clay, heavy with flat bases, with one to three internal handles of identical size on a single axis and frequently feature furrowing under the handles (Barber 1991: 70-77). Nagel (1938: 188) was the first to identify their function. The bowls contained water, and they performed a dual function. They were the medium for ensuring that the prepared rove was thoroughly wetted before it arrived to the spinner, and they also provided tension during spinning. The spliced rove, either ordered coils by the bowl or balls within the bowl, was led through the water, under the handle and up to the spinner standing and spinning in the drop spinning technique (Roth 1913: Fig. 4; VogelsangEastwood 2000: 272, Fig. 11.3).

Limestone whorls, 18% of the assemblage measure c. 5 cm in diameter and weigh c. 45 g. They are round and symmetrical with a straight or beveled edge and a central perforation with an hourglass profile. Surfaces are smooth and often polished. Basalt whorls, 5% of the assemblage, measuring 5 cm in diameter and weighing c. 65 g are round and symmetrical and well processed despite the hardness of the material. They are highly polished on all faces. Another three items of pink, flinty limestone, smooth and also highly polished fall within the range and morphology of spindle but whose attributes suggest a function other than spinning7 (Levy 2006; 113-122, Appendix B; Levy and Gilead 2012: 131-133). In sum, the observed attributes of the sherd whorls and of the chalk whorls appears to suggest that they were manufactured within households probably by the practitioners themselves. In contrast, the ceramic whorls and those of limestone and basalt feature symmetry and attention to finish suggesting production by skilled hands. The observations are not 100% so in Not all artefacts with a central perforation are spindle whorls even if they fall within the established criteria. Any potential whorl made of exotic material, or material that is difficult to process, or is highly polished, or whose quality of workmanship, particularly symmetry, is far in excess of utilitarian needs, must be evaluated with caution. Polish on all faces cannot develop from spinning. Polish can only develop along the rim, albeit slowly, since the yarn at the rim is stationary during spinning. Polish on all faces can occur as a specific manufacturing process or through handling or rubbing against skin or garments. 7 

Wet or damp spun flax, like all bast fibres, produces superior yarn. Dry spun yarn is fuzzy, hairy, and 183

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead unsuitable for textiles (Hess 1954: 304). Along the length of the fibre (in fact fibre bundles) ends stick out like the split ends of human hair. Water or saliva gum down these wayward ends. During spinning-twisting these ends are incorporated into the main strand, creating a sleek yarn (Hochberg 1979: 44). Four spinning bowls have been recovered from the Southern Levant; one from Neve Ur in the northern Jordan valley, two yet to be published from Abu Hamid in the central Jordan valley8 and a 4th from Bir es-Safadi in the Beer Sheva valley. Neve Ur with a representative Ghassulian repertoire was never excavated. The bowl is a surface find (Perrot et al. 1967: 223, PL. 42.9-10; Reich personal communication) (Fig. 9). The bowl c. 30% of the whole artefact is of biscuit coloured fabric with very fine temper, evenly fired throughout, smooth, very heavy with a flat base and a flaring profile and broken above the internal handle. The base is c. 20 cm in diameter, and the wall is 1 cm thick. A ledge handle, wide, flat and smooth parallel to the base and triangular in form, descends at 900 to the floor of the vessel. Calcareous concretions appear on the bowl particularly under the handle. Therefore, it is impossible to ascertain if furrowing, a characteristic of spinning bowls, exists.9

Figure 9. Spinning bowl, Neve Ur from Perrot, Zori and Reich 1967, photography by Alter Fogel.

others, probably from areas north of the Southern Levant, will appear once their presence in this time bracket is recognized. When published in 1967 (Perrot et al.) and1990 (Commenge-Pellerin) their function was not recognized although Perrot et al. (1967: 223 n.22) observed the morphological similarity between the Ghassulian artefacts and those of the Late Bronze Age housed in the Beit Shean Museum. Bar-Adon (1980: 179) astutely/intuitively suggested that the enigmatic artifact with two loops (handles?) next to the loom in the textile scene on the Badarian bowl (see Fig. 4) was a ‘spinning bowl with a piece of thread sticking out’. The blob next to it was not discussed. However, it is probably a representation of a ball of prepared and spliced rove. The suggestion went unheeded. Today with concrete evidence for spinning bowls from the period, the suggestion would appear to be valid.

As for the second bowl published by Commenge-Pellerin (1990: 66-67, Fig. 22.2), c. 50% of the bowl, is red-brown in colour, heavy with a flat base and a flaring profile, smooth on all surfaces with traces of slip on the inner face and a band of burgundy paint on the inner rim. The bowl, 12 cm in height with a base 20 cm in diameter and 36 cm at the rim, features a strap handle, parallel to the base, which begins from the mid height of the vessel wall and descends at 900 to the vessel floor. The presence of irregular rilling on the inner face indicates that the final stages in the manufacture of the bowl were carried out on a tournette (slow potter’s wheel) (Roux and Miroschedji 2009:155). There are shallow furrows, in three discrete groups on the tapered rim, at 900 to the axis of the handle. Approximately 75% of the base is present. Therefore, it is possible to ascertain that there was a single handle only. This is the case with respect to the 3rd and 2nd millennium BCE spinning bowls from Crete and those used in contemporary Japan (Barber 1991:73).

The fact that their number is very limited vis à vis the number of spindle whorls suggests that other, less sophisticated methods were generally used to impart the requisite moisture. Running the flax strands through the mouth to dampen with saliva and create smooth yarn is a practice widely attested historically and ethnographically (Baines 1989: 26). It is shown in the textile scene from the early second millennium BCE tomb of Thuthotep at Deir el Bersha, Egypt (Crowfoot 1931: 24, Fig. 6). In addition, it was observed as a contemporary practice amongst the spinners of Nahya, Egypt, the village where Crowfoot (1931: 32-36) observed flax processing. The spinners are known in the vicinity, as ‘the women who spin through their mouths’ (Crowfoot 1931: 33, PLS. 22-23). The women from the second village Kirdasseh, where flax is also cultivated and processed for textile purposes, were also reported to spin in the same manner. The Babylonian Talmud (Ketubot 47b) discusses the long-term physical damage of spinning flax and wetting it through the mouth. In addition, European folklore refers to broken lip muscles as a characteristic of flax spinners (Schneider 1989: 177).

These bowls from the Southern Levant are the earliest known examples of spinning bowls. Undoubtedly, Despite long and intensive efforts, I have failed to locate the thesis in which these two artefacts were discussed. A pen drawing in profile without metrics or further information was published by Breniquet (2008: 124, Fig. 29.3). 9  Thanks are extended to Prof. Avi Gopher and Amnon Assaf, the curator of the Museum of Prehistory at Mayan Baruch for assistance in locating the Neve Ur spinning bowl. Thanks are also extended to Ysrulich Reich who succeeded in locating a key to unlock the drawer where the artifact had been cached for so long and to him and his wife for so graciously hosting me at Kibbutz Neve Ur. 8 

184

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period The simplicity and effectiveness of wetting flax fibre through the mouth is compatible with the Ghassulian life style. The use of spinning bowls, a sophisticated concept instrumental in producing fine, smooth yarn, was incompatible with a village economy. The spinner was standing and stationary and capable only of spinning, a single task. On the other hand, wetting through the mouth when combined with drop spinning permitted mobility and multi-tasking (Hochberg 1980: 50-51).

centimetre or so creating a textured self-band. Also recorded is the use of coarse warps at the selvedges and at both ends of a textile to create a pleated effect (Schick 2002: 235). A number of poorly stitched features were observed among the repertoire of the Cave of the Treasure (Bar-Adon 1980:154; Cindorf et al. 1980: 230). Tiny repairs were observed on the shroud from the Cave of the Warrior (Schick 1998: 12-13, Figs. 3.28-3.29). Sewing needles are rarely encountered in all periods. Four bones needle were recovered from Gilat (Grigson 2006: 690, Fig. 13.2 J). The eye of the published needle measures approximately 1 mm. It is possible that a natural foramen was exploited to achieve such small dimensions. There are archaeological analogies from the Aegean and Norway (Hurcombe 2014: 71, Morten Kutschera personal communication).

The textile repertoire A remarkable number of textiles were recovered exclusively from the Judean desert and the Dead Sea basin. These finds emanate not from one specific site but from cave after cave associated with typical Ghassulian domestic repertoires of churns, cooking pots, sieves and whorls and also mortuary contexts (Aharoni 1961: 13, 14-15; Avigad 1962: PLS. 16-17; Bar-Adon 1980: 135, Fig. 1, 178, Fig. 51, 57-58, 191 Fig. 62; Schick 1998: 6-22). It is suggested that they were not manufactured within the caves (Schick 2002: 238: Shamir 2014: 140). Within a cave, there is neither sufficient space nor lighting for weaving. Cave residence in this stage of cultural development in the Southern Levant is anomalous. One may suggest, based on climatic conditions and water regime, that it was winter-spring occupancy by certain family members of the inhabitants of settlements in the Jordan valley or the Arad-Beer Sheva region, who exploited the winter pasture for their flocks. Similar economic strategies are practiced by settled Bedouin communities in the Dimona region (Eldar et al. 1992: 214).

Two cave sites feature exceptional repertoires – the Cave of the Treasure, Nahal Mishmar, and The Cave of the Warrior, Wadi el Makkukh. The Cave of the Treasure, known primarily for the copper artefact cache, is currently dated to the second half of the 5th millennium BCE (Davidovich 2008: 131-135; Gilead and Gošić 2014: 232-233, Table 3). Its highly inaccessible location, even for those ardent in archaeological pillage, guaranteed the integrity of its contents. The cave featured two occupational phases, Chalcolithic and Roman. Stratigraphic confusion prevailed. The Chalcolithic stratum had been grossly disturbed by later Roman occupancy. The site was excavated under the most difficult conditions. Access was by rappelling, and light was provided by emergency lights run from a command car engine. In addition, the excavation was conducted within a thick layer of bird and bat dropping, the accumulate of millennia. To circumvent the problem of stratigraphy, Bar-Adon (1980: 3) defined three strata: Roman, Chalcolithic, and Intermediate. The Intermediate was not only a physical stratum up to 15 cm in depth but also a category to which were attributed artifacts lacking identifiable stratigraphy or typology (Gilead and Gošić 2014: 227).

The textiles exhibit an overall homogeneity in visual attributes and techniques of production. They range from off-white to brown, attributable to differences in micro-loci of preservation or differences in retting conditions. They are neither coloured, apart from a single occurrence from the Cave of the Warrior, nor bleached. Fibres are spliced. Yarn is randomly s spun and s plied yarn (Schick 2002: 231, 238), the result of variability in the cohesive qualities of the flax fibres (Crowfoot 1982: 547). There are no observable differences between warp and weft threads. In later periods warp threads, which are subject to constant tension and friction are harder spun (more twist is inserted during spinning) (Shamir 2014: 149). All are plain weave. In most cases, the number of warp threads is slightly higher than the weft threads. Selvedges (the outermost edges of textiles), prone to attrition, are often reinforced using thicker warps or weaving through paired warp threads (Schick 2002: 224, 230, 234). Warp fringes and occasionally looped and knotted warp tassels are also encountered (Schick 2002: 235). Decorative features are rarely encountered. Bar Adon (1980:162; Cindorf et al. 1980: 230) record a textile with four weft threads of coarse yarns every

Cindorf et al. (1980: 231) working without a referential textile atlas, experienced difficulties in differentiating between Chalcolithic and Roman linen textiles. Both are s spun, both are of plain weave, and both have similar thread counts, a 4000 year textile continuum. However, they did observe that most Chalcolithic textiles were two-ply while most Roman textiles were single ply. Recent, re-evaluation of the textile assemblage using SEM technology has identified the presence of splicing, a characteristic of Chalcolithic yarn but absent from Roman yarn (Shamir 2014: 142). Thus 87 textiles, the majority of the assemblage, are now attributable to the Chalcolithic period, from coarse and loose to dense and fine, with a thread count range of 9-45 in the warp and 7-30 in the weft and a medium-range of 14 x 12 tpc, 16 x 13 tpc and 18 x 14 tpc (Shamir 2014: 150). 185

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead

Figure 10. Schematic of linen shroud, Cave of the Warrior, after Schick 1998 Fig. 3.2, courtesy of Yuval Shach.

A comparative evaluation of the Chalcolithic period textiles with those of considerably later periods evidences similar yarn density although the yarn of the later period is single. Linen, plain weave shrouds from 2nd century BCE Ein Gedi attest to 12 x 10 tpc and 16 x 12 tpc (Sheffer 1994: 9). A plain weave linen fragment from z spun yarn, indicative of European origin, was recovered from the Napoleonic siege trench at Acre. This textile, dated to 1799 CE, preserved as a pseudomorph on a copper button of a French soldier’s uniform, attests to a yarn density of 10 x 10 tpc (Schick 1997: 98).

repertoire (Bar-Adon 1980: 162, PL. 35.4: Cindorf et al. 1980: 229). The funerary outfit compromises three archaeologically intact textiles, with selvedges and ends – a shroud, woven as a single piece measuring 7 m x 2 m, (Fig. 10) a wrap-around kilt 1.4 m x 0.9 m and a sash measuring 2 m x 0.15-0.25 m. The textiles range from off-white to dark red-brown resulting from contact with body fluids and ochre used in the funerary rites (Schick 1998: 7, 15). Technologically the bodies of these textiles are, like the others of the Judean desert, exhibit spliced fibres, predominantly s spun and s plied yarn, plain weave and similar medium range thread counts. The differences are in the dimensions and state of preservation, the elaborations of the selvedges and ends, and the innovative use of coloured brown-black10 yarn (Koren 1998: 104).

As stated earlier, linen/flax is the only fibre type attested in the Chalcolthic period. However, the excavator attributed woolen textiles to all three strata (Bar-Adon 1980: 153, ILL. 26; Cindorf et al. 230-231). All are single ply, a feature alien to the spinning traditions of the Chalcolithic – all are two ply, but characteristic of the local Roman period spinning traditions. One fragment was z spun in the weft, a spinning direction alien to Chalcolithic traditions – all are s spun. A sample(s) of the woolen textiles was sent by Tamar Schick to England for analysis by Michael Ryder, the foremost authority on archaeological wool. He defined the wool, on morphological grounds, as late (Tamar Schick, personal communication). The laboratory report cannot be located. As Gilead and Gošić (2014: 229) suggest, it would be desirable to conduct C14 tests to resolve the issue.

The shroud features three coloured black, outermost warp threads at the right selvedge, woven as one unit and a narrow stripe, 11 warp threads wide, immediately before the outermost warp threads of the left selvedge (Schick 1998: Fig. 3.22). The colourant, of organic origin, has not been positively identified. However, it pertains to a group which includes resins, gum Arabic (acacia gum), collagen or bitumen, all locally available (Koren 1998: 104). The weft was not woven from continuous weft yarn, as generally observed, but from double weft lengths. Two ends of equal length of the double weft yarn were threaded through the outer four warps of the left selvedge and tied in a quasi-buttonhole stitch thus simultaneously securing the selvedge and forming a decorative weft fringe 10 cm in length (Schick 1998: 11-12, Fig. 3.26) (Fig. 11). Each unit of the weft fringe was created in sequence during weaving. It is suggested that this feature, which is present in Middle Kingdom (c. 2000-1600 BC) Egyptian textiles (Vogelsang-Eastwood

The Cave of the Warrior features an extraordinary mortuary repertoire, including a linen shroud of dimensions unprecedented in this time bracket in southwest Asia and the Nile Valley (Schick 1998: 20, 127). The cave assemblage dated to the first quarter of the 4th millennium cal. (Jull et al. 1998: 110-111) evidenced an undisturbed mature male burial with accompanying grave goods including leather sandals (Schick 1998: 34, Fig. 8.1). Three additional textiles, recovered from an interior niche, attributed to an earlier burial are dated to mid-5th millennium BCE (Jull et al. 1998: 111). Textile D from the internal niche features similar technological attributes spin, weave, selvedge type and thread count to that of textile 61-290/7-a from the Nahal Mishmar

This is not conventional dyeing – there is no chemical bonding. The colourant was mechanically impressed into the plied yarn prior to weaving. 10 

186

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period

Figure 12. Shroud, detail of decorative band and warp fringe from Schick 1998 PL. 3.7 courtesy of Orit Shamir and Israel Antiquities Authority, photography by Clara Amit.

Figure 11. Schematic of weft fringe, shroud, Cave of the Warrior, after Schick 1998 Fig. 3.26 courtesy of Natanel Levy.

The third textile, a sash, also a standard item of male attire in Egyptian and Mesopotamian cultural spheres (Vogelsang-Eastwood 1993: 86), measures 2 m x 0.150.25 m. Like other textiles of the Chalcolithic period, it is manufactured from s spun and S plied yarn, woven in plain weave from continuous weft yarn with a simple selvedge. At both ends of the sash are two rows of hand inserted countered weft twining creating a chain-like effect in low relief (Schick 1998: 15-16, Fig. 3.48; Shamir 2014: Fig. 7.5) (Fig. 13). The technique, also used as a fabric technique, is recorded at PPNB, 8th millennium BCE, Nahal Hemar (Schick 1988: 37, Fig. 16). The sash terminates in twisted and inter-looped tassels measuring ca 17 cm. The random, non-structured variations in the width of the sash along its length suggests that this was not a product of the same manufacturing hand as the other two textiles. The ability to retain an equal width is a primary marker for textile skill.

1992: 29), was not primarily a decorative effect. Rather, it was a technical solution to avoid loss of width when weaving wide textiles (Barber 1991: 151; Schick 1998: 21). Two identical decorative bands, interlaced by hand and not woven with heddle and shed rods, appear at each end of the textile. They feature alternating, contrasting narrow bands of naturally coloured yarn and coloured black yarn in basket weave (paired weft threads over paired warp threads), half basket weave (single weft thread over paired warp threads) and triple weft threads over paired warp threads. At the extreme ends of the textile are 180 equally spaced tassels, essentially elaborated warp fringes, 18 cm in length. While cutting the textile off the loom the individual warp threads were twisted around the outer weft threads to hold them in position once the source of tension had been removed. Subsequently groups of warp threads were twisted in a z direction (to the right), and then two groups twisted together in an s direction (to the left) and the ends of the tassels were knotted in an overhand knot to prevent unravelling (Schick 1998: 9-11, Figs. 3.13, 3.20) (Fig. 12). The second textile, found within the shroud with the deceased, is morphologically and technologically similar to the shroud, although less elaborate, with decorative bands at both ends in diverse thread combinations and contrasting dyed yarn, a looped and tied weft fringe and a fringe of twisted tassels at one end. The dimensions, 1.4 m x 0.9 m, suggest a wraparound kilt reaching mid-calf, a standard item of attire for males, of all social levels, from both Egypt and Mesopotamia (Schick 1998: 21; Vogelsang-Eastwood 1993: Fig. 5.4-6; Breniquet 2013: 4, Fig. 1. 2).

Figure 13. Sash, detail of countered weft twining and warp tassels, Cave of the Warrior, from Schick 1998 PL. 3.9, courtesy of Orit Shamir and the Israel Antiquities Authority, photography by Clara Amit.

187

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Textiles – mode, context and agents of production

Hebron Hills 50 kms distant from the Beersheva Valley sites. They were the product of household initiative undertaken by kinswomen, primarily for wedding needs (Vogel 1989: 85-86).

The finding of this extraordinary shroud, the like of which is unknown, in this time bracket, amongst contemporary societies of the region raised questions as to how could such a simple, peasant-type society (Gilead 1988: 436) as the Ghassulian produce such a finely crafted textile of mammoth dimensions. Could it not be the product of a more sophisticated society beyond the borders of the Southern Levant? The question was addressed at length, and the textile was broken down into its manufacturing components. In addition, the person hours in production were calculated, and a model proposed (Levy and Gilead 2013: 42) (Table 1).

