The Shephelah during the Iron Age: Recent Archaeological Studies 9781575064871

The area of the Judean Foothills – the biblicalShephelah – has in recent years become one of the most intensively excava

175 34 10MB

English Pages 218 Year 2017

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

The Shephelah during the Iron Age: Recent Archaeological Studies
 9781575064871

  • 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

The Shephelah during the Iron Age

This page left intentionally blank.

The Shephelah during the Iron Age Recent Archaeological Studies

“. . . as plentiful as sycamore-fig trees in the Shephelah” (1 Kings 10:27, 2 Chronicles 1:15)

edited by

Oded Lipschits and Aren M. Maeir

Winona Lake, Indiana Eisenbrauns 2017

© 2017 by Eisenbrauns Inc. All rights reserved Printed in the United States of America www.eisenbrauns.com

Library of Congress Cataloging-in-Publication Data Names: Maeir, Aren M., editor. | Lipschitz, Oded, editor. Title: The Shephelah during the Iron Age : recent archaeological studies / edited by Aren M. Maeir and Oded Lipschits. Description: Winona Lake, Indiana : Eisenbrauns, [2017] | Includes bibliographical references. Identifiers: LCCN 2016059410 (print) | LCCN 2017003148 (ebook) | ISBN 9781575064864 (hardback : alk. paper) | ISBN 9781575064871 (ePDF) Subjects: LCSH: Shephelah (Israel)—Antiquities. | Excavations (Archaeology)— Israel—Shephelah. | Iron age—Israel—Shephelah—Antiquities. Classification: LCC DS110.S555 S55 2017 (print) | LCC DS110.S555 (ebook) | DDC 933/.47—dc23 LC record available at https://lccn.loc.gov/2016059410

The paper used in this publication meets the minimum requirements of the American National Standard for Information Sciences—Permanence of Paper for Printed Library Materials, ANSI Z39.48–1984. ♾ ™

Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Aren M. Maeir and Oded Lipschits

vii

Four Seasons of Excavations at Tel Azekah: The Expected and (Especially) Unexpected Results . . . . . . . . . . . . 1 Oded Lipschits, Yuval Gadot, and Manfred Oeming Swinging on the “Sorek Seesaw”: Tel Beth-Shemesh and the Sorek Valley in the Iron Age . . . . . . . . Shlomo Bunimovitz and Zvi Lederman

27

Tel Burna: A Judahite Fortified Town in the Shephelah . . . . . . . . . . . . . 45 I tzhaq Shai Tel Gezer Excavations 2006–2015: The Transformation of a Border City . . . . 61 Steven M. Ortiz and Samuel R. Wolff Tell Halif in the Late Bronze and Iron Age . . . . . . . . . . . . . . . . . . . 103 Oded Borowski The Iron Age City of Khirbet Qeiyafa . . . . . . . . . . . . . . . . . . . . . . 115 Yosef Garfinkel Philistine Gath after 20 Years: Regional perspectives on the Iron Age at Tell eṣ-Ṣafi/Gath . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 Aren M. Maeir The Archaeology and History of Tel Zayit: A Record of Liminal Life . . . . . . 155 Ron E. Tappy Settlements and Interactions in the Shephelah during the Late Second through Early First Millennia BCE . . . . . . . . . . 181 I do Koch

v

This page left intentionally blank.

Introduction Aren M. Maeir and Oded Lipschits

The Judean Foothills—the biblical Shephelah (fig. 1)—has in recent years become one of the most intensively excavated regions in the world. The large number of archaeological projects, of various scopes and aims, that are (or in some cases, up until very recently) working in this area is a quite unique situation—so much so that the biblical quotation appearing in the book’s subtitle “. . . as plentiful as sycamore trees in the Shephelah” (1 Kgs 10:27; 2 Chr 1:15) could not have been more appropriate. Due to the large number of recent and ongoing excavations in this region, it appears that the study of the region and its ancient cultures can move beyond the standard baseline of much archaeological research—the basic chrono-stratigraphic and typological frameworks—and deal with broader, and ultimately more interesting and more significant issues. No less important is the recurring role of the Shephelah, throughout many historical periods, as a transitional zone between regions, cultures, and polities. This is especially clear during the Late Bronze and the Iron ages: time and again, the region was the transition zone between cultures and polities in the Coastal Plain (Egyptian in the Late Bronze; Philistine in the Iron Age) and the Central Hills (Canaanite and/or rural in the Late Bronze; Israelite/Judahite/Canaanite in the Iron Age). This transitional character has come into focus in many of the archaeological excavations conducted in the region. For these reasons, as part of the 16th World Congress of Jewish Studies, which was held in Jerusalem in July-August, 2013, the editors of the present volume (one of whom at the time had been excavating at Tell eṣ-Ṣafi/Gath for close to two decades [Maeir] and the other [Lipschits] had just commenced an excavation at Tel Azekah) initiated a double session on current research on the Late Bronze and Iron Ages of the Shephelah. In this double session, papers on seven different excavations (Beth Shemesh, Gezer, Burna, Tell eṣ-Ṣafi/Gath, Azekah, Tel Zayit and Khirbet Qeiyafa) were presented. Following and between these papers, interesting discussions developed, stressing the vibrant state of research in the region. Following the success of the session, with the consent of the lecturers who had presented papers, we decided to publish these papers in a proceedings volume and to add to the original papers several invited papers as well: one of them on another Shephelah site (Tel Halif, by O. Borowski) and one overview paper (by I. Koch). Almost all the major current excavations in the region are represented (save for Tel vii

viii

Aren M. Maeir and Oded Lipschits

Fig. 1.  Map of the Shephelah, showing major sites.

Eton, Tel Lachish, and Khirbet Arai; the latter two commenced after their excavators submitted their paper on Khirbet Qeiyafa). And these nine contributions are offered in the present volume. As can be seen from reading the various papers, various opinions, not always mutually agreed, are presented. This we see as one of the advantages of this volume—to present new data, to offer new interpretations of these data, but to demonstrate the lively, and at times even boisterous, scholarly discussions currently taking place on various issues relating to the archaeology and history of the Late Bronze and Iron Age Shephelah and its immediate environs. The first eight chapters in the volume are discussions of the excavations and discoveries from specific sites, and they are arranged alphabetically, according to site names. The final chapter, by I. Koch, is a general overview of the Shephelah during the Late Bronze and Iron Age. In the first chapter, Lipschits, Gadot, and Oeming, present an overview of the first seasons of the new excavations at Tel Azekah, a site that had previously been excavated by Bliss and Macalister in 1898–1899. In their article, they review previous research and then compare their initial results—those expected and those not.

Introduction

ix

Noteworthy is the fact that, despite earlier excavations, which were poorly conducted and poorly published, and the question of the overall preservation of the site, the finds from a wide range of periods were quite impressive and promising in regard to the long-term excavations of this important site. In the second essay, Shlomo Bunimovitz and Zvi Lederman, who have directed the excavations at Tel Beth Shemesh for about 25 years, discuss what they refer to as the “Soreq Seesaw”—the fluxes and changes in settlement patterns and cultural influences during the Iron Age. Accordingly, they identify four major changes in this “seesaw,” following the initial setting of the swing during the Iron Age I, between the Philistines and the Canaanites in Iron Age I. The four swings that they suggest are: (1) A westward push of early Judahite State toward Philistia in the early Iron IIA; (2) A resurgence of Philistia and a push eastward from the late 10th through the early 8th century BCE; (3) Hezekiah’s push westward in the late 8th century BCE; and finally, (4) the campaign of Sennacherib of Assyria and its aftermath in the late 8th and early 7th centuries BCE. The relatively recently commenced excavations at Tel Burna are described by Itzik Shai in the third chapter. In particular, he focuses on the fortified Iron Age phases at the site (which he identifies as Libnah, in contrast to Tappy [see chapter 8 in this volume], who suggests that Libnah be located at Tel Zayit), during the Iron IIA, Iron IIB, and Iron IIC. Based on various material finds, he argues, quite convincingly, that the site should be identified as a Judahite site, situated on the border with Philistine-controlled regions to the west. The fourth article, by Steve Ortiz and Sam Wolff, directors of the renewed expedition to Gezer, present an overview of the main discoveries of their expedition from the Middle Bronze, Late Bronze, and primarily from the Iron Age I and II. Importantly, they attempt to tie the results of their work with the results of the earlier excavations at Gezer, from Macalister in the early 20th century CE through other, more recent excavations. Oded Borowski, current director of the excavations at Tel Halif, provides a general overview of the Late Bronze and Iron Age finds at the site. Among other points, he notes that (1) during the Late Bronze Age the site served as a collection center for agricultural produce; that (2) there is no destruction between the Late Bronze and (rather limited) Iron I levels; and that (3) during the Iron Age II (9th through 7th centuries BCE) the site served once again as a center for the collection of agricultural produce. The sixth chapter, by Yosef Garfinkel, is an overview of the major finds from his excavations at Khirbet Qeiyafa, which were just ending at the time of his lecture in 2013. In his paper, which concentrates on the Iron Age remains, he reviews what he suggests are the urban aspects of the site, trade connections, and administrative aspects, and delves briefly into the chronology of the Iron Age levels at the site. Needless to say, Garfinkel debates other interpretations of these important remains. Aren Maeir has directed the excavations at Tell eṣ-Ṣafi/Gath since 1996, and many publications have resulted from his work there. So, instead of describing the