To reiterate, splicing the first stage in linen production entails joining end to end the flax fibres with water or saliva and sealing the join with pressure. The process lacks ethnographic analogies. Nevertheless, it can be stated with a degree of assurance, that it is a year round activity conducted while seated during the hours of daylight for it is visual and tactile. It was an activity suitable for a skilled older woman of the household not in the fields or out with the herds. Splicing is a critical stage in early linen, textile production; it determines its quality. Draft spun yarn (the standard method worldwide) can be spun and re-spun ad infinitum to reduce the dimensions or improve the uniformity of the yarn. This is not possible with spliced yarn. Once the splice is inserted and the fibres gummed end to end no further adjustments are possible. It is a fait accompli.

Spindle whorls, the lowest archaeologically visible rung of textile production, were recovered from the Beer Sheva valley sites occurs within houses, courtyards, subterranean structures, fill and middens (Levy 2006: 121). There is no evidence for concentrations or intentional storage. Likewise at Gilat, where the excavator proposed the presence of a sanctuary with an attached textile atelier, there is a similar pattern of dispersion (Levy et al. 2006: Table 12.25). These factors suggest that yarn manufacture was conducted within households for domestic consumption. Textiles of analogous dimensions to that of the shroud, albeit of wool, also woven on horizontal ground looms, were produced in the late 20th century in the hamlets of the

Spinning is a year round activity, conducted by young girls of the household in the drop spinning technique while herding their sheep and goats (Franquemont 1986: 311). Ethnographically, plying is conducted while seated in the supported on the thigh technique. Herding is a leisurely activity, and constant spinning does not interfere with it. Analogous behaviour was

Length of yarn (m) or number of items

Rate of production (meters per hour)

Number of working hours per day

Work-person days

Splicing*

82,6001

31

6

444

Spinning

82,600

62

8

167

Task

Plying

41,300

95

8

54

Winding yarn off spindle

123,900

672

8

23

Warping**

24,500

2150

4

9

Splicing leash yarn

2,1003

31

6

11

Spinning leash yarn

2,1004

62

8

4

Plying and replying leash yarn

1,575

95

8

2

Preparation of leashes

1,750

60

6

5

Insertion of leashes

2

1,750

30

6

10

Weaving body of textile***

41,3005

62

6

79

Knotting weft fringe****

4,2006

60

6

58

72

1.33

6

9

3607

12

6

5

Decorative bands Twisting, plying and knotting warp tassels Total

880

Notes: 1 See text; 2Based on 3,500 warp threads x 7m; 31750 alternate warp threads x 0.3m x4ply; 4 2100*3/4 (ply and reply yarn); 5Spun yarn divided by 7 (ratio of spinners to weavers); 6 Double weft threads; 7180 tassels in each warp fringe; *Splicer’s rate of production half that of spinner’s; ** Warping requires a crew of three people; *** Weaving requires a crew of five people; **** Knotting the weft fringe is done by one person while the other four members of the crew sit idle awaiting its completion.

Table 1. Person-power labour days in shroud manufacture, Cave of the Warrior (after Levy and Gilead 2013, 42)

188

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period recorded throughout the wet winter of 1965 in the hills above the Dead Sea (Carrington-Smith 1975: 8 n. 1). I suggest that the spliced flax was wound up in a ball, and subsequently soaked overnight, as in the experiment with local wild flax conducted by Abbo et al. (2014: 4) and carried out to the grazing grounds wrapped in a skin to retain the moisture. The spliced flax was ‘spun through the mouth’ and wetted with saliva, and the spun yarn was smoothed between thumb and index finger as observed by Crowfoot (1931: 34). An enzyme in the saliva reacts with the flax forming a very effective glue and hence smooth yarn (Barber 1991: 49 n. 7).

female, child snatching demon, is depicted holding a spindle and weaving comb (Wiggermann 2000: 235, Fig. 4.44). Throughout the ages, females of high and low status have spun yarn and woven cloth for domestic consumption and communal needs, including the Virgin Mary, the mother of God, depicted spinning in the Annunciation scene from the 11th century illuminated Codex Aureus of Speyer Cathedral and Queen Victoria, Empress of India (Hochberg 1979: 32, 1980: 20). It would be perverse to suggest that the Ghassulian culture was otherwise. In sum, textile production was a domestic activity, the domain of the females of the household. Flax fibres were spliced throughout the year in daylight hours by competent females, between household chores and wound up into balls of proto-yarn ready for spinning. The proto-yarn/rove was spun and plied throughout the year in the drop spinning technique by young girls while herding their flocks. The yarn, continuous and not double weft lengths like the shroud of the Cave of the Warrior, was woven during the summer months, agricultural down-time on horizontal ground looms into plain weave textiles with simple selvedges. Rectangular pieces of cloth cut off from the loom were conceived of as finished garments. They were primarily worn as is, wrapped or draped around the body held in position by pinning or belting. The limited evidence of sewing indicates that it was used primarily as a strengthening or repair technique but not for creating new garments. Manufacture of lengths of cloth, even those as large as the shroud, woven from strong linen yarn capable of withstanding sustained tension and friction during weaving, was well within the organizational abilities and technological capabilities of the female members of the Ghassulian. Their seasonal manufacture, within households, with ethnographic analogies of all phases apart from splicing but 50 km distant from the Cave of the Warrior, was seamlessly integrated into their agro-husbandry economy. The Ghassulian, despite its predominantly rural aspect and lack of overt hierarchy, was at the cutting edge of a suite of innovative technologies. The manufacture of textiles in hitherto unknown dimensions was but one aspect.

Weaving in the southern Levant is a summer activity carried out after the crops have been harvested and processed and a long period of dry weather is assured (Vogel 1989: 76). Schick (1998: 20) proposed that the textile was woven on a horizontal ground loom more than 2 m wide and raised c. 10-20 cm above the ground. Ethnographic observations from the Hebron hills attest to three women weaving in unison sitting on the already woven part of the textile (Schick 1998: Fig. 3.12: Vogel 1989: 85). A model was proposed for the production of the shroud from The Cave of the Warrior whereby two skilled female weavers sitting on either side of the loom to maintain the width and to insert the weft fringe with three small children sitting on the woven cloth passing a tiny ball of double weft yarn through the open shed from one to the other (Levy and Gilead 2013: 39). If their combined weight put too much pressure on the warp threads, small logs covered by skins could have been placed below their buttocks. Such procedures are practiced by modern mat weavers of Egypt and Persia interlacing on horizontal ground mat looms (Crowfoot 1933: 94, Fig. 6; Wulff 1966: 220). Periodically the tension is adjusted via ropes attached to stakes. Linen as opposed to wool has little elasticity, and loss of tension may have been less of a problem (Hess 1954: 176, 303). The earliest representations of textile production in southwest Asia and the Nile valley depict woman at the tasks (Barber 1991: Fig. 3.4; Vogelsang-Eastwood 2000: Fig. 11.3). Textual sources from Mesopotamia refer to captive female labour and bonded women in large-scale textile workshops with male overseers (Waetzoldt 2008: 112). Egyptian textile scenes from the Middle and New Kingdom tombs depict female workers as do as do Classical and Medieval representations (Hochberg 1980:11, 20, 41, 50-51, 55-56). So strongly is female labour associated with textiles, particularly spinning the lowest rung in production, that Hittite spells and rites to restore male potency involve holding and discarding spindles (Singer 2009: 206). In Biblical traditions (Samuel II: 3.39) a curse on the clan of Joab, of a magnitude equal to hunger or death by the sword, is that the male line will become spinners (Cassutto: 2008: 69) and even Lamaštu, the Assyrian,

Clad and shod – a new way of being A considerable number of frescos, all multi-layered, of total wall dimensions were excavated at Teleilat Ghassul from all chronological phases and across the site, by various expeditions from the 1920s until 1995. They are painted in a variety of styles, featuring geometric, zoomorphic and anthropomorphic motifs and scenes, in red, white, black and yellow mineral paint (Schwartzbaum et al. 1980: 3, 5). They were recovered from buildings of dimensions and artifact content character, which did not differ significantly from domestic units. Coloured plaster was also recovered 189

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead from a sanctuary, but motifs were not discernable. All the frescos had been significantly damaged by seismic activity. It was possible to consolidate and save a few fragments; most of them disintegrated. The evidence is primarily poor photography and interpretative watercolours painted at the time of discovery. Gilead (2002: 117-120) argues that the frescos are the works of shamans, created in their loci of activity or residence, recording, at a later date, what they had seen/ experienced during their out of body, altered state of consciousness, ecstatic experiences. Drabsch (2015), discussed below, offers an alternative reading.

descending height, wearing horned masks, in a formal, ceremonial situation, all facing left, with their arms bent at the elbow. The first and largest figure carries aloft a long narrow hockey stick-like element flaring at one end. Two artefact types of somewhat similar form (Bar-Adon 1980: 91#125, 105#156), manufactured specifically for ritual or ceremonial use (Sebbanne 2014: 128), were recovered from the Nahal Mishmar copper hoard. The identical position of the elbows of the two smaller figures suggests that they had also held liturgical artefacts. The gowns are all narrow with fitted sleeves lacking hem elaboration or fringes. The gowns of the two smaller figures are red, and they extend to the mid-calf. One features white stripes from the elbow to the wrist. The largest figure wears a multicoloured gown with multiple vertical lines of zig-zag on the upper part of the gown with a horizontal zig-zag line across the shoulders. The one extant sleeve shows a single line of zig-zag from shoulder to elbow and transverse lines from elbow to wrist (Cameron 1981: 14, Frontispiece). Most of the lower part of the gown is missing. A section in the middle of the gown features an amorphous patch with a grid of parallel vertical lines intersected by parallel horizontal lines. A similar, narrow, rectangular, grid configuration descends at an angle of 300 scarf-like from the smallest figure.

The Processional Frieze/The Procession (The Hennessey Fresco) (Fig. 14) excavated in 1977 was successfully reconstructed with UNESCO aid. The polychrome fresco measuring 4 m2 (Schwartzbaum et al. 1980: 1), is dated to the third quarter of the 5th millennium (Gilead 2003: 223). It depicts a line of three gowned individuals of

Ceremonies with individuals wearing horns and masks, characterizes shamanistic behavior, with the participants entering altered states of consciousness induced by hallucinogens, and/or intensive rhythmic music or dancing. Multi-cultural research attests to a series of re-occurring visual, entoptic phenomena experienced by the participants –shimmering, incandescent, zig-zags, parallel lines and grids accompanied by varied and saturated colours (LewisWilliams and Dowson 1988: 203). All these entoptic phenomena appear in the Professional Frieze. Since it is argued that the Processional Frieze was generated by an experience in an altered state of reality (Gilead 2002: 117-120), I will now evaluate the clothing from a technological textile aspect. Analysis of the ceremonial robes of Teleilat Ghassul The gowns of the two smaller figures feature straight transverse hemlines. This is indicative of a textile and not a skin. Textiles are rectilinear, and whether worn on the long axis or the short axis the hem will be straight. In contrast, skins feature an upcurving hemline drooping at the corners; the body and legs of the animal (Burnham 1973: 2; Heckett 2008: 209) as observed on the rock paintings at Cogul Lerida, Spain worn by clothed dancing females encircling an ithyphallic male (Obermaier 1925: PL XII; Sollas 1924: 405). The gowns of the two smaller figures are red, and the gown of the larger figure has a red and black background with dense lines of white zig-zag. The only fibre used for textiles

Figure 14. Fresco, The Procession, Teleilat Ghassul after Cameron 1981, frontispiece, courtesy of Yuval Shach.

190

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period in the Chalcolithic period was flax (Levy and Gilead 2012:127). Flax is particularly difficult to dye. Only with the advent of aniline dyes in the 1800s CE were the problems fully overcome. Linen clothing was worn in natural tones in Egypt and Europe, enhanced by natural bleaching in the sun on grassing greens or with the use of natron (Barber 1991: 15). However, since the palette of the representation maker was limited to four colours only, this may be evidence of artistic constraint and/or accommodation. Both white and yellow paint was used in the fresco. Although they approximate the colour of linen, they were not selected. Linen gowns dyed red or black do not reflect reality. Multi-coloured zig-zag motifs and parallel lines, both entoptic phenomena, are major motifs of the gown of the largest figure. To execute the zig-zag motifs the use of multiple heddle rods are required, inconsistent with weaving on a horizontal ground loom which has one heddle rod only, or tapestry weaving on an upright loom. In order that patterning be visually affective, there must be at least two contrasting colours. To recall, the shroud from the Cave of the Warrior features c. 240 m of coloured yarn out of a total of c. 41,300 m of spun yarn, and that is unique (Schick 1998: 8, Fig. 3.15). Such a textile as worn by the largest figure cannot be woven on a simple horizontal ground loom with a single heddle rod. Neither the requisite weaving nor dyeing technologies were present at this early stage of textile manufacture to achieve such results. Neither are there analogies of such motifs nor dyeing amongst the Egyptian textile repertoire, despite its magnitude.

and the Teleilat Ghassul clothing representations? It is theoretically possible that the zig-zag motifs were painted onto the surface of the textile robes. However, to date the only known painted textile, from the wider region, was recovered from a mid-4th millennium BCE burial at Gebelein, Egypt. The motifs are not geometric. They portray a riverine hunting scene (Read 1964: 129). Could not the motifs, repetitive geometric motifs, have been executed with a stamp seal carved in low relief, dipped in a colorant and impressed on the textile? Despite the many textiles recovered (Shamir 2015: Table 1), evidence of such activity has not been attested. Is there a problem of perspective? Is it possible that the zig-zag elements are not an integral element of the textile but jewelry of vegetable origin worn over the gown? This is possible for the vertical zig-zags but not for the transverse zig-zags across the shoulder. Moreover, the zig-zag motif on the gown is not an isolated occurrence, but also appears as a component of a suggested architectural feature from the same fresco (Hennessy 1982: 56) as a component of the Star Fresco, the Tiger Fresco and as the major component of the Zig-Zag Fresco (Cameron 1981: 5, 9, Figs. 4, 6; Drabsch 2015: 112, Fig. 136). Thus, it seems that the zig-zag motif is not an integral component of the gown. I argue, following the research conducted by LewisWilliamson and Dobson (1988) and Gilead (2002) that the fresco was created after the ceremony, subconsciously employing the principal of integration, decoded the entoptic zig-zag, observed during an altered state of consciousness, against a stored, referential mental database and effected a fit, imposing the zig-zag patterns on the gowns. This specific configuration has ethnographic analogies in rock art (Lewis-Williams and Dowson 1988: 203, 212, Fig. 4).

The grid of parallel vertical and horizontal lines, reminiscent of warp and weft, the basic structure of woven cloth, appears in the centre of the lower part of the gown of the largest figure and a section associated with the smallest figure. It is conceivable that during an out of body experience that the very essence of the cloth was perceived, but it also possible that this is an element of an underlying layer. We must recall that the fresco was reassembled from fragments (Schwartzbaum et al. 1980), and both Cameron (1980: 13) and Garfinkel (2003: 281) question elements of the reconstruction. The three gowns feature narrow fitted sleeves. To fit a sleeve requires a degree of tailoring skill. Several cobbled seams and a hem or two were recorded in the total Judean desert repertoire (Bar-Adon 1980: PLS. 27, 31, 36.3, 39). The observed sewing skills do not appear sufficient for the task. Burnham (1973: 30) observed that in antiquity weaving skills far outstripped sewing ability. There exists a long tradition throughout southwest Asia and the Nile valley of wearing clothing that is simply a textile taken from the loom, as is, and pinning, draping or belting it about the body (Barber 1994: 133-134; Vogelsang-Eastwood 2000: 282).

In sum, there exists little correlation between the actual textile findings of the Judean desert, including the most complete assemblage, the Cave of the Warrior, c. 12 km distant from Teleilat Ghassul, and clothing representations on the fresco. However, artistic representation, even without recall from an altered state of consciousness is not photographic veracity, and license is a recognized phenomenon. The core element is that all three participants and the said Warrior wore woven garments covering a large portion of their bodies. The correlation is conceptual rather than actual. Drabsch’s recent study of the frescos sheds new light and perspectives on the nature of ritual practices and social organization at the site. She argues that the reconstruction of the Procession is both erroneous and compressed. On-site photographs, at the time of discovery, were not taken. In her reconstruction, there are eight, not three, ranked figures, of a priestly class, three robed and five naked acolytes, three of which are equipped with wing-like appendages, similar to

How does one explain the technological incongruity between known textile techniques and capabilities 191

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead the scarf-like artifacts in the UNESCO reconstruction (Drabsch 2015: 119, Fig. 253). All are wearing masks and carrying liturgical artifacts, progressing towards a sanctuary. She reads the fresco as a visual record of an actual ritual event. She reconciles the problematic domestic aspect of the loci of the frescos redefining them as lineage houses, not specifically constructed as such but as evolved domestic units. The multi-layering of the frescos is explained as a physical reflection, documentation, of a sequence of significant events related to specific lineages. She suggests that the time, expenditure and skill invested in their production holds the potential for social differentiation. However, the reconstruction, primarily concerned with aspects of the cult, does not discuss technological aspects of textile manufacture and dyeing. Consequently, it thus does not significantly change the above observations on the zig-zag patterning and technological aspects of textile production apart from the suggested use of black gloves by the largest figure of the Procession (second figure according to Drabsch’s reconstruction) (Drabsch 2015: 158). Gloves require specialist tailoring in their manufacture. Such precision and skill was not observed on Ghassulian textiles.

that had been subject to vegetable tanning ensuring their preservation (Schick 1998: 34, 37). The sandals, which are constructed from two layers, an inner and outer sole, connected one to the other by interlocking slits and tongues, featured a cap to protect the big toe and a counter to protect the heel (Ashkenazi 2008: 7678). Thongs joined the cap to the counter where they were threaded through eyelets and subsequently tied around the ankles (Schick 1998: 36). There was no differentiation between left and right foot (Fig. 15). Representations of footwear, the oldest known, worn by a dignitary, appear at Teleilat Ghassul on the fresco ‘The Notables’ excavated by Mallon in the 1920s. The fresco, originally 4.5 m by an estimated 1.5 m collapsed during an earthquake. Only the lower 50 cm survived. It is a representation of a ceremony involving at least seven individuals, only feet and legs, six in a line orientated to the left facing a multi-rayed luminary. The first two individuals are seated: evidenced by their feet resting on footstools (Mallon et al. 1934: 141-142, Pl. 66). The first individual, with his footstool higher than that of the second, is shod. The soles, the ankles and part of the leg are outlined by short oblique white lines (Figs. 16, 17). Albright (1957: 143-144) and Schick (1998: 37) suggest that this is indicative of embroidered shoes. The reading is difficult. Embroidery, highly developed sewing, is not an Egyptian nor a Levantine tradition (Kemp and Vogelsang-Eastwood 2000: 443) and embroidery to be an effective display medium must be visible, which is not on the soles. The earliest

Drabsch (2015: 105-106) also realigns and revises the Star fresco. Several disconnected motifs peripheral to the central motif are reassembled as three gowned and masked figures also in a processional mode. Zig-zag motifs appear on the gowns both vertically and horizontally. The initial and the subsequent reconstruction of the ceremony depicted in the Procession is structured, formal and hierarchical; inconsistent with shamanistic behavior, despite the presence of horned masks. The fresco called ‘The Notables’, to be discussed below, also features a ceremony with clear indications of social/ religious hierarchy. Drabsch (2015: 21) reconciles these two incongruous aspects of interaction with the supernatural – evolving or transitional shaman-priests with older elements of inspired shamanism co-existing with priestly learned practices. The footwear To recall, the dignitary laid out in The Cave of the Warrior, probably for decarnization prior to secondary burial, was not only clothed but also shod. Next to the deceased lay two cowhide sandals with signs of use wear, not a pair but worn as such,

Figure 15. Leather sandals, Cave of the Warrior, from Schick 2003, 16 courtesy of the Israel Antiquities Authority, photography by Olga Negnevitsky.