x

Aren M. Maeir and Oded Lipschits

finds from the site once again, he chooses to discuss several general themes relating to the finds. These include: (1) the development of Philistine culture in light of recent finds at Tell eṣ-Ṣafi/Gath and other sites; (2) a critical engagement with the suggestion that there was a “Canaanite Enclave,” which supposedly existed in the Shephelah during the Iron I; (3) discussion of the status of Gath of the Philistines during Iron Age IIA (10th–9th centuries BCE), arguing that Gath’s lofty status during this phase had a substantial impact on the geopolitical situation in the Shephelah and a strong influence on the early Judahite Kingdom; (4) a review of the cultural influences between the kingdoms of Gath and Judah during the Iron Age IIA; and finally (5) an argument that the siege system around Tell eṣ-Ṣafi/Gath should be related to the siege and conquest of the site by Hazael of Aram Damascus and is properly identified as such. Ron Tappy, director of the recently completed excavations at Tel Zayit, provides a perspective on the finds from this site through the lens of a “liminal” perspective. First, he presents an extensive theoretical discussion of the importance of liminal sites and zones in historical, geographical, and archaeological research. Then, he focuses on three specific periods in the history of Tel Zayit: (1) Iron IIA, (2)  the late 8th century BCE (in both of these periods, the site, which he identifies as biblical Libnah, was on the Judahite side of the border between Philistia and the Judah, even if during these periods there were strong influences from Philistia), and (3) during the late 3rd and early 4th centuries CE, when the site served as a Roman fort. In these (and other) periods, one can see how the site, and this region, served as a border zone (a “liminal” zone) between different cultures, regions, and influences and how longue durée patterns can be seen at the site. The final chapter, by Ido Koch, is based by and large on his doctoral dissertation. Koch offers a synthetic view of the stages in the development of the settlement pattern and political structure of the Shephelah, from the late Late Bronze Age until the early Iron IIA. He suggests that, following the collapse of the Late Bronze political structure dominated by the Egyptian empire, the main polity in the ensuing Iron I is to be identified as Tel Miqne. At the same time, many parts of the Shephelah had limited settlement (save for a few central sites) but, nevertheless, the sites that existed had close cultural and trade connections. In the Iron IIA, following the repeated destruction of this Iron I settlement pattern, Tell eṣ-Ṣafi/ Gath became the major polity in the Shephelah, up until its destruction by Hazael in the late 9th century BCE. All together, the nine chapters in this volume provide both succinct reviews of the Late Bronze Age and Iron Age archaeological remains at some of the major sites in the Shephelah and at the same time discuss and debate various issues and topics. As can be seen from the various chapters, much is debated and not all is agreed upon, but the dynamic and lively discussions are evidence of the cutting edge archaeological research that is currently being conducted in this region. We can but hope for many more years of intensive research and important results in this important region.

Four Seasons of Excavations at Tel Azekah: The Expected and (Especially) Unexpected Results Oded Lipschits, Yuval Gadot, and Manfred Oeming

Geopolitical Locations and Historical Sources―and the Expected Archaeological Results Based on Them Tel Azekah (Tell Zakariya) is a pear-shaped mound encircled on three sides by Nahal Ha-Elah (Wādi ʿAjjur). The mound is ca. 127 m above the stream, atop steep slopes on the west, north, and east. On the south, at a drop of ca. 30 m, it is joined to a ridge by a low saddle. 1 On this saddle, a lower city grew around the southern slopes of the tell during the Late Bronze Age, adding about 13 dunams to the 45-dunam site. Both the size and the strategic location of Azekah in the heart of the Shephelah positions it as one of the main border sites between the coastal entities in the west (e.g., Philistines) and the hill entities in the east (e.g., Judah). It regulated and safeguarded the strategic junction of roads that led from Tell eṣ-Ṣafi (biblical Gath) in the west, through the Valley of Elah, to the Judean Hills in the east, and connected Beth-Shemesh in the north and Lachish in the south. Despite its size and importance, the name “Azekah” is not mentioned in any second-millennium BCE source. Since it was already well-fortified by the Middle Bronze Age (and possibly even before that; see below), and was both substantial and rich in the Late Bronze Age II–III, it might have become known as “Azekah” (and cf. Isaiah 5: 2) only when it became part of Judah—that is, not before the end of the 9th century BCE (and see below). If this is the case, it will remain for future research to determine the name by which the site was known in the second millennium BCE. According to Josh 10:10–11, 15:33–35, Neh 11:30, and 2 Chr 11:9, Azekah was an important stronghold on Judah’s western border. The first book of Samuel (1 Sam 17:1) designates it as the location of the legendary battle between David and 1.  The site can be approached from the south only, and for defensive purposes the saddle was probably artificially lowered in ancient times. Already Dagan (2011: 72–73) assumed that the city gate should be located on the southern slope of the tell and that the Assyrian and Babylonian armies, as indicated in historical documents, also attacked Azekah from this direction.