192

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period

Figure 16. Fresco, The Notables, Teleilat Ghassul from Mallon et al. 1934 PL. 66.

of fully clothed individuals, did not develop in a vacuum. It was but one of a cluster of innovations in settlement patterns, technologies, ideologies, enhanced investment in esthetics and concepts of self. It should be recalled that Ghassulian artisans were at the forefront of metallurgy with no known precursors or parallels (Sebbanne 2014: 119, 134), a proliferation of ceramic typology and intensification of painted and plastic decorations (Gilead 2007: 46) including the use of the potter’s wheel (Roux and Miroschedji 2009: 155), the production of monumental frescos in the buon fresco technique, with evidence of the developed concepts of landscape, ground line and bird’s eye perspective (Cameron 1981; Drabsch 2015: 50, 175, 179; Mallon et al. 1934), and a flowering of carving in bone, hippo and elephant ivory (Levy and Alon 1992: 65-71; Bar-Adon 1980: 16-23; Perrot 1959: 16). Bir es-Safadi, Abu Matar and Gilat, sites on the desert fringe, attest to a dramatic increase in the number of spindle whorls vis à vis other Southern Levantine sites of the period. Yet, despite extensive and intensive excavations of the sites, flaxseed was not found among the archaeo-botanics (Negbi 1955: 257-258; Zaitschek 1959: 48-52). To reiterate, flax is the only fibre attested in the Chalcolithic period, the previous PPNB and the subsequent Early Bronze Age. The water regime of the northern Negev is incompatible with flax cultivation and processing. Even in the Byzantine period with agriculture fueled by intensive water runoff management, flax was not one of the crops (Rubin 1996: 52-55). However, the Jordan valley with perennial water and marshy conditions, ideal for flax cultivation (Helback 1959: 106), attests to three Ghassulian sites, Teleilat Ghassul, Pella and Abu Hamid, from which flax seeds have been recovered, including two spinning bowls from the later (Breniquet 2008: 124, Fig. 29.3; Bourke1997: 410; Smith and Hanbury-Tension 1992: 24; Neef 1988: 29). Biblical sources (Joshua 2.6) and later Classical and Jewish sources also mention the Jordan and Beit Shean valleys as the primary loci of the region of flax cultivation and linen manufacture of the Southern Levant (Alon 1980: 168; Amar 1998: 114).

Figure 17. Fresco, The Notables, Teleilat Ghassul, detail of footwear, from Mallon et al. 1934 PL. 56.

unequivocal representation of shoes in the EgyptoLevantine cultural area appears more than 2000 years later in the tomb of Chnemhotep at Beni Hasan dated to 2000 B.C. (Newberry 1893: 69, PLS. XXVII, XXX) worn by the women of a clan of Semitic traders-cummetallurgists. Old World occurrences of footwear are recorded from the second half of the fourth millennium BCE. These include a single, moccasin type shoe from Armenia (Pinhasi et al. 2010) and from the same time bracket the more complex shoes of the Tyrolean mummy ‘Otzi’ from the Haslaboch Glacier, Austria-Italy (Hlavacek et al. 2004: 1). Sandals, elements of the royal wardrobe, are prominently displayed on the arm of an elite member of the entourage on the early 4th millennium BCE Namer Palette (Pritchard 1954: 296, 297). Discussion The Ghassulian textile industry, with evidence for linen textiles ranging from 2 cm to 7 m and representations 193

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead I suggest that the source of flax for the Beer Sheva valley sites was also from the same area. Spindle whorls are an element of the Judean desert cave sites repertoire (Bar-Adon 1980: 183-184; Ben-Yosef et al. 2017: Fig. 4). Thus, spinning was taking place. Flax was neither cultivated nor processed in the Judean desert because water and soil conditions are inappropriate. Therefore, flax was carried from sites where it was cultivated into the area as processed fibre. It is probable that the winter pasture of the Judean desert was exploited by members of settlements both to its north and south, with contemporary analogies (Eldar et al. 1992: 214; Levy 1992: 70) and that it was the locus for interaction and fibre transactions. Interaction may have also consisted of mating arrangements. Textile production is predominantly a female domain. The earliest graphic and textile sources attest to this (Barber 1991: 4). Raw materials and technology frequently travel in tandem. Furthermore, the presence of a spinning bowl at Bir es-Safadi, one of the four recovered from the Southern Levant, suggests north-south dispersion along the Jordan valley. It would be unlikely that the spinning bowl, that characterizes sophisticated manufacture of linen yarn, could develop in area where the raw material could not even be cultivated.

degree of formalization in transactions of economic nature, barter in processed flax to the fibre stage is very possible in the above-proposed modus. The adoption of drop spinning as the primary mode of yarn production was the motor that generated the textile industry. The yarn, well twisted during unimpeded free fall, is strong and uniform and capable of withstanding the unremittent tension imposed on the warp threads and constant friction during the changing of the sheds and compacting the cloth. Without drop spinning, the industry simply could not take off. Plied spindle yarn is manufactured approximately twice as quickly as thigh manufactured plied yarn with less energy investment (Tiedmann and Jakes 2006: 304). Rates of production of thigh manufacture plied yarn of inner bark tree bast from New Guinea and plant bast and inner bark tree bast from North America are comparable: 10-15 mph and 9 mph (MacKenzie 1991: 83; Tiedmann and Jakes 2006: 294, 301). Rates of yarn production in the drop spinning technique extrapolated from 10 minute session by 80 spinners from 20 Peruvian communities gives a rate of 62 mph for single yarn and 95 mph for 2 plied yarn (Bird 1968: 14). Yarn manufactured in the drop spinning technique, is more tightly twisted and uniform and therefore stronger. Thigh spun yarn can only be manufactured in a stationary mode, but drop spinning can be conducted in either a stationary or mobile mode. Increased speed in yarn production enabled the manufacture of clothing of woven cloth of full body covering dimensions as opposed to earlier minimal accessory type elements of attire, manufactured in labour intensive techniques. Examples of the latter include the headwear from Nahal Hemar and the girdle from Gilgal Hershman and Belfer-Cohen 2010: 187-188, Fig. 11.2; Schick 1988: 35-36, PL XVI).

The concept of the barter of raw materials, processed flax fibre, was not alien. It was simply grafted onto existing practices. There is ample evidence for the movement of heavy weight raw material and finished products through the system. Arsenic copper ore was imported from Anatolia or Caucasia and local copper ore from Wadi Feinan, Arava. In addition, archaeologically invisible but technologically crucial, hardwood charcoal was imported for copper smelting to the Beer Sheva valley interface sites (Perrot 1955: 7984 n. 15; Shugar 2000). Also evidenced is the movement of finely crafted basalt and phosphorite bowls, some c. one metre in diameter and elaborate censors from an unknown site in Transjordan to the northern Negev sites (Rosen 1990: 42). The latter are cultural markers for the domestication of the donkey. Worthy of mention are lightweight exotics from more distant sources, hippo ivory probably of Nilotic origin but possibly also from a residual population in local coastal waters (Horowitz and Tchernov 1990: 69), Nilotic and Red Sea shells, gold ingots (a single occurrence) of unknown provenance and without analogies (Gopher et al. 1990: 437), elephant ivory (Bar Adon 1980: PLS 16-21; Perrot 1955: 16), manufactured steatite beads (Bar-Yosef Mayer and Porat 2013: 345), and a single occurrence of lead (Langgut et al. 2016: Table 2, Fig. 6, 988). Also attested is a limited presence of seals and seal impressions indicative of ownership and socioeconomic transactions (Ben-Tor 1995: 361-375 with references therein; Mallon et al. 1934: 81; Tsori 1958: 47, PL. V.A). Given the above evidence, with an established practice of transactions in bulk raw material and a

Despite the increased speed in yarn production, there was a considerable disparity between yarn production and loom consumption. Spinning was a bottleneck throughout the ages until the very late Industrial Revolution (Patterson 1956: 161). Various scholars give figures for the ratio of spinners to weavers. According to Barber (1994: 87), it was 7:1; according to Patterson (1956: 161) with the presence of the treadle-operated spinning wheel it was 3-5: 1 while according to Seymour (2001: 332) it was 12:1. The Ghassulian textile industry is but the harbinger of what the future held for womankind. With the full-scale adoption of woven cloth as the primary mode of covering the human body, womankind became enslaved to the spindle. She could never produce enough. She was eternally shackled to the spindle. Ethnographic observations without number, literary sources, and graphics attest to woman ‘at their eternal task’ (Catallus died 54 BCE) (Levy and Gilead 2013: 134-135 with references therein). Spinning consumed every free minute. 194

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period The magnitude of the textile industry can be determined by the quantity, the quality, and the attitude towards the artefacts. The large number of textiles throughout the seasonally occupied Judean desert cave sites associated with domestic repertoires suggests widespread usage (Aharoni 1961: 14-15; Avigad 1962: PLS. 16-17; Bar-Adon 1980: 57-58, 135, 191). Perhaps even more telling is the use of a textile as a rag stuffed into a copper mace head recovered from Nahal Mishmar (Bar-Adon 1980: 119). The latter is a marker that textiles were sufficiently common to be treated in an off-hand manner.

do not allow the warp threads to be woven until the final centimeter. The tension is too great. Moreover, the first few centimetres of the cloth cannot be woven. The vertical distance between the two warp sets is too great. Among social groups of limited means, more frugal behavior is observed. The terminal warp threads are sewn around the outermost weft threads in overcast stitch to prevent the textile from unravelling. Subsequently the residual centimetres of warp ends are cut off and re-spun. The unspoken statement inherent in fringes was subsequently formalized in ritual and transactions in southwest Asia, particularly Mesopotamia (Munn-Rankin 1956: 92; Milgrom 1983: 61).

The earliest graphic evidence for woven gowns from the Teleilat Ghassul Processional Frieze fresco dates to the 3rd quarter of the 5th millennium BCE (Gilead 2003: 223). The gowns are narrow and without fringes. The major textiles from the Cave of the Warrior date to the 1st quarter of the 4th millennium BCE and the textiles of an earlier burial to the mid-5th millennium (Jull et al. 1998:111). The shroud is a marker for a developed industry. Novice weavers weave narrow textiles, and the products are characterized by weaving mistakes, loss of width, uneven weft rows and weaving wedges to adjust the discrepancies. None of these problems is present in the shroud. The shroud measures 7 x 2 m and speaks of skill, accumulated experience, and considerable expertise in handling ca. 3,500 warp threads. The manufacture of this textile of unsurpassed dimensions was obviously preceded by many of smaller sizes.

It is hypothesized that the male laid out in the Cave of the Warrior was a pastoral nomad dwelling in the uplands to the west or east of the Jordan valley with close connections with the settled communities of his area (Ashkenazi and Goren 2010). This conclusion is based on the osteological report, microscopic investigation of material adhering to the sandals, and a re-evaluation of the grave goods. Ceramic vessels were not found among the grave goods. However, we find there a wooden bowl fashioned from the burl of a Tabor oak, use of cowhide for the sandals and repair of the basket. In addition, we find an olive wood bow and reed arrows, which reflect an outdated technology in the Ghassulian. All of these finds were made from raw materials, which were alien to the Jordan valley but which were components of the upland Mediterranean type environments on either side of the valley (Ashkenazi 2008: 94).

It is difficult to suggest that this is the result of a long tradition of weaving. After all, direct and indirect evidence is lacking. However, the ability to manufacture such a large textile without loss of width and without weaving mistakes implies skills accumulated over at least several generations. The closest analogy is the documented trajectory of the Navajo of the Navajo in late historical times, 1680-1770 CE. During this period of less than a century the Navajo transformed themselves from hunter-gathers wearing skins to accomplished weavers and suppliers of woolen blankets to other Indian entities and also to the Spanish. The skill was not an indigenous development. It was acquired from Pueblo Indians concealed amongst them to escape reprisals by the Spanish (Kent 1961: 5-6). Mastery of skills would appear to be one of aptitude and motivation as opposed to long drawn out incremental acquisition.

The walking stick, buried with the individual would appear to negate this reconstruction. The individual had broken his left leg a short time before his death, and he limped. Heavy use wear on the left sandal confirms the osteological report (Schick 1998: 37). The stick is covered, on one face, with equidistant furrows 10-12 to the centimetre (Schick 1998: 28, Fig. 6.2), apparently a recycled heddle rod. Ashkenazi (2008: 100) suggests that a member of an agricultural family, that also practiced weaving, gave him the stick as a gift. The stick of simple non-worked willow was for utilitarian needs and not of a quality appropriate for a gift. Such recycling would be more consistent with an act from within a household than a gift from one of the proposed farming communities. A more plausible reconstruction is that the individual came from a weaving household, probably the source of the textiles rather than a nebulous upland encampment. Furthermore, I propose that the interred had been a resident of Teleilat Ghassul, only 12 kms distant. Teleilat Ghassul features storage bins with flaxseeds (Bourke 1997; 406, 410, 413) and evidence of a former, marshy environment suitable for their cultivation (Hennessy 1969: 21), a considerable spindle whorl assemblage (Mallon et al. 1934: 72), a fragment of a linen textile (Lee 1973: 304 citing R,

Clothing of woven cloth lends itself to manipulation. The shroud was not only a funerary garment but also a vehicle for enhancing the prestige of the individual and his extended family or clan. The excessive amount of cloth, far beyond pragmatic needs (Schick 1998: 22), is a social statement. The shroud features elaborate fringes in both the warp and weft. These are highly visible markers of conspicuous consumption (Veblen 1967: 157). The technological limitations of the loom 195

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Pfister 6.1.39) and a fresco depicting cultic dignitaries clad in woven gowns (Cameron 1981; Drabsch 2015).

The earliest clothing representations from Upper Paleolithic Eurasia do not depict utilitarian clothing, rather nude females wearing scanty items of attire, in which headwear predominates. Also observed are fringed girdles, belts and straps (Soffer et al. 2000: 514-522, Figs. 2-10), which are ritual wear or markers for a given state, condition or potential.12 Epi-palaeolithic and Neolithic representations and findings of southwest Asia evince a similar trajectory with a focus on headwear, whose high visibility lends itself to symbolic manipulation (Levy 2018). Headwear was recovered from Neve David (Kaufman and Ronen 1987: 337), from the Carmel (Garrod and Bate 1937: PL. VII.2), Nahal Hemar (Schick 1988: Figs. 2, 6), and their representations from Sha’ar Hagolan (Garfinkel 1999: 55). Also recovered was a schematic representation in the round of a female with a fringed girdle from Gilgal (Hershman and Belfer-Cohen 2010: 187-188, Fig. 11.2) and detailed, female clay figurines from Sha’ar Hagolan clad in waistcoats and scarfs (Garfinkel 1999: 47, 50). Male representations wearing loincloths are attested from Göbekli Tepe, Anatolia (Schmidt 2010: 245, Fig. 2); Ҁatal Hüyük (Mellaart 1966: PL. LIV), and also on Halafian ceramics (Garfinkel 2003: 37).

The shroud from the cave of the warrior lacks evidence of use-wear (Schick 1998: 21). Shamir (2014: 150) proposes that it was specifically manufactured as an item of funerary wear. The time required to manufacture this textile was calculated in person hours (Levy and Gilead 2013: 148). When splicing, spinning, plying and preparation of leash yarn, year round activities conducted in tandem with domestic activities and herding, are removed from the calculation, setting up the loom, weaving the body of the textile and elaboration of the ends and the weft and warp fringes the manufacture of the textile would have taken c. 50 six hour working days of two skilled adult females and c. 30 days labour of three, small, not economically productive children. Therefore, if the shroud represents funerary wear, manufactured specifically for the event, one must assume that certain households stockpiled mortuary textiles. Secondary burial, the predominant mortuary mode of the Ghassulian, is planned reburial. Possibly death was also planned, with the appropriate equipment on hand. It is not without parallels. In the dynastic periods of Egypt, the elite and specifically the pharaohs constructed mortuary structures and accumulated burial goods from the time of their ascension to power (David 1999: 119, 147).

It would appear that The Processional Frieze with full body covering and their correlates from the Cave of the Warrior constitute a watershed in patterns of human behavior. In these scenes clothed bodies are in, adorned nudity is out. From the 4th millennium BCE onward, and the advent of urban hierarchical societies, the fully clad body, clothed in woven cloth, becoming fuller overtime is de rigueur. In fact, in both the Mesopotamian and Levantine spheres this becomes the cultural norm. Nakedness, in both Egyptian and Mesopotamian representations, is reserved for the dehumanized and mutilated enemy. Woven cloth and full body covering was a point of no return. It was subsequently considered a mark of civilization, reflected in the myths of many cultures that place spinning and weaving at the beginning of their own history. They portray these fibre techniques as either inventions or gifts of the gods (Broudy1979: 9). As a counterpoise, and explicit in the Epic of Gilgamesh, clothing is a mark of humanity. To be without woven clothing is beyond the pale of civilization and an attribute of the animal world as personified by Enkidu before being brought into the realms of civilization (Foster 2001: 13-14,171-173). Already by the fourth millennium BCE, in the Mesopotamian cultural sphere, woven cloth garments are associated with power and status. Those in a position of authority are clothed in woven cloth while those bearing tribute or offerings are naked (Breniquet 2013: 12).

The fresco called ‘the Procession’ is a marker that heralds a new era with a new concept of worthy self, the fully covered human body, clothed in woven cloth. From this time forth men of standing, and subsequently women also, will be depicted fully clothed and frequently hatted and shod too.11 The rather narrow robes of Teleilat Ghassul are replaced by fuller models with elaborated hems and fringes. Total body covering was certainly worn prior to this, but it does not appear in representations. The presence of humans in the centre of continents and middle latitudes, known paleo-climates, the known physiological limits of humans’ exposure to sustained cold and tools associated with hide workings with ethnographic analogies attest to the use of even fitted and multi-layered clothing (Gilligan 2010: 20-21, 33-34). Consistent, not intermittent, use of clothing is dated by genetic studies to c. 100,000 years ago. Body lice are nourished on blood, but only breed in clothing. When they are separated from their hosts, they live only a few hours. Thus, their emergence as a sub-species with a distinct ecological niche. Their continued survival is entirely dependent on the daily use of clothing by some people (Bell 2015: 207; Shao et al. 2012). There are obviously exceptions as in the Hellenistic period with the glorification of the unadorned human form. Nevertheless, glorified nudity was appropriate for the gymnasium or sporting events and not for symposiums or the agora. 11 

An Upper Palaeolithic ivory figurine from Buret, Russia depicts an individual wearing a fitted pants suit with a hood with surface markings possibly indicating fur (Gilligan 2010: 57, fig. 15). 12 

196

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period Conclusions

Only with widespread access to insulating wool for thermal protection was the transition to a complete textile economy accomplished. When that happened, full body covering with woven cloth became the cultural norm in Mesopotamia and the Southern Levant.

It is proposed that flax processed to the fibre stage in the Jordan Valley was obtained through barter with seasonal herding members of the same communities by inhabitants of the Beer Sheva Valley sites. It was processed in households, spliced with saliva and pressure, and spun and plied in the drop spinning technique on high whorled spindles with discoid whorls. Spinning was carried out by the young girls of the household while herding the family flocks. Spinning bowls played a minimal role in yarn production. The 2ply (s2S) linen yarn was woven on horizontal ground looms with heddle technology into plain weave textiles with simple selvedges from continuous weft threads. Textiles were transformed into items of wear by draping, belting or pinning, not by tailoring.