1

2

Oded Lipschits, Yuval Gadot, and Manfred Oeming

Goliath: The Philistines “gathered their armies for battle; they were gathered at Socoh, which belongs to Judah, and encamped between Socoh and Azekah, in Ephesdammim.” Saul and the Israelite army were positioned against the Philistines at the Valley of Elah, and after the great victory of David over Goliath, the Philistines fled and were pursued by the Israelites “as far as Gath and the gates of Ekron.” With Azekah as a major site on the western border of Judah, it is easy to understand why the Assyrian army chose it as the first target of its 701 BCE attack. The Azekah Inscription, discovered in two fragments in the library of Ashurbanipal, was identified in 1974 as a single tablet by Nadav Naʾaman. 2 It describes an Assyrian campaign by Sennacherib against Hezekiah, King of Judah, including the conquest of Azekah. The text further describes Azekah’s strong fortifications, which were “[like the nest of the eagle?] located on a mountain ridge, like pointed iron daggers without number reaching high to heaven [. . . [Its walls] were strong and rivaled the highest mountains, to the (mere) sight, as if from the sky [appears its head? . . . . (lines 6–7, according to Naʾaman’s 1974 translation). The inscription also mentions the siege and conquest of the city and that a siege ramp was built as part of it: “[by means of beaten (earth) ra]mps, mighty? battering rams brought near, the work of [. . .], with the attack by foot soldiers, [my] wa[rriors . . . [. . .] they had seen [the approach of my cav]alry and they had heard the roar of the mighty troops of the god Ashur and [their] he[arts] became afraid [. . . .” Sennacherib also describes the results of this war and how “[The city Azekah I besieged,] I captured, I carried off its spoil, I destroyed, I devastated, [I burned with fire. . . .]” (lines 8–10, according to Naʾaman’s 1974 translation). A much-diminished Azekah was rebuilt sometime in the late 7th century BCE, probably after a long settlement gap (Koch and Lipschits 2013: 64–66), and by the early 6th century BCE, when Judah was attacked by the Babylonians, it had again become one of the strong, fortified cities on Judah’s western border. According to Jer 34:7, “. . . when the army of the king of Babylon was fighting against Jerusalem and against all the cities of Judah that were left, Lachish and Azekah; for these were the only fortified cities of Judah that remained.” An ostracon discovered in the burned gate of Lachish, dated to the 586 BCE Babylonian destruction, completes the data from the description in Jeremiah, since the last Judahite defenders report in the last lines of their letter (Lachish Ostracon IV; ANET: 321): “and let (my lord) know that we are watching for the fire signals of Lachish, according to all the indications which my lord hath given, for we cannot see Azekah.” Conclusions: Based on the geopolitical location of the site and on the biblical and extrabiblical sources, second-millennium BCE Azekah was barren territory. There would be nothing to find. The main discoveries to be expected would come from the Iron Age—the Iron IIB and IIC. Possibly, based on the mention of Azekah in Neh 11:30 and also 1 Chr 11:9, with textual indications for the existence of 2.  See Naʾaman 1974, as against the attribution of the inscription to the time of Tiglathpileser III or Sargon II, and see there further literature.

Four Seasons of Excavations at Tel Azekah

3

Azekah as a border site within Yehud, there would also be material evidence from the Persian and the Early Hellenistic periods.

Previous Archaeological Excavations at Azekah: The Anticipated Archaeological Results of the Renewed Expedition Tell Zakariya (Azekah) was one of the first sites excavated in the Holy Land. Between October and December, 1898, March and April 1899, and again in September, 1899, F. J. Bliss, assisted by R. A. S. Macalister, excavated the site for 17 weeks on behalf of the Palestine Exploration Fund. Bliss and Macalister published the results of their excavations in four preliminary reports (Bliss 1899a; 1899b; 1899c; 1900). Two years later, they published the final reports of the four excavations they had excavated in the Judean Shephelah: Tell eṣ-Ṣafi (Tel Zafit, identified with Philistine Gath), Tell el-Judeideh (Tel Goded), Tell Ṣandahana (Mareshah), and Tell Zakariya (Bliss and Macalister 1902; on Azekah, see pp. 12–27). A detailed study of the unpublished field diaries and plans from Bliss and Macalister’s excavations (Napchan-Lavon 2014) revealed that, in contrast to the common “archaeological legend,” according to which the fate of Azekah had been similar to that of nearby Mareshah (which, from the archaeological point of view, had been completely destroyed), Bliss and Macalister’s excavations at Azekah were focused on only three areas on the upper part of the tell: three trenches and one pit in the open area of the surface of the tell, the towers at the southwestern edge of the tell, and (mainly), the fortress on the acropolis. Already between October and December 1898, Bliss and Macalister dug 16 test pits along three parallel lines in the open area of the surface of the tell, sectioning the tell from east to west: Section A–B (northernmost), Section C–D (central), and Section E–F (southernmost). In the pits excavated in the northern row, there was a clear distinction, about 2 m below the surface, between the “Jewish” and “Phoenician” periods. Bliss and Macalister observed in their notebook that in the pits in the two other rows a wider variety of pottery types was unearthed, including, according to their analysis, “Jewish,” “pre-Israelite,” “Phoenician,” and “Amorite” fragments (Napchan-Lavon, Gadot, and Lipschits 2014: 86–87). These definitions are different from the official publication, where Bliss and Macalister identified two main strata above bedrock: the lower stratum was identified as “pre-Israelite,” while the upper strata were identified as “Jewish” (Bliss 1899a: 17, pl. 1; Bliss and Macalister 1902: pl. 2). Only one more area was excavated in the upper part of the tell: in the north of the site, a trial pit was excavated, approximately 30 × 20 m, reaching a depth of 3.5 m on average (Napchan-Lavon, Gadot, and Lipschits 2014: 81). Only a few of the pottery vessels, bronze, iron and stone objects, and coins that were recorded in the finds list registered the exact location and depth (and cf. Napchan-Lavon 2014: 129–44); some of them were also published in the preliminary and final reports. Study of these finds, and combining them with the few available details concerning their location and level, may demonstrate that there are clearly three different