The Catholic marriage of comparatively rapidly spun strong yarn and the horizontal ground loom engendered far-reaching changes in cultural behavior. The skilled manipulation of a stick and a stone with a hole in the middle and another five sticks changed the way we interact one with the other, our display mechanisms, our rapid person evaluations and concepts of appropriate behavior. Ninety nine point nine percent of the people reading this article are wearing full body clothing from spun yarn woven in plain weave, undoubtedly spun and woven with greater rapidity. However, the basic technology remains unchanged. Likewise, the basic concept of appropriate social behavior, full body covering with woven cloth, is still current.

The overwhelming majority of spindle whorls were recovered from the Beer Sheva Valley sites, an area with a water regime inconsistent with cultivation and processing of flax. However, it would seem unlikely that this was the centre of the textile industry. It is far more probable that the upper Jordan Valley was the centre of gravity because of its appropriate water regime, archaeo-botanical evidence, and a later reputation. Indeed, in the Classical period, the Jordan Valley was known throughout the Mediterranean world, not only as the centre of the local linen industry but also as the source of the finest linen of the period (Graser 1959: 386-390). Direct evidence for the Ghassulian period is currently lacking. The Ghassulian sites are undoubtedly hidden under the alluvium. Like so many other fortuitous discoveries, they will probably be exposed during construction of lines of infrastructure.

Acknowledgements I would like to acknowledge the astute insights of my advisor, Professor Isaac Gilead, who has guided me through this research. I also wish to take this opportunity to extend my gratitude for the help and encouragement of my colleagues Yehuda Mansell and Dr. Karni Golan, for the photography by Alter Fogel and Evgeny Ostrovsky and the graphic wizardry by Patrice Kaminsky. Special thanks are extended to graphics artist Yuval Shach and to my son, Natanel Levy. I also thank the Israel Antiquities Authority for granting me permission to use images from one of their publications and to the deceased Professor Jean Perrot, may his name be an archaeological blessing, who granted me permission to publish his material.

Textiles from those of ribbon like dimensions to those measuring 7 x 2 m were recovered exclusively from the Judean desert and the Dead Sea basin. They are obviously only an infinitesimal percentage of what was produced. Considerable time elapsed between the first faltering steps of the industry and the production of the 7 x 2 m shroud. Spinning and weaving is characterized by a marked disparity between yarn production and loom consumption. The theme of incessant demand and the unbridgeable gap was elaborated through the ages in graphics, folklore and texts. The textile industry was a female domain. The spread of the technology on a north to south axis was dispersed via a movement of females in formalized mating arrangements. From the mid-5th millennium BCE onward, men of worth are depicted in full body, woven robes and footwear. Physical correlates from several hundred years later, in the Cave of the Warrior expose a veritable dignitary’s wardrobe, with elaboration of fringes and hems consistent with conspicuous consumption. Despite the manufacture of linen clothing, skins and furs continued to be worn for a considerable time. Linen is not warm.

References Abbo, S. Zezak, I. Lev-Yadun, S. Shamir, O. Friedman, T. and Gopher, A. 2014. Harvesting Wild Flax in the Galilee, Israel and Extracting plant Fibres – Bearing on Near Eastern Plant Domestication. Israel Journal of Plant Sciences: 1-13. Aharoni, Y. 1961. Expedition B. Israel Exploration Journal 11: 11-24. Albright, W. 1957. From the Stone Age to Christianity. Garden City, New York: Doubleday. Alfaro-Ginar, C. 1984. Tejidos y Cestería en la Peninsula Iberica. Madrid, Institut Español de Prehistoria del Consejo Superior de Investigaciones Cientificas. Alfaro-Ginar, C. 2012. Textiles from the Pre-Pottery Neolithic Site of Tell Hallula (Euphrates Valley, Syria). Paléorient 38.1-2: 41-54. Allen, S. 1997. Spinning Bowls: Representations and Reality. In J. Phillips, L. Bell, B. Williams, J. Hoch and R. Leprohon (eds), Ancient Egypt, the Aegean and Near 197

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead East: Studies in Honour of Martha Rhodes Bell, Vol I: 1738. San Antonio, Van Siclen. Alon, G. 1970. The Jews in their Land in the Talmudic Age. Jerusalem, Hebrew University. Amar, Z. 1998. Written Sources Regarding the Jazirat Far’un (Coral Island) Textiles. Atiqot XXXVI: 114-119. Anderson, E. 1987. The Spinner’s Encyclopedia. Newton Abbot, U.K., David and Charles. Avigad, N. 1962. Expedition A. Israel Exploration Journal 12: 169-183. Ashkenazi, H. 2008. The Archaeology of the Individual: Reconstructing the Life of the Deceased from the Cave of the Warrior. Unpublished MA thesis, University of Tel Aviv (Hebrew). Ashkenazi, H. and Goren, Y. 2010. The Archaeology of the Individual: Reconstructing the Life of the Deceased from the Cave of the Warrior. In P. Matthiae, F. Pinnock, L. Nigro, and N. Marchetti (eds), Proceedings of the 6th International Congress of Archaeology of the Ancient Near East Vol. I: 177187. Weisbaden, Harrassowitz Verlag. Avigad, N. 1962. Expedition A – Nahal David. Israel Exploration Journal 12: 169-183. Baines, P. 1989. Linen: Handspinning and Weaving. London, Batsford. Bar-Adon, P. 1980. The Cave of the Treasure. Jerusalem, Israel Exploration Society. Barber. E. 1991. Prehistoric Textiles. Princeton, New Jersey, University Press. Barber, E. 1994. Women’s Work: The First 20,000 Years. New York, Norton. Bar-Yosef Mayer, D. and Porat, N. 2013. Beads. In D. Shalem, Z. Gal and H. Smithline (eds), Peqi’in: A Late Chalcolithic Burial Site Upper Galilee, Israel: 337-369. Kinneret College, Ostracon. Bedigan, D. 1985. Is š-gis-ī Sesame or Flax. Bulletin on Sumerian Agriculture: 159-178. Bell, G. 2015. The Evolution of Life. Oxford, Oxford University Press. Ben-Tor, A. A Stamp Seal and a Seal Impression of the Chalcolithic Period from Grar. In I. Gilead (ed.), Grar: A Chalcolithic Site in the Northern Negev: 361-375. BeerSheva VII. Beer Sheva, Ben-Gurion University of the Negev Press. Ben-Yosef, E., Shamir, O. and Levy, J. in press. The Earliest Near Eastern Wooden Spinning Implements? A Rejoinder to Langgut et al. 2016. Antiquity. Bergfjiord, C., Karg. S., Rast-Eicher, A., Nosch, M-L., Mannering, U., Allaby, R., Murphy, B., and Holst, B. 2010. Comment on 30,000 Wild Flax Fibers. Science 328: 1634. Bird, J. 1968. Handspun Yarn production Rates in the Cuzco region of Peru. Textile Museum Journal 2/3: 9-16. Bonani, G. 1995. AMS Analysis Procedures. Atiqot XXVII: 205-206. Bourke, S. 1997. The ‘Pre Ghassulian’ Sequence at Teleilat Ghassul: Sydney University Excavations 1975-1995.

In H. Gebel and Z. Kafafi (eds), The Prehistory of Jordan II. Perspectives from 1997: 395-416. Berlin, ex oriente. Breniquet, C. 2006. ‘Ce Lin, Qui Me Le Peignera?’ Enquête sur la Fonction des Peignes en Os du Néolithique Précéramique Levantin. Syria 83: 167-176. Breniquet, C. 2008. Essai sur le Tissage en Mesopotamie. Paris, de Boccard. Breniquet, C. 2013. Function and Use of Textiles in the Ancient Near East. In M.-L. Nosch, H. Koefoed and E. Andersson Strand (eds), Textile Production and Consumption in the Ancient Near East: 1-25. Ancient Textile Series 13. Oxford: Oxbow. Breniquet, C. 2014. The Archaeology of Wool in Early Mesopotamia. In C. Breniquet and C. Michel (eds) Wool Economy in the Ancient Near East and Aegean: 5278. Ancient Textile Series 17. Oxford, Oxbow. Broudy, E. 1979. The Book of Looms. New York, Brown University Press. Brunton, G. and Caton-Thompson, G. 1928. The Badarian Civilization. London, Bernard Quaritch. Burnham, D. 1973. Cut my Cote. Toronto, Royal Ontario Museum. Cameron, D. 1981. The Ghassulian Wall Paintings. London, Kenyon and Deane. Carrington-Smith, J. 1975. Spinning Weaving and Textile Manufacture in Prehistoric Greece. Unpublished PhD dissertation, University of Hobart. Cassuto, D. 2008. Bringing the Artifacts Home: a Social Interpretation of Loom Weights in Context. In B. Nakhai (ed.), The World of Women in the Classical and Ancient Near East: 63-77. Newcastle upon Tyne, Cambridge Scholars Publishing. Chang, E. 1995. A History of Lace. In J. Turner and P. van de Griend (eds), A History and Science of Knots: 347377. Singapore, Scientific World. Cindorf, E., Horowitz, S., and Blum, R. 1980. Textile Remains from Nahal Mishmar. In P. Bar-Adon (ed.), The Cave of the Treasure: 229-234. Jerusalem, Israel Exploration Society. Commenge-Pellerin, C. 1990. La Potterie de Safadi (Beersheva) au IVe millenaire avant l’ère Chretiénne. Paris, Association Paléorient. Crewe, L. 1998. Spindle Whorls. Jonsered, Sweden, Paul Åstöms Förlag. Crocker-Jones, G. 1990. Traditional Spinning and Weaving in the Sultanate of Oman. Muscat, Historical Association of Oman. Crowfoot, E. 1982. Appendix B. Textiles, Matting and Basketry. In K. Kenyon and T. Holland (eds), Excavations at Jericho Vol. IV: 546-550. London, British School of Archaeology at Jerusalem. Crowfoot, G. 1931. Methods of Handspinning in Egypt and the Sudan. Halifax: Bankfield Museum Notes. Crowfoot, G. 1933. The Mat Weaver from the Tomb of Khety. In F. Petrie, M. Murray and D. Mackay (eds), Ancient Egypt and the East: 93-99. London, Macmillan. Crowfoot, G. 1945. The Tent Beautiful. Palestine Exploration Quarterly 77: 34-46. 198

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period Crowfoot, G. 1954. Textiles, Basketry and Mats. In C. Singer, E. Holmyard and A. Hall (eds), A History of Technology: 411-447. Oxford, Clarendon Press. David, R. 1999. Handbook to Life in Ancient Egypt. Oxford, Oxford University. Davidovich, U. 2008. The Chalcolithic Period in the Judean Desert: Identification, Settlement Pattern and Material Culture as a Basis for Environmental Reconstruction. Unpublished MA thesis, Hebrew University at Jerusalem (Hebrew). Dempsey, J. 1975. Fiber Crops. Gainesville, Fla, University of Florida. Dothan, T. 1963. Spinning Bowls. Israel Exploration Journal 13/2: 97-112. Eldar, I. Nir, Y. and Nahlieli, D. 1992. The Bedouin and their Campsites in the Dimona Region of the Negev: Comparative Model for the Study of Ancient Settlements. In O. Bar-Yosef and A. Khazanov (eds), Pastoralism in the Levant: 205-217. Madison, Wisconsin, Prehistory Press. Ejstrud, B., Andresen, S., Appel, A., Gjerlevsen. S., and Thomsen, B. 2011. From Flax to Linen. Esbjerg: University of Southern Denmark. Emery, I. 1966. The Primary Structure of Fabrics. Washington, D.C: The Textile Museum. Fischer, P. 2008. Tell Abu Al-Kharaz in the Jordan Valley. Vienna, Osterreichischen Akademie der Wissenschaften. Foster, B. 2001. The Epic of Gilgamesh. New York, W.W. Norton and Co. Franquemont, E. 1986. Cloth Production in Chinchero Peru. In A. Rowe (ed.), The Junius B. Bird Conference on Andean Textiles: 309-329. Washington: Textile Museum. Frödin, O. and Nordenskiöld, E. 1918. Über Zwiren und Spinnen bei den Indianern Südamericas. Göteborg, Wald, Zachrissons Boktryckrei. Galili, E. and Sharvit, J. 1994-95. Evidence of Olive Oil Production from the Submerged Site at Kfar Samir, Israel. Journal of the Israel Prehistoric Society 26: 122-133. Garfinkel, Y. 1999b. The Yarmukians. Jerusalem, Bible Lands Museum. Garfinkel, Y. 2003. Dancing at the Dawn of Agriculture. Austin, Texas, University of Texas Press. Garrod, D. and Bate, D. 1937. The Stone Age of Mount Carmel Vol. I. Oxford, Clarendon Press. Gilead, I. 1988. The Chalcolithic Period in the Levant. Journal of World Prehistory 2: 397-439. Gilead, I. 2002. Religio-Magic Behaviour in the Chalcolithic Period of Palestine. In S. Ahituv (ed.), Beer-Sheva XV: Aharon Kempinski Memorial Volume: 103-128. Beer Sheva, Ben-Gurion University of the Negev Press. Gilead, I. 2003. Book Review of Lovell, J. The Late Neolithic and Chalcolithic Periods in the Southern Levant: New Data from the Site of Teleilat Ghassul, Jordan. British Archaeological Reports International Series 974. Mitekufat Haeven 30: 218-224.

Gilead, I. 2007. The Besorian: a Pre-Ghassulian Cultural Entity. Paléorient 33/1: 33-49. Gilead, I and Gošić, M. 2014. Fifty Years Later: a Critical Review of Stratigraphy, Chronology, and Context of the Nahal Mishmar Hoard. Journal of the Israel Prehistoric Society 44: 226-239. Gilligan, I. 2010. The Prehistoric Development of Clothing: Archaeological Implications of a Thermal Model. Journal of Archaeological Theory 17: 1580. Gopher, A., Tsuk, T., Shalev, S., and Gophna, R. 1990. Earliest Gold Artifacts in the Levant. Current Anthropology 11/4: 436-443. Graser, E. 1959. Appendix: The Edicts of Diocletian on Maximum Prices. In T. Frank (ed.), An Economic Survey of Ancient Rome Vol. I: 305-421. Paterson, New Jersey, Pageant Books. Grigson, C. 2006. The Worked Bone from the Chalcolithic Site of Gilat. In T. Levy (ed.), Archaeology, Anthropology and Cult: 685-702. London, Equinox. Hardy, K. 2008. Prehistoric String Theory. How Twisted Fibres Helped to Shape the World. Antiquity 82: 271280. Harvey, J. 1996. The Textiles of Central Asia. London, Thames and Hudson. Helbaek, H. 1959. Notes on the Evolution and History of Linnum. Kuml Aarhus: 103-129. Hennessy, J. 1969. Preliminary Report on a First Season of Excavations at Ghassul. Levant 1: 1-24. Hershmann, D. and Belfer-Cohen, A. 2010. ‘It’s Magic!’: Artistic and Symbolic Material Manifestations from Gilgal Sites. In O. Bar-Yosef, N. Goring-Morris and A. Gopher (eds), Gilgal. Early Neolithic Occupations in the Lower Jordan Valley: The Excavations of Tamar Noy: 185216. Oxford, Oxbow. Hess, K. 1954. (5th ed.). Textile Fibers and their Uses. Chicago, J. B. Lippincott. Hlavacek, P., Ostravska, L., Vaclav, G., Blaha, A. and Vaculik, J. 2004. Testing the Characteristics of Replicas of Stone Age Footwear Recovered from the Oetzi Italian Alps. Paper presents at Emed Award Conference, Leeds, UK. Hochberg, B. 1979. Spin, Span, Spun. Santa Cruz, Cal, SelfPublished. Hochberg, B. 1980. Handspindles. Santa Cruz, Cal, SelfPublished. Horowitz, L. and Tchernov, E. 1990. Cultural and Environmental Implications of Hippopotamus Bone Remains in Archaeological Contexts in the Levant. Bulletin of the American Schools of Oriental Research 280: 67-76. Hurcombe, L. 2014. Perishable Material Culture in Prehistory. London: Routledge. Joffee, A., Dessel, J. and Hallote, R. 2001. The Gilat Woman. Near Eastern Archaeology 64/1: 8-23. Jull, A., Donahue, D., Carmi, I. and Segal, I. 1998. Radiocarbon Datings of the Finds. In T. Schick (ed.), 199

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead The Cave of the Warrior: 110-111. Israel Antiquities Authority Reports 5. Jerusalem, Israel Antiquities Authority. Kania, K. 2013. The Spinning Experiment: Influences on Yarn in spinning with a Hand-Spindle. In H. Hopkins (ed.), Ancient Textiles Modern Science: 11-29. Oxford: Oxbow. Kaufman, D. and Ronen, A. 1987. Le Sépultre Kébarienne de Névé David Haïfa, Israël. Anthropologie 91/1: 335342. Kemp, B. and Vogelsang-Eastwood, G. 2000. The Ancient Textile Industry at Amarna. London, Egyptian Exploration Society. Kent, K. 1961. The Story of Navajo Weaving. Phoenix, Arizona, Heard Museum of Anthropology and Primitive Art. Kislev, M., Simchoni, O., Melamed, Y. and Maroz, L. 2011. Flax Seed Production: Evidence from the Early Iron Age Site of Beth-Shean, Israel and from Written Sources. Vegetation History and Archaeobotany 20/6: 579-584. Kissell, M. 1918. Yarn and Cloth Making. New York, Macmillan. Koren, Z. 1998. Color Analysis of the Textiles. In T. Schick (ed.), The Cave of the Warrior: 100-109. Israel Antiquities Authority Reports 5. Jerusalem, Israel Antiquities Authority. Kramer, C. 1982. Village Ethnoarchaeology. New York, Academic Press. Kvavadze, E., Bar-Yosef, O., Belfer-Cohen, A., Boaretto, E., Jakeli, N., Matskevitch, Z. and Meshvelani, T. 2009. 30,000- year-old Wild Flax Fibers. Science 328: 1359. Kvavadze, E., Bar-Yosef, O., Belfer-Cohen, A., Boaretto, E., Jakeli, N., Matskevitch, Z. and Meshvelani, T. 2010. Response to ‘Comment on 30,000-year-old Wild Flax Fibers’ Science 328: 1634. Landi, S. and Hall, R. 1979. The Discovery and Conservation of an Ancient Egyptian Tunic. Studies in Conservation 24: 141-152. Langgut, D., Yahalom-Mack, N., Lev-Yadun, S., Kremer, E., Ullman, M. and Davidovich, U. 2016. The Earliest Near Eastern Wooden Spinning Implements. Antiquity 90: 973-990. LeBreton, L. 1957. The Early Period at Susa. Iraq 19: 79124. Lee, R. 1973. Teleilat Ghassul: New Aspects and Master Typology. Unpublished PhD dissertation. Hebrew University at Jerusalem. Levy, J. 2006. The Chalcolithic Textile Industry in the Southern Levant: Tools, Technology and Products. Unpublished MA thesis, Ben-Gurion University of the Negev. Levy, J. 2018. From Adorned Nudity to a Dignitary’s Wardrobe: Symbolic Raiment in the Southern Levant 13,500 -3,900 BCE. In M. Siennicka, I. Rahmstorf, and A. Ulanawaska (eds), First Textiles, The Beginnings of Textile Manufacture: 39-50. Oxford, Oxbow.