4

Oded Lipschits, Yuval Gadot, and Manfred Oeming

levels of occupation discovered in Bliss and Macalister’s excavations on the upper plateau: Early Bronze Age II–III, Late Bronze Age I–II, and Iron Age II. The presence of Early Bronze material close to bedrock can be found in only one location; the appearance of Late Bronze material is clear, as is the Iron Age level on top, found in nearly all the excavated pits, at a depth of between 0.6 and 1.8 m (Napchan-Lavon, Gadot, and Lipschits 2014: 87). The towers at the southwestern edge of the tell were assigned by Bliss and Macalister (1902: 13–14) to the Roman/Byzantine period. From the unpublished material, however, and since no indicative ceramic finds or other materials are mentioned in the diaries as having come from this area, it is unclear on what they based their conclusion (Napchan-Lavon, Gadot, and Lipschits 2014: 89–90). 3 The area most thoroughly excavated by Bliss and Macalister was the fortress on the acropolis. It is therefore the best-documented area among the unpublished material preserved at the PEF and also in the published reports. The fortress’s walls and towers were trenched, disconnecting the walls from occupation levels on both the internal and external sides, making it almost impossible to associate any floors with the fortress walls. However, this method created a clear outline of the building. Four “Clearance Pits” were excavated inside the citadel, exposing many walls and floors and revealing many small finds that were carefully described in the field diaries, in most cases, however, without a clear description of the location of the finds. Based on the interpretation of the evidence from the fortress, Bliss and Macalister (1902: 23) summarize the occupation of the tell: Though not founded in the earliest period, it was already inhabited when Joshua entered the land, and was fortified in Jewish times, possibly by Rehoboam; during the Seleucidan period the acropolis was strengthened by the addition of towers, and finally, after a brief occupation in Roman and Byzantine times, the place was deserted.

For many years, this dating was unanimously accepted by scholars. 4 However, from the descriptions of the masonry in the field diaries and notebooks, it is clear that the walls of the structure were not of homogenous construction. Furthermore, in certain places, the fortress’s walls rest on bedrock, while in others they seem to rest on debris and remains of earlier buildings. The majority of lmlk stamped handles found at Azekah were discovered inside the fortress, as were several lamp-in-bowl and bowl-in-bowl foundation deposits that date to the Late Bronze or early Iron 3.  In his reanalysis of the excavation results, Dagan (2011) suggests, contrary to Bliss and Macalister, an Iron I–II date for these towers. From the section drawings in the published reports, he deduced that the foundations of these towers do not resemble those of the walls of the fortress (which he dates to the Hellenistic period; see below). In the renewed excavations of the northern tower, it became clear that this tower is part of the Middle Bronze fortifications (and see further below). 4.  See, e.g., Avi-Yonah and Yeivin (1955: 289–90); Stern (1971: 133–37; 1993: 124); Aharoni (1987: 266); Seger (1997: 243); Negev and Gibson (2001: 64).

Four Seasons of Excavations at Tel Azekah

5

Age. In a number of places in the walls and towers, drafted and bossed stones appear, typical of the Hellenistic period (Geva 1985: 28, fig. 4; Sharon 1987: 21–42; Dagan 2011: 81–83), and the large rock chambers found within the fortress have been reinterpreted as ritual baths or miqwaot, typical of the end of the Second Temple period (Reich 1990: 281–82; Zissu 2006: 88; Dagan 2011). It seems that these features were also noticed by Bliss and Macalister. From a purely archaeological perspective, they managed to identify four main occupational levels within the fortress, and they also dated the occupation levels at the site accordingly. The earliest level was dated to “the pre-Israelite period,” and among the objects attributed to this level was a vessel containing assorted Egyptian jewelry, including two scarabs, one with the name Thutmose III, the other with the name of Amenhotep II (Bliss and Macalister 1902: 22–23). Based on these finds, we can date this phase to the Late Bronze Age II–III. The second level, corresponding to Bliss and Macalister’s “late pre-Israelite period,” included a crude plaster floor that contained handles with lmlk stamp impressions of the two-winged scarab variety. The level above this contained a second plaster floor of higher quality than the earlier floor, containing jar handles with four-winged lmlk stamp impressions. This level was also dated to “the late pre-Israelite period” (Bliss and Macalister 1902: 22–23). It appears that these two levels should be dated to the Iron II (perhaps parallel to Lachish Levels IV–III and II). The latest level was attributed to “the Jewish period” (Bliss and Macalister 1902: 20–22), and it can safely be dated to the Hellenistic period. Conclusions: After studying the published and unpublished material from the Bliss and Macalister excavations, we may conclude that it is possible to reconstruct the settlement history of Azekah according to their discoveries. Settlement began in the Early Bronze II–III, and the finds from this period can be found very close to the natural bedrock. In the Bliss and Macalister excavations, there is no evidence for the existence of the site in the Intermediate and Middle Bronze Age. On the other hand, there is evidence of a large and significant occupation level during the Late Bronze Age (I–II[?]), as indicated by the many scarabs and Egyptian-style artifacts found in the excavations. Again, there is no evidence of settlement at the site in the early Iron Age; there are, however, many finds and much evidence for at least two different phases during the Iron II. The Hellenistic period is evidenced by the large fortress, but it is not clear from Bliss and Macalister’s excavations if and to what extent the occupation at that time extended beyond this central building.

Archaeological Surveys in Azekah: The Anticipated Archaeological Results of the Renewed Expedition In the late 20th century, Dagan surveyed Tel Azekah as part of the regional surveys of the Shephelah (Dagan 2000: 46–47; 2011, with further literature). His general observations, based on the collection of pottery from the site’s slopes, was that Tel Azekah was settled during the Early Bronze II–III, Intermediate Bronze Age,

6

Oded Lipschits, Yuval Gadot, and Manfred Oeming

Middle Bronze Age, Late Bronze Age, Iron I and II, Persian, Hellenistic, Roman, and Byzantine periods. In 2009, Tel Azekah was subjected to a thorough archaeological survey and ground-penetrating geophysical survey as part of the preparation for the Lautenschläger Azekah Expedition and plans for the renewed archaeological research of the site (Emmanuilov 2012). A geophysical survey, using Electrical Resistivity Tomography (2D/3D), was conducted; it showed the existence of a fortification wall surrounding at least the western side of the surface of the tell and some architectural remains under the surface of the lower southern terrace. 5 The intensive archaeological survey was aimed at identifying more precisely the periods during which the site had been settled and at estimating the size and nature of the site during each of those periods. The surface was divided into nine areas, based on the topography of the site; six of the areas were further divided into survey fields. This division enabled each field to be defined chronologically, independent of the others, and to trace the shifting of the settlements on the surface of the mound in various periods. The results of the intensive survey showed that there were two settlement peaks at the site: in the Late Bronze Age and in the Iron II. These results went hand-in-hand with the results of Bliss and Macalister’s excavations, as described above. In addition, the Early Bronze II–III remains reinforced Bliss and Macalister’s discoveries just above bedrock from the same period, as well as their Late Hellenistic and Early Roman finds. However, the finds from the Middle Bronze IIA, Persian, and Late Roman, Byzantine, Early Islamic, and Ottoman periods (Emmanuilov 2012) had no parallels in Bliss and Macalister’s published and unpublished materials.

Archaeological Expectations Based on Historical Sources and Previous Archaeological Investigations: Interim Conclusions There is good reason to anticipate that careful study of available historical sources, together with a comprehensive assembly and analysis of finds from previous excavations and the results of modern archaeological surveys, will facilitate a detailed reconstruction of the settlement history of Azekah. The results of this preliminary study demonstrate that the earliest phase of settlement at the site was from the Early Bronze II–III. Evidence was found for a limited settlement in the Middle Bronze Age (Emmanuilov 2012: 57, Table 1) and possibly even earlier, during the Intermediate Bronze Age (Dagan 2011: 76–77, Table II). On the other hand, there is evidence of a large and significant occupation during the Late Bronze Age, when the settlement area was expanded to the Lower Terrace in the southern part of the site. The strength of Azekah in the Late Bronze Age, based especially on the presence of Late Bronze sherds in the lower city, came as a surprise, since there are no historical documents supporting the existence of such a large and important site during that period. 5.  The geophysical survey was conducted by Prof. Olaf Bubenzer and Dr. Stefan Hecht on behalf of the Geographical Institute of Heidelberg University.

Four Seasons of Excavations at Tel Azekah

7

Fig. 1.  The location of Tel Azekah.

In the early Iron Age, the site was again home to a relatively small occupation, but it grew significantly in the Iron Age II, when there was a second settlement peak—and this is much more understandable in light of the biblical and extra­ biblical sources. Based on historical reconstruction, we may hypothesize that Judahite settlement at the site began only after the destruction of nearby Tell eṣ-Ṣafi (biblical Gath) in the second half of the 9th century BCE. Azekah developed as a central Judean border city in the late 9th and into the 8th century BCE (similar to Lachish Levels IV–III) and was destroyed in the Assyrian campaign of 701 BCE. It was probably restored in the late 7th century as a fort on the renewed Judean border in the Shephelah (similar to Lachish Level II) and was destroyed again in the Babylonian campaign of 588–586 BCE. It continued to exist as a border town in the Persian period; in the Hellenistic period, it was a large fortress. However, it is not clear if and to what extent the occupation at that time extended beyond this central structure (Emmanuilov 2012: 68, fig. 31). In the Roman period, the settlement continued below the tell itself, and no other buildings were erected on the summit of the tell until the present. It was used as agricultural land and was constantly cultivated, and the marks of the plow can be seen on the upper part of the stones that were excavated just a few centimeters below the topsoil. In the 1970s, the Israel

8

Oded Lipschits, Yuval Gadot, and Manfred Oeming

Defense Forces used the summit of the hill for training; their presence is marked by signs for trenches that were used as field toilets, as well as other earth-works, such as the fill of a large pit on the top of the mound, which may have served as the central water system of the site. Based on this information, one can understand why Keel and Küchler (1982: 826–27) understood Azekah as a place with only some traces of settlement in the Early and the Middle Bronze Ages, but it is hard to understand why they reconstructed what amounted to not much more than a guard-post from the Late Bronze Age. During the Iron Age, they reconstructed a fortified city connected to the administration and army of the Kings of Judah.

The Lautenschläger Azekah Expedition: Plans and Actual Field Work (2012–2015) When they finished excavating, Bliss and Macalister refilled their trenches, leaving no architectural remains visible above the surface of the tell. And since documentation of their excavation trenches is poor, we have no basic knowledge regarding the order of the layers at the site, of the city plan, and/or the extent of settlement in each of the periods. Thus, there were basic questions about the site’s history that had to be dealt with prior to more sophisticated archaeological research. As a result, we chose to plot the first excavation areas so they would deal with these very basic concerns. Three 10-m-wide sections were excavated along the southern (Area S1), eastern (Area E1), and western (Area W1) slopes. Area W2 was added in 2013, 50 m south of Area W1, in order to gain more information on the stratigraphy and fortifications of the western slopes, and Area W3 was opened in 2014 around Bliss and Macalister’s Tower 3 , in order to connect it to the Middle Bronze fortifications. Area E3, located on the southeastern corner of the tell, was added in 2015. The excavations were organized to create a section toward the east, searching for the fortifications of the site and possibly also a siege ramp that was used by Sennacherib in his assault on the city. Area S2 had already been opened in 2012, on the southern lower terrace of the site, in order to study the history of the settlement in this sector, and Areas T1 and T2 were excavated at the top of the mound. These areas were designed to be wide sectors, exposing each layer extensively. Area N was opened at the northern corner of the site, very close to the pit that was excavated by Bliss and Macalister. The excavations in this area were intended to expose the flat, wide area at this strategic location of the site, which had a view of and controlled the main junction of roads below the site, as well as to expose the northeastern corner of the site’s fortifications from the various periods. Many other small trial pits were excavated as part of the training of the M.A. students in our summer educational courses, especially along the line of the presumed fortifications on the eastern and southern slopes of the site. Four seasons of excavations have been conducted at the site thus far (2012– 2015; the results of the fifth season, conducted in the summer of 2016 are not