Levy, J. and Gilead, I. 2012. Spinning in the 5th Millennium: Aspects of the Textile Economy. Paléorient 38/1-2: 127-139. Levy, J. and Gilead, I. 2013. The Emergence of the Ghassulian Textile Industry in the Southern Levant Chalcolithic Period (c. 4500-3900 BC). In M.-L. Nosch, H. Koefoed and E. Andersson Strand (eds), Textile Production and Consumption in the Ancient Near East: Archaeology, Epigraphy, Iconography: 26-44. Oxford, Oxbow. Levy, T. 1992. Transhumance, Subsistence and Social Evolution in the Northern Negev Desert. In O. BarYosef and A. Khazanov (eds), Pastoralism in the Levant: 65-82. Madison, Wisconsin, Prehistory Press. Levy, T. and Alon, D. 1992. A Corpus of Ivory Artifacts from Shiqmim. Eretz Israel 33: 65-71 (Hebrew). Levy, T., Connor, W., Rowan, Y. and Alon, D. 2006. The Intensification of Production at Gilat: Textile Production. In T. Levy (ed.), Archaeology, Anthropology and Cult: 705-738. London, Equinox. Leuzinger, U. and Rast-Eicher, A. 2011. Flax Processing in the Neolithic and Bronze Age Pile Dwelling Settlements of Eastern Switzerland. Vegetation History and Archaeobotany 20: 535-542. Lewis-Williams, J. and Dowson, T. 1988. The Signs of all Times: Entoptic Phenomena in Upper Paleolithic Art. Current Anthropology 29: 201-245. Liu, R. 1978. Spindle Whorls Part 1. Some Comments and Speculations. Bead Journal 3: 87-103. Macdonald, E. 1932. Prehistoric Fara. In Beth Peleth II. London, British School of Archaeology in Egypt and Bernard Quaritch. MacKenzie, M. 1991. Androgynous Objects: String Bags and Gender in Central New Guinea. Chur, Switzerland, Harwood Academic. Mallon, A., Koeppel, R. and Neuville, R. 1934. Teleilat Ghassul I. Rome, Institut Biblique Pontifical. McCorriston, J. 1997. The Fiber Revolution. Current Anthropology 38/4: 517-549. Médard, F. 2012. Switzerland Neolithic Period. In M. Gleba and U. Mannering (eds), Textiles and Textile Production in Europe: From Prehistory to 400 AD: 367377. Ancient Textiles Series 11. Oxford, Oxbow. Milgrom, J. 1983. Of Hems and Tassels. Biblical Archaeology Review 9/3: 61-65. Mellaart, J. 1966. Excavations at Ҁatal Hüyük, 1965. Fourth Preliminary Report. Anatolian Studies XVI: 165- 192. Minar, J. 2002. Spinning and Plying. In P. Drooker and L. Webster (eds), Beyond Cloth and Cordage: 85-99. Salt Lake City, University of Utah. Munn-Rankin, J. 1956. Diplomacy in Western Asia in the Early Second Millennium B.C. Iraq XVIII: 68-110. Nagel, G. 1938. La Ceramique du Nouvel Empire a Deir el Medineh l. Cairo, Imprimerie de l’institut francais d’archeologie orientale. Neef, R. 1988. Les Activités Agricoles et Horticoles. In G. Dollfus and Z. Kafafi (eds), Abu Hamid Village du 200

J. Levy: Clothes Maketh (Hu)Man: Textile Production in the Southern Levant in the Chalcolithic Period 4e Millénaire de la Vallée du Jourdaine: 29-30. Amman, Economic Press. Negbi, M. 1955. The Botanical Finds at Tell Abu Matar, near Beersheba* Israel Exploration Journal 5: 257258. Newberry, P. 1893. Beni Hasan I. London, Egyptian Exploration Fund. Obermaier, H. 1928. Fossil Man in Spain. New Haven, Conn, Yale University Press. Patterson, R. 1956. Spinning and Weaving. In C. Singer, E. Holmyard and E. Hall (eds), A History of Technology, Vol. III: 151-186. Oxford, Clarendon Press. Perrot, J. 1955. Tell Abu Matar Near Beersheba. Israel Exploration Society 5: 17-40, 73-84, 167-189. Perrot, J. 1959. Statuettes en Ivoire et Autres Objects en Ivoire et Os Provenant des Gisements Prehistoriques de la Region Beersheba. Syria XXXVI: 8-19 PL. II, III. Perrot, J., Zori, N. and Reich, Y. 1967. Neve Ur, un Nouvel Aspect du Ghassoulien. Israel Exploration Journal 17/4: 201-23. Picton, J. and Mack, J. 1989. African Textiles. New York, Harper and Row. Pinhasi, R., Gasparian, B., Areshian, G., Zandaryan, D. and Smith, A. 2010. First Direct Evidence of Chalcolithic Footwear from the Near Eastern Highlands. PLoS ONE 5(6): e10984.doi: 10.1371/journal.pone.0010984 Pritchard, J. 1954. The Ancient Near East in Pictures. Princeton, Princeton University Press. Rast-Eicher, A. 2012. Switzerland Bronze and Iron Ages. In M. Gleba and U. Mannering (eds), Textiles and Textile Production in Europe: 378-396. Oxford, Oxbow. Raven, L. 2003. Spin It. Loveland, Colorado, Interweave. Read, H. 1964. The Art of Ancient Egypt. In Read, H. (ed.), Discovering Art, vol. 1. 129-192. London, Purnell. Renfrew, J. 1973. Paleoethnobotany. New York, Columbia University. Rosen, S. 1990. Metals, Rocks, Specialization and the Beginning of Urbanization in the Northern Negev. Supplement of Biblical Archaeology Today: 41-56. Roth, L. 1913. Ancient Egyptian and Greek Looms. Halifax: Bankfield Museum Notes. Roth, L. 1918. Studies in Primitive Looms. Halifax, King and Sons. Roux, V. and Miroschedji, P. 2009. Revisiting the History of the Potter’s Wheel in the Southern Levant. Levant 41/2: 155-173. Rowan, Y. and Golden, J. 2009. The Chalcolithic Period of the Southern Levant. Journal of World Prehistory 22: 1-92. Roy, J. 2011. The Politics of Trade: Egypt and Lower Nubia in the 4th millennium BC. Leiden: Brill. Rubin, R. 1996. Urbanization, Settlement and Agriculture in the Negev. The Impact of the RomanByzantine Empire on the Frontier. Zeitschrift des Deutschen Palästina Vereins 112:49-64. Ryder, M. 1965. Report of Textiles from Ҁatal Hüyük. Anatolian Studies XV: 175-176.

Ryder, M. 1969. The Origins of Spinning. Textile History 1: 73-82. Schick, T. 1988. Nahal Hemar: Cordage, Basketry and Fabrics. Atiqot 18: 31-43. Schick, T. 1994. Textile Remains. Atiqot XXXI: 97-98. Schick, T. 1995. A 10,000 Year Old Comb from Wadi Murabba’at in the Judean Desert. Atiqot XXVII: 199204. Schick, T. 1998. The Textiles. The Cave of the Warrior. Israel Antiquities Authority Report 5. Jerusalem, Israel Antiquities Authority. Schick, T. 2002. The Early Basketry and Textiles from the Northern Judean Desert. Atiqot 41: 223-239. Schmidt, P. 2010. Göbelki Tepe –The Stone Age Sanctuaries. New Results of Ongoing Excavations with a Special Focus on Sculptures and High Reliefs. Documenta Prehistorica XXXVll: 239-256. Schneider, J. 1989. Rumpelstiltskin’s Bargain. In A. Weiner and J. Schneider (eds), Cloth and the Human Experience: 177-213. Washington, D.C., Smithsonian. Schwartz, A. 1947. The Reel. Ciba Review 59: 2130-2167. Schwartzbaum, P., Silver, C., Wheatley, C., Dangas I. and Searight, A. 1980. Consolidating and Mounting of Chalcolithic Mural Painting Fragments from the Site of Teleilat Ghassul. Paris, UNESCO. Sebbane, M. 2014. The Hoard from Nahal Mishmar, and the Metal-working Industry in Israel in the Chalcolithic Period. In M. Sebbane, O. Misch-Brandl and D. Masters (eds), Masters of Fire: Copper Age Art from Israel: 114-136. Princeton, Princeton University Press. Seymour, J. 2001. The Forgotten Arts and Crafts. New York, Dorling Kindersley. Shamir, O. 2005. Textiles in the Land of Israel from the Roman Period until the Early Islamic Period in the Light of Archaeological Findings. Unpublished PhD thesis, Hebrew University of Jerusalem (Hebrew). Shamir, O. 2014. Textiles, Basketry and other Organic Artifacts of the Chalcolithic Period in the Southern Levant. In M. Sebbane, O. Misch-Brandl and D. Masters (eds), Masters of Fire: Copper Age Art from Israel: 139-152. Princeton, Princeton University Press. Shamir, O. 2015. Textiles from the Chalcolithic Period, Early and Middle Bronze Age in the Southern Levant. Archaeological Textile Review 57: 12-25. Shamir, O. and Baginski, A. 1998. Research into Early Textiles Discovered in the Land of Israel. Qadmoniyot 115: 53-60 (Hebrew). Shao, R., Zhu, X-Q., Barker, S. and Herd, K. 2012. Evolution of Extensively Fragmented Mitochondrial Genomes in the Lice of Humans. Genome Biological Evolution 4/11: 1088-1101. Sheffer, A. 1994. Textiles from Tomb 2 at En Gedi. Atiqot XXIV: 9*. Shimony, C. 1995. Fibre Identification. Atiqot XXVII: 204. Shugar, A. 2000. Archaeometallurgical Investigation of the Chalcolithic Site of Abu Matar, Israel: A Reassessment 201

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead of Technology and its Implications for the Ghassulian Culture. Unpublished PhD thesis, University of London. Singer, I. 2009. The Hittites and their Civilization. Jerusalem, Bialik Institute (Hebrew). Smith, R. and Hanbury-Tension, J. 1992. The Pottery Neolithic and Chalcolithic Periods at Pella. In A. McNicoll, P. Edwards, J. Hanbury-Tension, T. Potts, R. Smith, A. Walmsey and P. Watson (eds), Pella in Jordan 2. The Second Interim Report of the Joint University of Sydney and Wooster College Excavations at Pella 19821985: 17-27. Sidney, Mediterranean Archaeology. Soffer, O., Adovasio, J. and Hyland, D. 2000(a). The ‘Venus’ Figurines: Textiles, Basketry, Gender and Status in the Upper Palaeolithic. Current Anthropology 41: 511537. Sollas, W. 1924. Ancient Hunters and their Modern Representatives. London, Macmillan. Speir, R. 1970. From the Hand of Man. Boston, Houghton Mifflin. Tiedmann, E. and Jakes, K. 2006. An Exploration of Prehistoric Spinning Technology: Spinning Efficiency and Technology Transition. Archaeometry 48/2: 293-307. Tzori, N. 1958. Neolithic and Chalcolithic Sites in the Valley of Beth-Shan. Palestine Exploration Quarterly 90: 41-55. Veblen, T. 1967 (1899). The Theory of the Leisure Class. New York, Viking. Verhecken, A. 2010. The Moment of Inertia: a Parameter for the Functional Classification of Worldwide Spindle Whorls for all Periods. In E. AnderssonStrand, M. Gleba and U. Mannering (eds), North European Symposium for Archaeological Textiles, NESAT X: 348-363. Ancient Textile Series, vol. 5. Oxford: Oxbow. Verhecken, A. 2013. Spinning with the Hand Spindle: Analysis of the Mechanics and Implications of Yarn Quality. Archaeological Textile Review 55: 97-101. Vogel, T. 1989. Village Weaving in the Southern Hebron Hills. Ariel 75: 75-89. Vogelsang-Eastwood, G. 1987. A Re-examination of the Fibres from Ҁatal Hüyük Textiles. Oriental Carpet and Textile Studies. 111: 15-19. Vogelsang-Eastwood, G. 1992. The Production of Linen in Pharaonic Egypt. Leiden, Stichting Textile Research Centre.

Vogelsang-Eastwood, G. 1993(a). One of the Oldest Textiles in the World? Archaeological Textile Newsletter 16: 4-7. Vogelsang-Eastwood, G. 1993(b). Pharaonic Egyptian Clothing. Leiden, Brill. Vogelsang-Eastwood, G. 2000. Textiles. In P. Nicholson and I. Shaw (eds), Ancient Egyptian Materials and Technology: 268-298. Cambridge, Cambridge University. Voigt, M. and Meadows, R. 1983. Haji Firuz Tepe, Iran; the Neolithic Settlement University Museum Monograph. Philadelphia, University of Pennsylvania Press. Waetzhold, H. 2007. The Use of Wool for the Production of Strings, Ropes, Braided Mats and Similar Fabrics. In C. Gillis and M-L. Nosch (eds), Ancient Textiles: Production, Craft and Society: 112-121. Ancient textile Series 1. Oxford, Oxbow. Weindling, L. 1947. Long Vegetable Fibers. New York, Columbia University Press. Weir, S. 1970. Spinning and Weaving in Palestine. London, British Museum. Weir, S. 1976. The Bedouins. London, British Museum. Weiss, E. and Zohary, D. 2011. The Neolithic Southwest Asian Founder Crops. Current Anthropology 52/4: S237-S 254. Werker, E. 1995. Wood Identification. Atiqot XXVII: 203. Wiggermann, F. 2000. Lamaštu, Daughter of Anu, a Profile. In M. Stol (ed.), Birth in Babylonia and the Bible: Its Mediterranean Setting: 217-252. Groningen, Styx Publications. Wilson, T. 1896. The Swastika. Annual Report of the Smithsonian Institution: 763-996. Wincott-Heckett, E. 2007. Clothing Patterns as Constructs of the Human Mind: Establishment and Continuity. In C. Gillis and M-L. Nosch (eds), Ancient Textiles: Production, Crafts and Society: 208-214. Ancient Textile Series 1. Oxford, Oxbow. Wulff, H. 1966. Traditional Crafts of Persia. Cambridge, Mass, Massachusetts Institute of Technology. Zaitschek, D. 1959. Remains of Cultivated Plants from Horvat Betar (Beersheba). Atiqot 2: 48-52. Zohary, D., Hopf, M. and Weiss. E. 2012. Domestication of Plants in the Old World. Oxford, Oxford University.

202

Stylistic Devices and Exegetical Techniques in ‘Rewritten Bible’ Compositions Atar Livneh ‘Rewritten Bible’ texts have been the subject of much study in recent years (Falk 2007; Laato and van Ruiten 2008; White Crawford 2008; Machiela 2010; Zahn 2011; Zsengellér 2014). One of the primary questions discussed is the relation between the original source and the secondary artifact (e.g., Tov 1994; Alexander 1998:116118; Brooke 2000; Bernstein 2013a). This issue is closely associated with the methods ‘Rewritten-Bible’ authors employ to rework the biblical text and if and how these differ from the techniques whereby the biblical text was transmitted (e.g., Tov 1998; Ulrich 1998:88-89; Segal 2005; Brooke 2010:361-371; Zahn 2011). The much longer scholarly history of the latter discipline and growing acknowledgment of the similarities between the two enterprises have influenced the way in which the compositional techniques typical of ‘Rewritten Bible’ texts are defined. Thus, for example, additions (+) and omissions (-) and other devices characteristic of textual criticism—such as harmonization and word substitution—are identified and examined (e.g., Tov 1994; Brooke 1997; Zahn 2011:1-23; Bernstein 2013c). Scholars frequently refer to these as ‘exegetical techniques’ because the reworking imbues the original text with fresh meaning (e.g., Brooke 1997; Segal 1998; Falk 2007:21-25; Bernstein 2013c; but cf. Zahn 2011:12-17).

attention to the ‘poetical elements in Jubilees’ in the introduction to his commentary, he focused on the way recognizing them helps us to reconstruct the text. Although Doran (1989) contends that the book is primarily a literary rather than historical artifact, no systematic examination of its stylistic devices has yet been conducted. While Endres (1987:100) notes some, he concludes that ‘In terms of style his rewriting of this story [Jacob’s marriage to Leah and Rachel] demonstrates again that the author did not enjoy narration of a story; thus the dramatic elements, such as suspense and ambiguity, are often ‘remedied’ by omission in his rewriting of the biblical narrative’ (my italics). Contra this view, I wish to argue that the Jubilean author was not only aware of but also skilled in the art of narrative. The awareness and the skill are exemplified by his frequent use of stylistic devices that serve both aesthetic and exegetical functions. The significance of such a survey goes beyond Jubilees, suggesting that, together with the so-called ‘exegetical techniques’, these formed part of the set of tools at the disposal of ancient Jewish authors, who reshaped and reinterpreted the biblical texts. In what follows, rather than addressing all of the features and the passages in which they occur, I shall merely illustrate some.

While understanding these factors is important for analyzing the form and exegesis characteristic of ‘Rewritten Bible’ texts, some ancient authors also blended them with stylistic devices so skillfully that the two cannot always be distinguished. In contrast to the exegetical techniques, the stylistic features and their contribution to the author’s message (based on his exegesis) have been relatively neglected—the exception to the rule being the Genesis Apocryphon (e.g., Nickelsburg 1998; Falk 2007:26-106; Weingold 2010; Bernstein 2013b, 2013d). One of the reasons for this is the fact that many of the Qumran ‘Rewritten Bible’ compositions are so fragmentary that even their basic literary character, scope, and structure cannot be identified. The broken lines further preclude any recognition of the stylistic means employed in smaller literary units (e.g., chiasm, parallelism). Rather surprisingly, some of the works, which have been preserved in a complete form, have also suffered a similar fate.

Repetition of word(s) from the biblical source Widely attested in the Hebrew Bible (see recently Yona 2013), repetition is also a common feature of Jubilees. On occasion, a clear allusion to a biblical verse is followed by a clause employing one or more words from the same passage. Thus, for example, the retelling of the episode of the covenant between the pieces contains the following sequence: ‘He [God] brought him [Abraham] outside and said to him: ‘Look at the sky and count the stars if you can count them’. When he had looked at the sky and seen the stars, he said to him: ‘Your descendants will be like this’ (Jub. 14:4-5 [VanderKam 1989; my italics]). While God’s instruction to Abraham: ‘look at the sky …’ is drawn from Gen 15:5, Jubilees adds that Abraham followed it, thus highlighting his obedience. Employed in the Samaritan Pentateuch— in the context of Moses’ bringing the plagues in accordance with God’s instructions, for example (cf. SAM Exod 9:19), this compositional technique has been adduced by text-critical scholars, who often regard

This chapter examines Jubilees (C2 BCE) as a test case. Although early on Charles (1902:xlii-xliii) paid brief 203

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead the ‘command and its fulfillment’ as a sub-type of harmonization (e.g., Tov 2001:85-89). Consisting of two elements—the insertion of additional material and repetition—the first is commonly examined in the study of ‘exegetical techniques’ typical of ‘Rewritten Bible’ texts (e.g., Vermes 1986:308; Alexander 1988:116118; Falk 2007:23), The second is also a staple subject in handbooks relating to the art of biblical narrative (e.g., Alter 1981:88-113; Sternberg 1987:365-440; BarEfrat 1989:211-216; Polak 1999:59-80). Jubilees is thus distinctive in combining both.

of the words ‘when evening came’. Hereby, he presents Noah in a more favourable light, stressing his innocence by underscoring that the transgression was an ‘unconscious’ one (van Ruiten 2000:283; Werman 2015:235). In its retelling of the Exodus narrative, Jubilees (48) seeks to present the story in terms of divine vengeance upon Israel’s enemies, employing the root b-q-l (‘revenge’) six times (Berger 1981:542). Although this does not appear in the biblical account of the Exodus, it is suggested by the synonym šĕpāṭîm (judgments) in Exod 12:12 (cf. Ezek 25:17, which combines the two terms; Livneh 2011:171-174).