Four Seasons of Excavations at Tel Azekah

9

Fig. 2.  Tel Azekah: Excavated Areas

included in this paper). Other unexpected results surprised us during these seasons, along with the more expected results, as indicated in the preliminary study of the site.

The Expected Archaeological Results The Early Bronze Period The surveys conducted by Dagan and by our expedition predicted that the site was first settled during the Early Bronze Age. It is therefore no surprise that finds dating to the Early Bronze Age were found all along the western and southern slopes of the tell in Areas W1, W2, W3, S1, and N. Except for Area S1, the finds were always from earth fills sealed by the Middle Bronze fortifications. Analysis of

10

Oded Lipschits, Yuval Gadot, and Manfred Oeming

the pottery’s typology helped us to determine that the site was settled during the EB IIIA–B periods, parallel to Tel-Yarmuth Phase B-III/II and C-IV/II and Tell eṣ-Ṣafi/ Gath Phases E5 and E6 (de Miroschedji 2006; Greenfield, Shai, and Maeir 2016). Some of the finds may hint at earlier settlement, dating to the EB II, parallel to Phase C-V at Tel Yarmuth. The finds from Area S1 are exceptional. In this area, we exposed a well-made crushed lime floor. Under it, we found two articulated skeletons of donkeys that had been slaughtered and intentionally buried (Sapir-Hen, Gadot, and Lipschits forthcoming). The donkeys had died while young, and their corpses had been buried whole. The sacrifice of donkeys is a well-recorded Early Bronze Age phenomenon (see Arnold et al. 2016 for earlier references), but in most cases the animals were adults. The Azekah sacrificial burial is exceptional: it consists of two apparently special donkeys (as in all other cases): the two are quite young specimens, sacrificed well before they could have served for transport. Future expansion of the excavated section may help to better understand their context. No Early Bronze finds were found in Area E1, which is on the eastern slope of the tell. This was already noted in the survey (Emmanuilov 2012: fig. 26). Our explanation is that the eastern slope is an expansion of the site that began in the Late Bronze Age, when Azekah expanded beyond the Middle Bronze fortifications (see below), continued in the Iron Age, when a city wall was built at the edge of the site as it existed in this period, about 10 m to the east of the assumed location of the Middle Bronze fortifications (still to be confirmed in the field), and especially in the Hellenistic period, when the eastern side of the tell was reshaped (see below). All of this (still quite theoretical) reconstruction indicates that the site did not expand much to the east during the Early Bronze Age, and the top of the natural hill was much smaller than in the post-Middle Bronze period. Determining the nature of the site and whether it was fortified or not during this period awaits further excavation.

The Late Bronze Period Following our preliminary survey and its results, and based on the reanalysis of Bliss and Macalister’s excavations, we realized that during the Late Bronze Age the site had reached its peak. Therefore, the fact that occupational remains dating to this period had already been found in eight of the ten excavated areas did not come as a surprise. Though we already had some evidence of the existence of the site during the LB I and LB IIA, the excavations at Tel Azekah have yet to expose occupational layers from these periods. Two exceptional finds that attest to human settlement at the site should be mentioned in this respect. The first is a sherd of a White Slip II Type 2 Cypriot krater found in a fill in Area W1. This is a very rare find in the southern Levant, securely dated to Late Cypriot IIB, which corresponds to the Levantine LB IIA (Yasur-Landau et al. 2014). The second is an Egyptian bifacial plaque found out of context in Area S2, exhibiting iconographic motifs dated to the mid-18th Dynasty (Koch et al. forthcoming). These two discoveries, though found