In Jubilees’ reworking of the account of Potiphar’s wife, the command that is obeyed serves as proof and evidence:

The nouns ‘love’ and ‘son’ recur in Jub. 19:15-31 and ‘the days of X’s life’ in Jub. 35:6-8 form similar examples, albeit in extra-biblical additions.

When that woman saw that he would not lie with her, she accused him falsely to his master: ‘Your Hebrew slave whom you love wanted to force me so that he could lie with me. When I shouted, he ran away, left his clothes in my hands when I grabbed him …’ When the Egyptian saw Joseph’s clothes … he believed what his wife said. (Jub. 39:10-11 [VanderKam 1989; my italics]).

Inclusio On several occasions, Jubilees uses the same word(s) and/or motif at the beginning and end of an account, thereby forming an inclusio. Thus, for example, it reworks Gen 25:27 [NJPS]— ‘… Esau became a skillful hunter, a man of the outdoors; but Jacob was a mild man who stayed in camp’—by placing Jacob at the opening and conclusion in order to accentuate his righteousness: ‘Jacob was perfect and upright, while Esau was a harsh, rustic, and hairy man. Jacob used to live in tents’ (Jub. 19:13 [VanderKam 1989]).

While Potiphar’s wife’s report generally follows Gen 39:17-18, the second reference to Joseph’s clothes serves as proof of physical evidence. Supporting Potiphar’s wife’s claim, the repetition establishes that Potiphar’s judgment was based on what he heard and saw. Together with the additional material, it thus imparts a new significance to the biblical text.

A more complex example occurs in Jub. 40:8:

Leitmotif

(A) So Joseph became ruler over the entire land of Egypt. All of the pharaoh’s princes, all of his servants, and all who were doing the king’s work loved him because he conducted himself in a just way. (B) He was not arrogant, proud. (C) or partial, nor did he accept bribes. (A’) because he was ruling all the people of the land in a just way. (VanderKam 1989).

Unlike the previous device, the leitmotifs in Jubilees are not necessarily words/roots derived from the principle biblical source. The account of Noah’s intoxication, for example, employs three variations on the root ‘to sleep’: When evening came, he went into his tent. He lay down drunk and fell asleep. He was uncovered in his tent as he slept. Ham saw his father Noah naked and went out and told his two brothers outside. Then Shem took some clothes, rose—he and Japheth— and they put the clothes on their shoulders as they were facing backwards. They covered their father’s shame as they were facing backwards. When Noah awakened from his sleep, he realized everything that his youngest son had done to him. (Jub. 7:7-10 [VanderKam 1989; my italics]).

Sections A and A’ both refer to Joseph’s ruling over all the land of Egypt. This duplication resembles the repetition within the biblical story itself: ‘I [Pharaoh] put you [Joseph] in charge of all the land of Egypt … Thus he placed him over all the land of Egypt’ (Gen 41:41, 43). However, Jubilees adds that Joseph ruled justly. Sections B and C elucidate this fact, referring to his lack of arrogance (B) and refusal to accept bribes (C)—each of these traits being depicted via a pair of synonyms. Joseph is thus configured as a just ruler via an inclusio and pairs of synonyms.

Although Gen 9:24 suggests that Noah was asleep (‘When Noah woke up …’ [NJPS]), this fact is nowhere explicitly noted in the biblical story (van Ruiten 2000: 283). As Werman (2015:235) observes, the depiction of Noah as sleeping accords with the author’s insertion 204

A. Livneh: Stylistic Devices and Exegetical Techniques in ‘Rewritten Bible’ Compositions In some small literary units, the motif that reoccurs at the beginning and end is elucidated by the middle member (cf. the structure of Exod 25:31; Deut 26:14). In his testament, for example, Noah depicts the circumstances preceding the flood in the following words: ‘When everyone sold himself to commit injustice and to shed innocent blood, the earth was filled with injustice’ (Jub. 7:26). Herein, the motif injustice (ḥāmās) which twice depicts humanity’s sins (Gen 6:11, 13), is elaborated by the term ‘shedding blood’ drawn from Gen 9:6. Human crimes prior to the flood are thus clarified in light of the divine ordinance given immediately after the cataclysmic event (van Ruiten 2000:300-301; Werman 2015:237).

(A) No one who consumes blood or who sheds blood on the earth will be left. (B) He will be left with neither descendants nor posterity living beneath heaven. (C) because they will go into sheol. (D) and will descend into the place of judgment. (E) All of them will depart into deep darkness through a violent death. (Jub. 7:29 [VanderKam 1989]; my numbering). Following Lev 17:12-13, the first statement juxtaposes eating and shedding blood (van Ruiten 2000:300-301). The brief depiction of the sins (section A) is set within a passage (A–E), whose primary interest is in the nature of the punishment. The latter thus explains the meaning of the somewhat obscure term karet in detail (cf. Werman 2015:238). It is structured in a poetical style, the first statement paralleling the second and both referring to the disappearance of the sinner and/or his progeny from upon the earth in a formula drawn from the curses in Deuteronomy (sections A and B; cf. Deut 28:63, 29:19, 30:18). Sections C–E all describe the sinner/his descendants’ residence beneath the earth via a specific designation and/or typical trait. With the exception of the phrase ‘place of judgment’, the parallels between Sheol/going down/darkness/death are all based on standard biblical locutions (cf. Ps 55:16; Job 17:13, 16; Prov 5:5). While C, D, and E exhibit parallels, the latter (E) complements the two preceding members (C, D) by noting not only the (final) destination of the sinners but also the ‘violent death’ that will remove them from ‘upon the earth’ (cf. A, B).

Similar examples of the A–B–A structure, in which the middle member unpacks the motif that envelops the unit, also occur in Jub. 7:23, 33:21-23 (Livneh 2014: 3636), and 36:2, 4 (Livneh 2011b:189). For further examples of inclusio in the book, see Jub. 12:29; 19:15-31 (Endres 1987:25-28; van Ruiten 2010:244-251), 25:3-11, 27:14-17, 35:13-17 (the latter inclusio occurs in 4Q223-224 II 3-13— but not in the Geʿez translation). Repetition via synonymous words within prose clauses While synonymous words are more common in prayers and blessings/curses in Jubilees (e.g., 21:21, 24:28, 36:10; but cf. also the ‘apocalypse’ in Jubilees 23), the author likewise employs this device in his rewriting of the biblical narratives. In addition to the depiction of Joseph as a just ruler noted above (§3), we may note Jubilees’ reworking of Gen 17:14 [NJPS]: ‘And if any male who is uncircumcised fails to circumcise the flesh of his foreskin, that person shall be cut off from his kin; he has broken My covenant’. Herein, the consequences of disobedience are adduced via three synonymous terms, thereby stressing the graveness of the divine punishment: ‘Anyone who is born, the flesh of whose private parts has not been circumcised … (is meant) for destruction (la-ʾamāsəno), for being destroyed (wa-la-taḥagwəlo) from the earth, and for being uprooted (wa-la-taśarrəwo) from the earth because he has violated the covenant of the Lord our God’ (Jub. 15:26 [VanderKam 1989; my italics]). While this might be defined as an ‘expansion’, the term ‘expansion’ does not adequately describe how the repetition is fundamental to the exegesis.

If this passage is examined from the perspective of ‘exegetical techniques’, it appears to elaborate the karet referred to in Lev 17:10-14 in Deuteronomistic language and standard phrases from Psalms and the wisdom literature. However, this reading fails to take account of the manner in which the biblical components have been arranged. The poetical structure highlights the bitter end of those who eat and/or shed blood on the earth, thereby clearly playing an exegetical function. For further examples of parallelism, see especially the blessings/prayers (e.g., Jub. 21:22, 25:3, 31:8; Charles 1902:xliii) as well as the ‘apocalypse’(e.g., Jub. 23:11, 23, 28) and less commonly in the narrative sections (e.g., Jub. 38:1). Chiasmus

Parallelism

Jubilees also employs the well-known device of chiasmus. On two occasions, it does so in order to stress the contrast between younger/older siblings and their rights. Thus, for example, it reconfigures Laban’s words ‘It is not the practice in our place to marry off the younger before the older’ (Gen 28:26) chiastically: ‘No one should give his younger daughter before his

Jubilees’ fondness for depicting the divine punishment via synonyms is further reflected in its reworking of the karet imposed on those who eat and/or shed blood (Jub. 7:29). While Charles (1902:63) observes that the passage is written in verse, its structure has not been discussed in detail: 205

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead older one, but he should first give the older and after her the younger’ (Jub. 28:6). It also states that Esau’s sons complained to their father: ‘Why is it that when you are the older and Jacob the younger your father gave Jacob the birthright and neglected you?’(Jub. 37:2 [VanderKam 1989]). While the juxtaposition of Esau the older/Jacob the younger is drawn from Gen 27:15, Jubilees inserts the Esau-Jacob-Jacob-Esau sequence.

Here, sections A and A’ depict Canaan’s residence in Lebanon. The geographic portrayal of this territory as stretching as far as the (stream of) Egypt is taken from Judg 3:3; 1 Kgs 8:65; 1 Chr 13:5; and 2 Chr 7:8. These details envelop sections B and B’, which relate to the curse destined to fall upon Canaan for choosing Lebanon as his abode. At the heart of the concentric structure stands the justification for the curse—the land of Lebanon belongs to Shem and his sons (C). The central place of the latter statement reflects the importance Jubilees ascribed to the notion that Canaan belonged to Shem’s descendants—and them alone.

The same device occurs in the reworking of Isa 3:5—‘So the people shall oppress one another, each oppressing his fellow: The young shall bully the old; and the despised [shall bully] the honored’. [NJPS]. Seeking to highlight the group conflict, the Jubilean author repeats the references to the first two sectors (young and old), arranging them in a chiastic structure: ‘One group will struggle with another—the young with the old, the old with the young (warāzut məsla liqānāt wa-liqānāt məsla warāzut); the poor with the rich, the lowly with the great; and the needy with the ruler—regarding the law and the covenant’ (Jub. 23:19 [VanderKam 1989; my italics]; cf. Kister 1986:6).

While this extra-biblical passage is constructed out of building blocks from various biblical verses, the same device also occurs within passages that allude to a specific biblical source. One example occurs in the reworking of the story about Potiphar’s wife in Gen 39:8-10: ‘But he [Joseph] refused. He said to his master’s wife, ‘… How then could I do this most wicked thing, and sin before God?’ And much as she coaxed Joseph day after day, he did not yield to her request to lie beside her, to be with her’ (NJPS). This both starts and culminates with the notation that Joseph refused to lie with Potiphar’s wife:

For further examples of chiasmus in Jubilees, see Jub. 37:20-23 (Livneh 2012:85-91); Jub. 48:5-6 (HalpernAmaru 2015:69-70); Jub. 48:13-14a (Segal 2007:218219).

(A) But he did not surrender himself. (B) He remembered the Lord and what his father Jacob would read to him from the words of Abraham. (C) that no one is to commit adultery with a woman who has a husband; that there is a death penalty which has been ordained for him in heaven before the most high Lord. The sin will be entered regarding him in the eternal books forever before the Lord. (B’) Joseph remembered what he had said. (A’) and refused to lie with her. (Jub. 39:6-8 [VanderKam 1989; numbering added]).

Concentric circles The account of Canaan’s dwelling in Shem’s lot in Jub. 10:29-33 is structured as a concentric circle: (A) 10:29 When Canaan saw that the land of Lebanon as far as the stream of Egypt was very beautiful, he did not go to his hereditary land to the west of the sea. He settled in the land of Lebanon, on the east and west, from the border of Lebanon and on the seacoast. (B) 10:30 His father Ham and his brothers Cush and Mizraim said to him: ‘You have settled in a land which was not yours and did not emerge for us by lot. Do not act this way, for if you do act this way both you and your children will fall in the land and be cursed with rebellion, because you have settled in rebellion and in rebellion your children will fall and be uprooted forever. (C) 10:31 Do not settle in Shem’s residence because it emerged by their lot for Shem and his sons. (B’) 10:32 You are cursed and will be cursed more than all of Noah’s children through the curse by which we obligated ourselves with an oath before the holy judge and before your father Noah’. (A’) 10:33 But he did not listen to them. He settled in the land of Lebanon—from Hamath to the entrance of Egypt—he and his sons until the present. (VanderKam 1989; my numbering).

While A and A’ follow the primary biblical source, they frame an explanation of Joseph’s piety unattested in the Bible—namely, that Joseph did not surrender to Potiphar’s wife’s charms because he remembered his father’s instruction (sections B and B’). The central section (C) depicts the contents of the patriarch’s instruction—the prohibition against adultery and its punishment (Livneh 2013:12). While the words in this section are not drawn from the principle biblical text, they obviously constitute an interpretation of Joseph’s claim that lying with his master’s wife would be a ‘sin before God’ (cf. Gen 39:9). By placing the ordinance at the centre and elaborating it in more detail than the other sections, Jubilees indicates that Joseph obeyed the divine commandment. Hereby, he not merely connects the Genesis story with the law—a well-known Jubilean tendency—but literarily places the ordinance at the heart of the narrative. 206

A. Livneh: Stylistic Devices and Exegetical Techniques in ‘Rewritten Bible’ Compositions For further examples of concentric cycles, see Jub. 30:56 (Werman 1997:10; Werman 2015:411) and 36:22-24 (Livneh 2013:13-14).

Jubilees. The author skillfully interweaves them in both small-scale units and longer accounts throughout the book. While we may never know whether or not he ‘enjoyed the art of narration’, he was evidently well aware of it and well versed in it. While the stylistic devices and literary structures in the book obviously serve aesthetic purposes, they also play a significant role in constructing the meaning of the text. Jubilees almost invariably reworking biblical sources, the stylistic devices thus also serve exegetical purposes.

2+3 structure This device typically adduces two similar items followed by an essentially different third. Like the concentric circle, Jubilees employs this method on occasion in accounts with no biblical parallel. Thus, for example, it relates that Noah’s sons each built a city, contrasting Shem with Ham and Japhet in this regard:

The test case of Jubilees suggests that the ‘pool’ of compositional techniques upon which (some of) the ‘Rewritten Bible’ authors drew is broader than suggested by the current lists of ‘exegetical techniques’. The latter lists rely heavily on the field of textual criticism, However, ‘stylistic devices’ belong to a different modern discipline—namely biblical rhetoric and stylistics. The two phenomena of exegetical techniques and stylistic devices are interwoven in Jubilees, and they both serve to rework the biblical text. Consequently, the modern distinction between them is difficult to maintain in the context of ‘Rewritten Bible’ texts.

(A) When Ham realized that his father had cursed his youngest son, it was displeasing to him that he had cursed his sons. He separated from his father ... He built himself a city and named it after his wife Neelatamauk. (B) When Japheth saw (this), he was jealous of his brother. He, too, built himself a city and named it after his wife Adataneses. (C) But Shem remained with his father Noah. He built a city next to his father at the mountain. He, too, named it after his wife Sedeqatelebab. (Jub. 7:13-16 [numbering added]).

References

This passage is inspired by Gen 4:17: ‘And he then founded a city, and named the city after his son Enoch’— the kin after whom the city is named being changed from son to a wife and the deed ascribed to each of Noah’s sons. While the same formula recurs in the three sequential depictions, the location of the towns and their names differ, the repeating formula thereby drawing an analogy between Ham and Japheth and a contrast between these two brothers and Shem. While scholars have long noted the comparison (VanderKam 2001:39-41; VanderKam 2002:485-486) the 2+3 pattern, which serves to draw the reader’s attention to Shem and his righteous behaviour, has not been identified.

Alexander, P. S. 1988. Retelling the Old Testament. In B. Lindars, (ed.), It is Written: Scripture Citing Scripture: Essays in Honour of Barnabas Lindars, 99-121, Cambridge, Cambridge University Press. Alter, R. 1981. The Art of Biblical Narrative, New York, Basic Books. Bar-Efrat, S. 1989. Narrative Art in the Bible. Bible and Literature Series 17, Sheffield, Sheffield Academic Press. Berger, K. 1981. Das Buch der Jubiläen. Jüdische Schriften aus hellenistisch-römischer Zeit, 2/3, Gütersloh, G. Mohn. Bernstein, M. 2013a. ‘Rewritten Bible’: A Generic Category which has Outlived its Usefulness? In M. Bernstein, Reading and Re-reading Scripture at Qumran: 39-62. Studies on the Texts of the Desert of Judah 107, Leiden, Brill. Bernstein, M. 2013b. From the Watchers to the Flood: Story and Exegesis in the Early Columns of the Genesis Apocryphon. In M. Bernstein, Reading and Re-reading Scripture at Qumran, 151-74, Studies on the Texts of the Desert of Judah 107, Leiden, Brill. Bernstein, M. 2013c. Rearrangement, Anticipation and Harmonization as Exegetical Features in the Genesis Apocryphon. In M. Bernstein, Reading and Re-reading Scripture at Qumran, 175-94. Studies on the Texts of the Desert of Judah 107, Leiden, Brill. Bernstein, M. 2013d. Is the Genesis Apocryphon a Unity? What Sort of Unity Were You Looking for? In M. Bernstein, Reading and Re-reading Scripture at Qumran, 239-65, Studies on the Texts of the Desert of Judah 107, Leiden, Brill.

Jubilees uses the same literary device to distinguish between God’s chosen people and the nations in 15:30: ‘For the Lord did not draw near to himself either Ishmael, his sons, his brothers, or Esau … But he chose Israel to be his people’. This unit is heavily dependent upon Deuteronomy (esp. 7:6-7, but cf. also 10:15, 14:2). The synonyms ‘draw near/choose’ are adopted from Num 15:6. The contrast of God’s favouring Jacob but not Esau follows Mal 2:2-3. However, the reference to Ishmael in this context is a Jubilean innovation. The innovation presents Ishmael and Esau as the two first objects of God’s acts while Israel constitutes the climax. God chose neither Ishmael nor Esau but a third, unique nation—Israel. Conclusion Despite the brevity of this survey, it adequately illustrates the nature and use of stylistic devices in 207

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead Brooke, G. J. 1997. Exegetical Strategies in Jubilees 1-2:  New Light from 4QJubileesa. In M. Albani, J. Frey, and A. Lange (eds), Studies in the Book of Jubilees, 39-57, Texte und Studien zum Antiken Judentum 65, Tübingen, Mohr Siebeck. Brooke, G. J. 2000. ‘Rewritten Bible.’ In L. H. Schiffman and J. C. VanderKam (eds), Encyclopedia of the Dead Sea Scrolls, 2 vols., 2:77-81, Oxford, Oxford University Press. Brooke, G. J. 2010. Genre Theory, Rewritten Bible and Pesher, Dead Sea Discoveries 17:361-86. Charles, R. H. 1902. The Book of Jubilees or the Little Genesis, London, Adam and Charles Black. Doran, R. 1989. The Non-Dating of Jubilees, Journal for the Study of Judaism 20:1-11. Endres, J. C. 1987. Biblical Interpretation in the Book of Jubilees, Catholic Biblical Quarterly Monograph Series 18, Washington, Catholic Biblical Association. Falk, D. K. 2007. The Parabiblical Texts: Strategies for Extending the Scriptures among the Dead Sea Scrolls, Companion to the Qumran Scrolls 8, Library of Second Temple Texts 63, London, T&T Clark. Halpern-Amaru, B. 2015. The Perspective from Mt. Sinai: The Book of Jubilees and Exodus, Journal of Ancient Judaism Supplements 21, Göttingen, Vandenhoeck & Ruprecht. Kister, M. 1986. Concerning the History of the Essenes. Tarbiz 56:1-18 (Hebrew). Laato, A., and van Ruiten, J. T. A. G. M. 2008. Rewritten Bible Reconsidered: Proceedings of the conference in Karkku, Finland, August 24-26, 2006, Studies in Rewritten Bible 1, Winona Lake, IN, Eisenbrauns. Livneh, A. 2011a. Judgment and Revenge: The Exodus Account in Jubilees 48, Revue du Qumran 98:161175. Livneh, A. 2011b. ‘Love your fellow as yourself ’: The Interpretation of Leviticus 19:17-18 in the Book of Jubilees, Dead Sea Discoveries 18:173-99. Livneh, A. 2012. ‘Can the Boar Change its Skin?’ Esau’s Speech in Jubilees 37:18-23, Henoch 34:75-94. Livneh, A. 2013. Not at first sight: Gender Love in Jubilees, Journal for the Study of the Pseudepigrapha 23: 3-20. Livneh, A. 2014. ‘United We Stand’: Two Jubilean Clan Rosters Adapted to Demonstrate Israelite Cohesion, Journal of Jewish Studies 65:26-37. Machiela, D. 2010. Once More, with Feeling: Rewritten Scripture in Ancient Judaism—A Review of Recent Developments, Journal of Jewish Studies 61:308-20. Nickelsburg, G. W. 1998. Patriarchs Who Worry about their Wives: A Haggadic Tendency in the Genesis Apocryphon. In M. E. Stone and E. G. Chazon (eds), Biblical Perspectives: Early Use and interpretation of the Bible in Light of the Dead Sea Scrolls, 137-58, Studies on the Texts of the Desert of Judah 28, Leiden, Brill. Polak, F. 1999. Biblical Narrative: Aspects of Art and Design, 2nd ed., Biblical Encyclopedia Library XI, Jerusalem, Bialik Institute (Hebrew).

Ruiten, J. T. A. G. M. van 2000. Primaeval History Interpreted: The Rewriting of Genesis 1-11 in the Book of Jubilee, Journal for the Study of the Pseudepigrapha Supplement Series 66, Leiden, Brill. Ruiten, J. T. A. G. M. van 2010. Abraham in the Book of Jubilees: The Rewriting of Genesis 11:26-25:10 in the Book of Jubilees 11:14-23:8, Journal of Ancient Judaism Supplements 161, Leiden, Brill. Segal, M. 1998. Biblical Exegesis in 4Q158: Techniques and Genre, Textus 19:45-62. Segal, M. 2005. Between Bible and Rewritten Bible. In M. Henze (ed.), Biblical Interpretation at Qumran, 10-18, Grand Rapids, Eerdmans. Segal, M. 2007. The Book of Jubilees: Rewritten Bible, Redaction, Ideology and Theology, Journal for the Study of Judaism Supplement Series 117, Leiden, Brill. Sternberg, M. 1987. The Poetics of Biblical Narrative: Ideological Literature and the Drama of Reading, Indiana Studies in Biblical Literature, Bloomington, Indiana University Press. Tov, E. 1994. Biblical Texts as Reworked in Some Qumran Manuscripts with Special Attention to 4QRP and 4QParaGen-Exod. In E. Ulrich and J. C. VanderKam (eds), The Community of the Renewed Covenant: The Notre Dame Symposium on the Dead Sea Scrolls, 111-34, Notre Dame, University of Notre Dame Press. Tov, E. 1998. Rewritten Bible Compositions and Biblical Manuscripts, with Special Attention to the Samaritan Pentateuch, Dead Sea Discoveries, 5:33454. Tov, E. 2001. Textual Criticism of the Hebrew Bible, 2nd ed., Minneapolis, Fortress. Ulrich, E. 1998. The Dead Sea Scrolls and the Biblical texts. In P. W. Flint and J. C. VanderKam (eds), The Dead Sea Scrolls after Fifty Years: A Comprehensive Assessment, 2 vols. 1:79-100, Leiden, Brill. VanderKam, J. C. 1989. The Book of Jubilees. Corpus Scriptorum Christianorum 510-11, Scriptores Aethiopici 87-88, Leuven, Peeters. VanderKam, J. C. 2001. The Book of Jubilees, Sheffield, Sheffield Academic Press. VanderKam, J. C. 2002. Putting Them in Their Place: Geography as an Evaluative Tool. In J. C. VanderKam, From Revelation to Canon: Studies in the Hebrew Bible and Second Temple Literature, 476–99. Leiden, Brill. Vermes G. 1986. Biblical Midrash. In G. Vermes, F. Millar, and M. Goodman (eds), The History of the Jewish People in the Age of Jesus Christ, 3 vols. 3:1:308-41, Edinburgh, T&T Clark. Weingold, M. 2010. One Voice or Many? The Identity of the Narrators in Noah’s Birth Story (1QapGen 1-5.27) and the ‘Book of the Words of Noah’ (1QapGen 5.2918.23), Aramaic Studies 8:89-105. Werman, C. 1997. Jubilees 30: Building a Paradigm for the Ban of Intermarriage, Harvard Theological Review 90:1-22. 208

A. Livneh: Stylistic Devices and Exegetical Techniques in ‘Rewritten Bible’ Compositions Werman, C. 2015. The Book of Jubilees: Introduction, Translation, and Interpretation, Jerusalem, Yad BenZvi (Hebrew). White Crawford, S. 2008. Rewriting Scripture in Second Temple Times, Grand Rapids, Eerdmans. Yona, S. 2013. The Many Faces of Repetition: Basic Patterns of Repetition in Construct-State Expressions in Biblical, Post-Biblical and Ancient Near Eastern Rhetoric, Beer Sheva, Ben-Gurion University Press (Hebrew).

Zahn, M. M. 2011. Rethinking Rewritten Scripture: Composition and Exegesis in the 4QReworked Pentateuch manuscripts, Studies on the Texts of the Desert of Judah 95, Leiden, Brill. Zsengellér, J. 2014. Rewritten Bible after Fifty Years: Texts, Terms, or Techniques?: A Last Dialogue with Geza Vermes, Journal for the Study of Judaism Supplement Series 166, Leiden, Brill.

209

Conus Ornaments from Tel Bareqet in an Early Bronze Age Near East Context Daniella E. Bar-Yosef Mayer, Sarit Paz and Yitzhak Paz Abstract The discovery of 16 Conus apex beads from the Early Bronze Age II site of Tel Bareqet in central Israel, prompted research concerning this type of personal ornament. Theses ornaments were made of Indo-Pacific Conus shells, and they were discovered in numerous third millennium BCE sites in the Levant, the Sinai Desert, as well as in Mesopotamia, suggesting long-range contacts. The existence of a workshop of such artifacts in Oman might point to their actual origin. Ethnographic analogies, coupled with the size and distribution of these artifacts, suggest that these were prestige items.

would have been displayed on the body or on clothes, possibly with a specific meaning attached. These items are the focus of this paper.

Introduction Conus shells are known in the archaeological record form as early as the Middle Stone Age (from Border Cave: Villa et al. 2012), and from a few Upper Palaeolithic sites in Europe (Vanhaeren and d’Errico 2006). In the Near East they were found in the Early Ahmarian levels of Ksar Akil (Bosch et al. 2015), from the Ahmarian of Üçagızlı Cave (Stiner 2013), and Hayonim Cave (personal observation). Usually, these objects form a minute component in each of the Palaeolithic shell assemblages. Consequently, they did not attract much attention thus far in Palaeolithic research (contra BarYosef Mayer, 2015).

Conus shells and conus beads Conus mollusks are today famous for the beauty of their shells, which are very valuable to collectors, and notorious as deadly venomous animals. At the same time, the venom of certain Conus species has been developed in recent years into a potent pain reliever called Prialt®. A recent molecular phylogenetic study concludes that the family Conidae Fleming, 1822, which Linnaeus identified as including 30 species, is made of four genera, and 803 species (Puillandre et al. 2015). The genus Conus Linnaeus, 1758 contains 85% of the known species within the family Conidae. According to that study, the genus Conus is divided into many sub-genera, and most inhabit tropical and subtropical waters. Today a single species, Conus Museum 60 ventricosus Gmelin, 1791 (previously also referred to as C. mediterraneus), inhabits the Mediterranean. The Red Sea contains many of the Indo-Pacific species in this group, with over 40 different species (Bosch et al. 1995, Dekker and Orlin 2000). Twenty-eight species are known from the Gulf of Eilat/Gulf of Aqaba (Heiman 2002).

Conus shells are more common in Neolithic sites, especially Red Sea species from the desert regions of the Levant (Bar-Yosef Mayer 1997; Spatz et al. 2014). In most occurrences, the shells were perforated, and they undoubtedly served as personal ornaments. However, it should be noted that many different Conus shell species were collected, and they were made into several types of beads (Bar-Yosef Mayer 1997). During the Early Bronze Age, however, Conus shells were collected apparently for more prestigious ornaments. This conclusion is based upon the fact that these shells are significantly larger than the older shell beads. The artifact discussed is called a Conus disc, a Conus ring, or a Conus apex bead. In all cases, it is the part of the Conus shell that consists of the spire, where the rest of the body has been removed, and a hole opened in its center (Fig. 1). The result is a round ring, varying in outer diameter depending on whether it was made of a large or a small species of the Conidae family. The central perforation varies in diameter (probably depending on the size of the drill used to make the hole). The thickness is also variable, depending on the size determined by the bead maker. Mostly, large specimens were collected, and they were made into large and impressive artifacts that

Conus shells were used as beads since prehistoric times, usually by perforating a hole in their apex, and sometimes by turning them into disk beads, by removing a part of their body whorl and drilling a hole in the apex. The apex is also sometimes perforated naturally as a result of abrasion on the shore. Rarely, more elaborate manufacturing practices are evident, such as filing of the body whorl (Bar-Yosef Mayer 1997: Fig. 4). These types of Conus beads are very abundant in many periods and in many regions of the world. They are usually made of relatively small Conus species. The terms Conus disc, Conus ring, and Conus apex bead 210

D. E. Bar-Yosef Mayer et al.: Conus ornaments from Tel Bareqet in an Early Bronze Age

Figure 1. The Conus apex beads from Bareqet. Each bead is viewed from two sides. Photography and layout: Oz Rittner.

least two occupation phases dated to Early Bronze Age II (EBII, ca. 3100-2850, Regev et al. 2012). Both phases lie directly on the bedrock, with no later strata above. The architectural remains of this ca. 5 hectares site included fortifications and domestic compounds, which consist of dwelling and storage units and courtyards, streets and alleys, and rock-cut installations. Finds included typical EBII pottery vessels, some of them complete, ground stone tools, various bone and copper implements, some textile remains, flint blades and cores, and beads made of stone, metal, and shell.

refer to a specific type of a Conus ornament present in the Early Bronze Age, or third millennium BCE in the Near East. These were made of much larger and heavier Conus species, and they can reach a diameter of about 4 cm. These are discussed below. Conus apex beads in Early Bronze Age sites The ornaments from Tel Bareqet To date, Tel Bareqet is the northern-most site in the southern Levant, in which Conus apex beads were found. The site is located in central Israel, and its lower part was excavated as a part of salvage excavations in 2004 (Paz and Paz 2007). In 2006 and 2008 a small-scale community excavation was conducted in the higher part of the site (Paz, 2010). It is a unicultural site with at

An assemblage of 16 Conus apex beads was discovered in the northwestern quarter of the lower part of the site, where remains of large domestic compounds were discovered near the city wall, separated by streets and courtyards. The beads were found on the rock surface 211

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead in Locus 258 (square DD24) that appears to have been the southernmost part of broadroom 286. However, the walls in this part were not preserved.

Salah, C. arenatus was represented by a complete shell and another consisting of only the spire. Complete shells of C. taeniatus, C. tessulatus and C. textile may have been collected as raw material for producing Conus apex beads, while C. coronatus and C. parvatus, which are smaller, probably served to produce other types of Conus beads. Thus, it seems that Conus apex beads were items placed as grave goods in Jericho, Bab-edh-Dhra and the nawamis sites, but potentially may have been produced in the Early Bronze Age habitation sites of Sheikh Muhsein and Nebi Salah. Unfortunately, apart from complete and broken shells, only one shell with an incision can support this idea. Ground stone, flint, and copper artifacts from these sites do not include artifact types associated specifically with this task, nor were finished Conus apex beads found in these sites. Furthermore, the sites are inland within the Sinai Peninsula, away from the coast (Beit Arieh 2003, and see further discussion below).

The outer diameter of the beads ranges between 17 and 34 mm (average 25.5 mm), while the diameter of the holes ranges between 3 and 9 mm (average 5.75 mm). The height of the beads ranges from 3 to 6 mm. Because they were all worked by grinding all around and by perforation of the center (apex) in the spire, it was impossible to determine the species of which they were made. Yet their large size indicates that they were not made of the Mediterranean C. ventricosus. On three specimens, there are traces of coronation at the edge of the spire. This feature further supports an Indo-Pacific range origin, including the Red Sea. Observations under a binocular microscope revealed that on a few of the specimens there are slight indentations at the edge of the hole. These indentations suggest that the beads were strung or sewn onto clothes or other artifacts (Fig. 1). The lack of signs in others does not necessarily mean that they were not strung (Pauc and Reinhard, 2002). Most shells also exhibit a sheen that is the result of wear.

Other occurrences in the Near East

Additional specimens, also associated with graves, were found in the nawamis sites of Southern Sinai (Bar-Yosef et al. 1977, 1986). These species were identified when possible, and they included C. arenatus and C. taeniatus. Both of these species come from the Red Sea (Bar-Yosef Mayer 1999). When specimens were measured, they ranged in diameter between 6 and 38 mm.

Conus apex beads were identified in various sites of the third to second millennia BCE. In Mesopotamia, the site of ‘Usieyh, dated to early second millennium B.C. (Oguchi 1992:61), is particularly rich in shell rings. These are the perforated and polished Conus spires. Oguchi (1992:68-69) distinguishes between type A, which have very large holes, and indeed look more like rings that might fit on human fingers and type B. The latter are characterized by smaller perforations, and they are consistent with the specimens known from the Levant described above. At ‘Usiyeh, of a total of 2509 shell rings, 754 belong to his type B ranging between 14 and 31 mm in external diameter. Importantly, there is no evidence for their manufacture on site, and they were all imported. Their context within the site is unclear, and they are only described as coming from underground structures. Several other sites in Mesopotamia also contain this type of shell artifact, especially along the Euphrates, but apparently not in such large numbers (Oguchi 1992:69-70; Moorey 1994:133). Graves of the third millennium Umm an-Nar in the United Arab Emirates contained a single Conus apex bead from Tomb V. Additional sites along the Persian Gulf coast are mentioned as containing similar artifacts (Reese 1991:184, fig. 240).

In southern Sinai Conus shells were also present in Early Bronze Age habitation sites. At Sheikh Muhsein three complete shells of C. taeniatus, and one complete C. textile were present, as well as two C. textile with a hole in the apex and a Conus cf. miliaris, with an incision in its body. Among the other unidentifiable Conus sp., four exhibit the spire (top) of the shell that seems naturally abraded and three are shells with a hole in the apex. In addition, one Conus apex bead was found, ca. 45 mm in outer diameter (Bar-Yosef Mayer, 2003, 2011). At Nebi

A unique case is that of the third millennium site of Ra’s al-Junayz in Oman. There, a workshop for the production of Conus apex beads was discovered. The raw material there was identified as consisting of C. flavidus, C. terebra thomasi, C. pennaceus, C. textile and C. virgo. The presence of shells in various stages of production enabled the reconstruction of the chaîne opératoire, and tools used in this operation were also found. These tools included anvils (called supports), blades and borers, as well as abraders and polishers

Conus apex beads in other Levantine sites There are several known cases of Conus apex beads in the southern Levant. They were found in tombs dated to EBII from Jericho, Tombs A127 and F5 (Kenyon 1960:92, 172-173; Figs. 28, 65). Biggs (1963:127) mentions specifically a holed example of C. taeniatus from the Early Bronze Age. Unidentified as to species and not described in detail, Conus are mentioned in a number of tombs at Bab-edh-Dhra. A few photos suggest that at least some of them were Conus apex beads (Wilkinson 1989:461-470, figures 266, 269). All of these burials are dated to EBII/III in Jericho and Bab-edh-Dhra (Kenyon 1960, 1965; Schaub and Rast 1989).

212

D. E. Bar-Yosef Mayer et al.: Conus ornaments from Tel Bareqet in an Early Bronze Age (Charpentier 1994). Other sites in Oman seem to display similar workshop attributes (Azzarà 2009), yet in all the cited references, these objects never reached the Indus Valley, which is known to have had contacts with Oman during the third millennium BCE (Charpentier 1994:165). Thus, we may conclude that the Conus apex bead is indeed a Near Eastern artifact, which probably had a specific meaning understood by the inhabitants of the region that spans the Levant, Mesopotamia, and part of Arabia.

be too light to serve as spindle whorls (although this is not impossible). Ethnographic studies may provide some clues that help interpret archaeological finds, although we may never be certain how accurate these analogies are. A case in point is the use of Conus apex beads in East Africa. Hingston Quiggin mentions their use of as ornament or currency (Hingston Quiggin 1949: 50, 103). In the early 1960’s, based on her observations and historic accounts, J. R. Harding, described Conus shell disc ornaments in Tanganyika (today Tanzania) and other African countries that have been in use for hundreds of years, (Harding 1961, 1962). She mainly described in detail how the shells were manufactured and worn among various African tribes. She witnessed men drilling holes in the Conus spires using drills consisting of long steel pins set into wooden handles (Harding 1961:60). Harding identified the same artifact bearing different names. Sometimes the name changed within the same tribe, depending on the way it was used. In Tanganyika, it was called Kibangwa (plural: Vibangwa). In Rhodesia (today Zimbabwe), this object was called Kilungu (plural: Shilungu). There, sometimes half a Conus Museum 60 ring was used and made into a triangular shape, called lupingu. The Conus rings were usually made of Conus betulinus, C. litteratus, or C. virgo (all Indo-Pacific species). Harding noted that the artifacts were apparently manufactured on the Indian Ocean coast and traded inland, into the tribes in which she found them.

Other spires of Conus shells in Bahrain were made into shell seals, using other types of perforation and decoration on the Conus spires after they were cut and filed (al Khalifa 1986). To date no such artifacts were discovered in the Levant. Discussion and conclusions During the third millennium BCE, alongside various small beads made of different types of minerals, metal and shell (Wygnańska and Bar-Yosef Mayer 2018), Conus apex beads made of large Indo-Pacific shells stand out as unique artifacts with a very wide geographic distribution, often associated with graves. The excavations at Tel Bareqet stand out in that for the first time in the archaeological research of the Land of Israel, a central coastal plain fortified urban settlement that was well-preserved, was widely exposed. Fortunately, the rich material culture revealed at the site, included many metal tools and weapons, complete vessels and prestige items. Conus apex beads found at Bareqet, and particularly the use-wear traces found on them, demonstrate that they were used by the living populations. When the beads were inserted into graves, they were probably intended to reflect the status of the interred.

According to various local traditions, these are ancient artifacts, originally manufactured on the East African coast by Muslim traders. In some cases there were historic documents linking them to Portuguese traders of the 16th century. The further away from the coast they moved the harder it would be for them to obtain the shell out of which the ornaments were made, and consequently, it gained its high value (Harding 1961:56). The Conus disc ornament was usually worn in East Africa as a body decoration, using a piece of lion hide or that of some other carnivore. It was placed on the forehead, the chest, the arms, the wrists or the ankles. In various tribes, it was used to decorate clothes or sandals, as well as the loincloths of dancers. In some tribes, it was tied with fiber from a Mhoja tree or placed on an iron plate. In one case, Harding noted that the chief was buried with one or more beads on his forehead. Among some tribes, the vibangwa served in certain rituals, especially those that have to do with healing. In other tribes, they had monetary value. Often, they were used as tribal insignia while in some tribes they were used to mark the enemy. In still other groups, the vibangwa were used to identify important personalities. Harding cites David Livingstone’s accounts, which equated the great value of the vibangwa with the Lord Mayor’s badge in

Conus apex beads with a similar shape are also found in Late Bronze (LB) and Early Iron Age sites. But those sometimes differ from the EB shells in the extra step of manufacturing around the edges, when they were filed to produce a square or rectangular shape. All stages of production seem to be present in the LB and Iron Age levels of Tel Bet Shemesh (Grant et al. 1931: Pl. 20; Brandl 1985; Tamar and Mienis 2008). It is likely that the LB and Iron Age specimens are more closely related, or influenced by, similar items from the Aegean world (Reese 1983), but this topic requires further research. When considering the possible use of these beads, the first option is that they were simple beads used for stringing on a necklace. However, their relatively large holes are atypical in comparison with other shell beads. Moreover, they are considerably larger than most other types of beads from the EBII in the Levant (Wygnańska and Bar-Yosef Mayer 2018). They seem to 213

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead London (Harding 1961:54). Harding notes (1962:170) that in both East and West Africa only people of distinction possessed this type of Conus shell disc.

Daily Life, Materiality, and Complexity in Early Urban Communities of the Southern Levant: Papers in Honour of Walter E. Rast and R. Thomas Schaub: 185-195. Winona Lake, Indiana, Eisenbrauns. Bar-Yosef Mayer, D. E. 2015. Nassarius Shells: Preferred Beads of the Palaeolithic. Quaternary International 390: 79-84. Bar-Yosef, O., Belfer, A., Goren, A., Smith, P. 1977. The Nawamis Near ‘Ein Huderah (Eastern Sinai). Israel Exploration Journal 27 (2-3): 65-88. Bar-Yosef, O., Belfer-Cohen, A., Goren, A., Hershkovitz, O., Ilan, O., Mienis, H. K. and Sass, B. 1986. Nawamis and Habitation Sites Near Gebel Gunna, Southern Sinai. Israel Exploration Journal 36 (3-4): 121-167. Beit-Arieh, I. 2003. Archaeology of Sinai: The Ophir Expedition. Archaeology of Sinai: The Ophir Expedition. Tel Aviv: Emery and Claire Yass Publications in Archaeology Institute of Archaeology, Tel Aviv University. Biggs, H. E. J. 1963. On Mollusca Collected during the Excavations at Jericho, 1952-1958, and their Archaeological Significance. Man August 1963: 125128. Bosch, M. D., Wesselingh, F., Mannino, M. A. 2015. The Ksâr ‘Akil (Lebanon) Mollusc Assemblage: Zooarchaeological and Taphonomic Investigations on Upper Palaeolithic Shells. Quaternary International 390: 85-101. Bosch, D., Dance, S. P., Moolenbeek, R. G., Oliver, P. G. 1995. Seashells of Eastern Arabia. London, Motivate Publishing. Brandl, B. 1985. Scarab, Scaraboid and Beads made of Shell from the Persian Period at Yafit. Erets Israel 18: 290-292. Charpentier, V. 1994. A Specialised Production at Regional Scale in Bronze Age Arabia: Shell Rings from Ra’s Al-Junayz Area (Sultanate of Oman). In K. Koskemienni, and A. Parpola (eds), South Asian Archaeology, Helsinki 1993: 157-170. Helsinki, Suomalainen Tiedakatemian Toimituksia. Dekker, H. and Orlin, Z. 2000. Check-List of Red Sea Mollusca. Spirula 47: 1-46. Grant, E., Wood, I. F. and Wright, G. E. 1931. Ain Shems Excavations (Palestine). Biblical and Kindred Studies, Vol. 3. Haverford, Haverford College. Harding, J. R. 1961. Conus Shell Disc Ornaments (Vibangwa) in Africa. Journal of the Royal Anthropological Institute of Great Britain and Ireland 91: 52-66. Harding, J. R. 1962. Conus-Shell Disc Ornaments in Africa. Man November 1962: 170. Heiman, E. L. 2002. Shells of East Sinai, an Illustrated List: Conidae. Triton 5: 19-23. Hingstonm Quiggin, A. 1949. A Survey of Primitive Money. London, Methuen and Co. Kenyon, K. M. 1960. Excavations of Jericho. Vol. I. London, British School of Archaeology in Jerusalem. Kenyon, K. M. 1965. Excavations of Jericho. Vol. II. London, British School of Archaeology in Jerusalem.

Further support for this latter notion comes from Safer and Gill. They point to Conus spires used to decorate and/or identify a high-status woman in Angola, and that of a male chief in Zambia (Safer and Gill 1982:9899). In the same volume, other occurrences of large Conus disks are mentioned as decorating noblemen in Taiwan (p. 106) and other parts of Southeast Asia and the Pacific. In the Samoan Islands, funerary sticks decorated with Conus rings were kept with the skulls of dead ancestors in small funerary houses. The shells represented the spirits of dead chiefs, and they were used by living chiefs to communicate with their dead ancestors (Safer and Gill 1982:173). As stated above, there is no way to connect directly between these ethnographic accounts and the archaeological finds from the third millennium BCE. However, the presence of Conus apex beads mostly in graves, their relatively rare occurrence, and their outstanding size and impressive appearance lend support to their probable use as prestige items. Acknowledgements This paper is dedicated with appreciation to Isaac (Itzik) Gilead. We are grateful to the editors of this volume for inviting me to participate. We thank Oz Rittner for photography of the artifacts. References Al Khalifa, H. 1986. The Shell Seals of Bahrain. In Shaikha Haya Ali Al Khalifa, and M. Rice (eds), Bahrain through the Ages: The Archaeology: 251-261. London, Routledge. Azzarà, V. M. 2009. Domestic Architecture at the Early Bronze Age Sites HD-6 and RJ-2 (Ja’Alān, Sultanate of Oman). Proceedings of the Seminar for Arabian Studies 39: 1-16. Bar-Yosef Mayer, D. E. 1997. Neolithic Shell Bead Production in Sinai. Journal of Archaeological Science 24: 97-111. Bar-Yosef Mayer, D. E. 1999. The Role of Shells in the Reconstruction of Socio-Economic Aspects of Neolithic through Early Bronze Age Societies in Southern Sinai. Unpublished PhD dissertation, The Hebrew University of Jerusalem. Bar-Yosef Mayer, D. E. 2003. Mollusc Shells and Shell Beads. In I. Beit-Arieh, (ed.), Archaeology of Sinai: The Ophir Expedition: 229-241. Tel Aviv: Emery and Claire Yass Publications in Archaeology Institute of Archaeology, Tel Aviv University. Bar-Yosef Mayer, D. E. 2011. Nawamis, Shells, and Early Bronze Age Pastoralism. In M. S. Chesson, (ed.), 214

D. E. Bar-Yosef Mayer et al.: Conus ornaments from Tel Bareqet in an Early Bronze Age Moorey, P. R. S. 1994. Ancient Mesopotamian Materials and Industries: The Archaeological Evidence. Oxford, Clarendon Press. Oguchi, K. 1992. Shells and Shell Objects from Area A of ‘Usiyeh. Al-Rafidan XIII: 61-81. Pauc, P., Reinhard, J. 2002. Protohistoric Shell Bead Manufacture and the Problem of String Suspension: Recent Studies in the Northwest Mediterranean Region. The Malaco+Archaeology Group Newsletter 2: 2-5. Paz, Y. 2010. Community Archaeology in ProtoHistorical Tel Bareqet, Israel: School Children and Agency for Active Public Engagement in Cultural Heritage Projects. Public Archaeology 9 (1): 34-47. Paz, Y. and Paz, S. 2007. Tel Bareket – Excavations in a Fortified City of the Early Bronze Age II in the Central Coastal Plain. Qadmoniot 40 (134): 8288. Puillandre, N., Duda, T. F., Meyer, C., Olivera, B. M., Bouchet, P. 2015. One, Four Or 100 Genera? A New Classification of the Cone Snails. Journal of Molluscan Studies 81: 1-23. Reese, D. S. 1983. The use of Cone Shells in Neolithic and Bronze Age Greece. The Annual of the British School at Athens 78: 353-357. Reese, D. S. 1991. Shells from Umm an-Nar and Hafit Graves, Abu Dhabi. In K. Frifelt (ed.), The Island of Umm an-Nar, Volume 1: Third Millennium Graves: 184186. Moesgaard, Aarhus University Press. Regev, J., Miroschedji, P. de, Greenberg, R., Braun, E., Greenhut, Z., Boaretto, E. 2012. Chronology of the Early Bronze Age in the Southern Levant: New Analysis for a High Chronology. Radiocarbon 54: 52566.

Safer, J. F. and Gill, F. M. 1982. Spirals from the Sea: An Anthropological Look at Shells. New York, Clarkson N. Potter. Schaub, R. T. and Rast, W. E. 1989. Bab Edh-Dhra: Excavations in the Cemetery Directed by Paul W. Lapp (1965-67). Winona Lake, Indiana, Eisenbrauns for the American Schools of Oriental Research. Stiner, M. C., Kuhn, S. L. and Gülec, E. 2013. Early Upper Paleolithic Shell Beads at Üçagızlı Cave I (Turkey): Technology and the Socioeconomic Context of Ornament Life-Histories. Journal of Human Evolution 64: 380-398. Tamar, K. and Mienis, H. K. 2008. Conus Arenatus Aequipunctatus at Tel Bet Shemesh, Israel: From Cone ‘crowns’ to Holed Cone Rectangles. The Archaeo+Malacology Group Newseltter 14: 2-5. Vanhaeren, M., d’Errico, F. 2006. Aurignacian EthnoLinguistic Geography of Europe Revealed by Personal Ornaments. Journal of Archaeological Science 33: 1105-1128. Villa, P., Soriano, S., Tsanova, T., Deganof, I., Higham, T. F. G., d’Errico, F., Backwell, L., Lucejkof, J. J., Colombini, M. P., Beaumont, P. B. 2012. Border Cave and the Beginning of the Later Stone Age in South Africa. Proceedings of the National Academy of Science of the U.S.A. 109: 13208-13213. Wilkinson, A. 1989b. Objects from the Early Bronze II and III Tombs. In R. T. Schaub and W. E. Rast, (eds), Bab Edh-Dhra: Excavations in the Cemetery Directed by Paul W. Lapp (1965-67): 444-470. Winona Lake, Indiana, Eisenbrauns, for the American Schools of Oriental Research. Wygnańska, Z. and Bar-Yosef Mayer, D. E. 2018. Beads. In M. Lebeau (ed.), ARCANE Interregional II. Artifacts: 283-294. Turnhour: Brepols.

215

Flint Knapping in a Brush Hut: A Case Study from Ohalo II, a 23000 Year-old Camp in the Sea of Galilee Dani Nadel, Daniel Kaufman, Udi Grinburg and Dan Malkinson Introduction

camp. These include large quantities of charred seeds and fruit (Kislev et al. 1992; Weiss et al. 2008). There are also many thousands of animal bones, including bones of mammals (Rabinovich 2002), micro-mammals (Belmaker et al. 2001), birds (Simmons and Nadel 1998), and fish (Zohar 2002). Also, there are hundreds of shell beads (Bar-Yosef Mayer 2002), bone implements (Rabinovich and Nadel 1994-5), and basalt/limestone artifacts. The latter include an in situ grinding stone used for cereal processing (Nadel et al. 2012; Piperno et al. 2004). However, the most common finds are flint artifacts, totalling more than 100,000 pieces (Nadel 2001; Nadel and Sarel 2005).

Spatial analysis of material remains has a long history in the archaeological literature. Mostly, it was developed from ethnographic and ethno-archaeological works, in which human activities were consistently documented to be spatially organized (Binford 1983; Leroi-Gourhan and Brezillon 1972). Commonly, the hearth was the focal place for a wide array of mundane and ritual activities in a wide range of sites and periods (e.g. Binford 1998; Brooks and Yellen 1987; Galanidou 1997, 2000; Goring-Morris 1988; Hayden and Cannon 1983; O’Connell 1987; Stapert 1989, 1990; Yellen 1977).

A wide range of remains, field observations, and micromorphological analyses indicate in situ preservation not only of brush hut floors and hearths, but also of large and small remains (Nadel and Werker 1999; Nadel et al. 1994, 2004a,b, 2006; Tsatskin and Nadel 2003). These observations and results prompted the spatial analyses

Accordingly, many spatial analyses of prehistoric sites focused on the distribution patterns of artifacts around hearths (e.g. Alperson-Afil and Goren-Inbar 2010; Brooks and Yellen 1987; Stapert 1989), especially, as in many cases, no other in situ features were preserved. The Ohalo II site, with its high quality of preservation, provides an excellent opportunity to study the distribution patterns of remains within a prehistoric camp in general and in locales such as brush hut floors, in particular. The goals of this paper are to describe the Brush Hut 2 flint assemblage and its spatial patterning. This, in turn, provides a case study for better understanding flint-related activities within a brush hut, especially in terms of actual knapping locales. The Ohalo II site The site is located on the shore of the Sea of Galilee, and it was excavated during 1989-1991 and 1999-2001. It was submerged for millennia, and it was exposed only due to severe droughts and heavy pumping of water out of the lake. The remains include six brush huts, of which four were fully excavated (Loci 1, 2, 3 and 13) and two widely excavated (Loci 12 and 15) (Figure 1). In Locus 1 the remains include the oldest known wall bases (Nadel 2003a; Nadel and Werker 1999) and grass bedding (Nadel et al. 2004a). A wealth of finds was preserved on the brush hut floors, around the nearby hearths, and in other parts of the

Figure 1. a) Location map.

216

D. Nadel et al.: Flint Knapping in a Brush Hut

Figure 1. b) plan of the Ohalo II site.

217

Studies in Archaeology and Ancient Cultures in Honor of Isaac Gilead of material remains, here with a focus on the flints on one brush hut floor. The site is radio-carbon dated by more than 45 readings from various loci as well as preand post-depositional layers to c. 23,000 cal BP (Nadel 2002; Nadel et al. 1995).

In certain ways the Ohalo II assemblage is typical of southern Levantine Late Upper Paleolithic/Early Epipaleolithic industries, while in others it differs (e.g. Bar-Yosef 1970, 1991; Goring-Morris 1987, 1995; Kaufman, 1987, 1988, 2003; Nadel 2003). This aspect is beyond the scope of the current paper.

The Ohalo II flint assemblage

The Brush Hut 2 assemblage (N=5,621) derives from one of the smallest brush huts at the site, 7.5 m2 in area and with an excavated volume of 0.75 m3. The brush hut was fully excavated, and the entire flint assemblage was studied. Furthermore, this is one of the two assemblages subjected to a refitting endeavour (Grinburg 2005). In most parameters the assemblage falls within the average of the six brush huts (Table 1).

The flint assemblage from six brush huts is large. It includes 45,429 pieces (excluding debris; Table 1). Noteworthy, all identifiable debitage specimens larger than 10 mm are included in these counts. The flints are mostly very sharp, although a variety of patination colors are present. The main target products were bladelets and small blades, ranging between 48.2% (Locus 15) and 61.6% (Locus 12) of the assemblages. The cores are mostly small, made of local pebbles. Their total number in the brush huts (N=204) could not have been the source for more than 45,000 debitage pieces and tools. The ratio of products/cores (227:1) is too high for the size of the cores, and there is no way that more than 200 specimens were knapped from each core. Even the ratio of blade/lets to cores (124:1) is much too high. These data indicate either that many cores were discarded outside the huts or taken elsewhere, or that many of the products were manufactured elsewhere and brought into the huts (or a combination of both).

Methodology The characterization of the spatial distribution of flints on the floor of Brush Hut 2 includes three major aspects. The first is the mapping of selected debitage components. The second tests with statistical tools the spatial correlations between these components. The third focuses on the mapping of the refitted items. Mapping

All stages of the reduction sequence are present in each locus, including cores, primary elements, core trimming elements, and tools. A large proportion of the blade/lets are twisted, and the ratio between right and left twist is 9:1.

A common method for mapping archaeological finds is the spline interpolation approach, fitting a smooth surface to the counts in archaeological grid cells (e.g. Goren-Inbar et al. 2004; Metcalf and Heath 1990; Simms and Heath 1990; Weiss et al. 2008). Addressing the advantages and disadvantages of this procedure is beyond the scope of this paper. However, it suffices to say here that the main problem is that the remains were collected in well-delineated square units, while the results are presented in contours crossing the original grid. Thus, the plotted results may not be accurate, and they may contain artificial features (Enloe et al. 1994; Simek and Larick 1984).

Most of the tools are fashioned on bladelets. Adhesive remains and impact fractures were identified on several retouched bladelets, including twisted bladelets. These were interpreted as barbs incorporated in composite projectiles (Yaroshevich et al. 2013). It is noteworthy that the scalene/proto triangle type was found exclusively in two brush huts (1 and 15), with none in Locus 2 (Nadel 1999).

Locus

Tools

Cores

Blade/Lets

Flakes

Primary elements

Cte

Total

N

L.1

7.1%

0.4%

55.3%

21.5%

12.7%

3.0%

100.0%

8,955

L.2

2.8%

0.4%

53.1%

29.4%

10.3%

4.0%

100.0%

5,621 12,419

L.3

2.8%

0.2%

56.4%

27.3%

9.6%

3.7%

100.0%

L.12

1.6%

0.3%

61.6%

24.7%

9.4%

2.4%

100.0%

4,393

L.13

2.5%

0.7%

55.5%

27.3%

10.5%

3.5%

100.0%

13,093

L.15

10.0%

0.3%

42.8%

28.6%

16.2%

2.0%

100.0%

948

Total

3.6%

0.4%

55.7%

26.2%

10.7%

3.4%

100.0%

45,429

average

4.5%

0.4%

54.1%

26.5%

11.5%

3.1%

7,572

std. dev.

3.3%

0.2%

6.2%

2.9%

2.6%

0.8%

4,768

Table 1. The flint assemblages from six brush huts.

218

D. Nadel et al.: Flint Knapping in a Brush Hut A different approach was adopted here, keeping the presentation and all analyses to the original grid squares. Such an approach is commonplace for Levantine sites, especially Natufian sites (e.g. Edwards and Hardy-Smith 2013; Samuelian 2013a,b; Valla et al. 2013). For Ohalo II, we calculated the density of artifacts per excavation unit (usually 0.5 x 0.5 x 0.05 m; however, some units are smaller – especially by the edges of the brush hut floor where the units are not square), and then the mean for the studied feature (e.g. the floor of Brush Hut 2). On the maps, each unit represents the density of finds as expressed in standard deviation (std. dev.) units, compared to the mean of the floor. The units were then grouped into intervals of 1 std. dev., according to their divergence from the mean.

In order to represent best the spatial patterning of the flint assemblage, maps of most debitage components were produced. These include the total assemblage, the most common products (blade/lets) with cores, primary elements with cores, and core trimming elements with cores. In addition, the tools (macroliths and microliths combined) and the microliths were also mapped. Finally, all minute remains (2.5 std. dev.). The core trimming elements differ from the above (N=182, Figure 3c). They also have two peaks although these are somewhat more diffused. The one in the north is offset relative to the peaks of other debitage components. However, in terms of Pearson’s r, all debitage components have a high positive correlation among themselves (Table 3). Types

2

1

2 4

11

479

548

Table 2. The Brush Hut 2 refitted products, by raw material. Note that refitted fragments are counted in the total when relevant (e.g. raw material 10) and not counted when the refitted debitage by itself is composed of fragments (e.g. raw materials 3, 18).

All refitted sets were studied for their technological aspects, e.g., their place within the reduction sequence. The ensuing mapping was conducted according to raw materials, both of refitted debitage (e.g. bladelets, primary elements) and refitted fragments (e.g. two pieces of a broken bladelet). Results

r

p(tr)

cores – total

0.84