Four Seasons of Excavations at Tel Azekah

11

out of context, may be interpreted as evidence for social stratification at Tel Azekah during the LB II. The main data from the Late Bronze Age derives from Area T2, located on the top of the mound, and Area S2, located on the lower terrace that surrounds the western and southern slopes of the mound, possibly an extramural quarter of the town. In Area S2, the chief finds dating to the LB IIB include remains of buildings that were constructed inside an abandoned rock-cut ditch dating to the Middle Bronze Age. The earliest occupational levels (S2-8, S2-7) consist of walls and floors. The limited exposure of these occupational levels does not allow reconstructing any clear plan, and thus the nature, function, and exact date of the buildings remain unknown. In the absence of pre-LB II pottery, the only conclusion that can be reached is that the buildings do not pre-date the 14th century BCE. The next occupational level, dated to the LB IIB, is characterized by the erection of a large building in the rock-cut ditch (S2–6). The building consists of two longitudinal rooms separated by a massive wall made of unworked boulders. This building was enlarged during the next occupational level (S2–5), by establishing a row of pillars on the western end of the western room. The pillars were laid on the edge of an artificial “step” in the rock-cut bedrock, thus creating a third longitudinal room to their west. Even though the exact nature and function of these buildings (the “boulder building” of Phase S2-6 and the “pillar building” of Phase S2-5) could not be determined, the use of large boulders together with the extent of the buildings imply that they were not merely of a domestic nature. Mud-brick debris and a few in situ broken vessels found on the floors of the latter “pillars” building indicate that it was destroyed sometime during the LB IIB. Other finds that can be associated with this phase include a scarab of Ramses II found out of context in Area N1 and other items that can be dated to the 19th Dynasty (Koch et al. forthcoming). During Phase S2-4, the entire formation of the building changed: the rock-cut ditch was filled with earth and stones, burying the former buildings under it and creating a leveled space in which an open-air paved plaza was built. The plaza included a cistern and a stone-built silo. A new building, which probably functioned as a warehouse, was erected next to it. Both the public plaza and the adjacent building were found under thick destruction debris dated to the LB III, probably during the second half of the 12th century BCE. 6 This destruction should be equated with a destruction layer also encountered in Areas T1, W2, and E3, located on the upper mound (below). Following the destruction of the public plaza, the extramural quarter was abandoned and its habitation resumed only in the Iron IIB. The most dramatic exposure of the end of the Late Bronze destruction was in Area T2, where we unearthed a destroyed architectural compound (Building T2/ F627). The structure includes two main parts: a northern room and an area that was 6.  While some scholars prefer to use the term Iron IA when referring to the second half of the 12th century BCE (e.g., Mazar 1990), we prefer to use the term LB III, because it reflects the continuation of Canaanite culture and of the Egyptian governing system.

12

Oded Lipschits, Yuval Gadot, and Manfred Oeming

Fig. 3.  A skeleton from the destruction of Building T2/F627 from the end of the 12th century BCE.

partitioned into three sub-spaces to the south, with evidence of several sub-phases of floor-raising. The ground plan, as far as can be inferred from its fragmentary nature, generally recalls the design of the pillared buildings or “patrician houses” that were popular during the Late Bronze Age (e.g., Mazar 1997: 157–69). 7 An analysis of the formation process shows that the northern room was roofed and that the upper area of the house was used for storage (Metzer 2015: 127–28; Metzer, Gadot, and Lipschits forthcoming). On the ground floor, in the middle of the room, an elaborate grinding installation with an adjacent collecting vat was built. This specialized architecture points to the exceptional function of the room, which is further emphasized by the pottery objects and other finds from this area. The destruction of Building T2/F627 was severe and complete. To date, four skeletons have been unearthed beneath the destruction. More than 100 complete vessels were unearthed in the destroyed building. The assemblage includes almost the entire range of pottery that can be found in southern Canaan during the LB IIB and III (Metzer, Gadot, and Lipschits forthcoming). A few vessel types make it possible, however, to narrow down the timeframe to the LB III, parallel to Lachish VI and close to the end of the 12th century BCE. It is important to note that missing 7.  To the north of the building at Area T2, evidence of another structure belonging to this phase was exposed; the relation to Building T2/F627 needs further investigation.

Four Seasons of Excavations at Tel Azekah

13

from the assemblage are Egyptian or Egyptianized ceramic items and Mycenaean IIIC 1b pottery. Five scarabs, eight amulets, a bulla, and a conoid seal were found in this complex (Koch et al. forthcoming). Three of the scarabs and three of the amulets were found together in the same context, close to the remains of a single skeleton (T2/ L220). Another concentration of finds includes one scarab and four amulets, found in the northeastern section of Area T2, close to the remains of another skeleton (T2/407). An additional scarab was found in the courtyard, located in the center of the complex, and an additional amulet was found in a room located in the southeast. A concentration of items such as these by the sides of the two skeletons most probably indicates that they were used by these individuals and thus raises questions regarding the reception and adaptation of Egyptian amulets by the local population (Koch et al. forthcoming). Another part of the destroyed city was exposed in Area E3, where the southeastern corner of the Acropolis was exposed. The architectural context of this destruction is presently unclear. The only wall that was part of the structure was destroyed; it was made of very large worked stones and seems to have been of a public nature. The collapse includes stones, burned mud brick, smashed pottery vessels, and charred wooden beams. In between the stones of the collapse, a standing stone was identified. We believe that the destroyed structure, along with other finds unearthed in the topsoil, was part of the city’s acropolis and served cultic purposes.

The Gap in the Iron I Following the 12th century BCE destruction, Tel Azekah remained abandoned for more than two centuries (the Iron I). The chronological gap was already noticed in the survey, when the absence of Philistine Bichrome style pottery (Philistine 2) was observed. After four excavation seasons, no such pottery, not even residual sherd material, has been uncovered. Azekah seems to suffer a fate similar to that of Lachish (Ussishkin 2004: 62–64, 70–71), unlike that of Beth-Shemesh (Bunimovitz and Lederman 2009: 120–21; Lederman and Bunimovitz 2014: 65) and Gezer (Dever et al. 1970: 23–24; Dever et al. 1974: 50–55; Dever 1986: 60–87; Ortiz and Wolf 2012: 12), both of which were destroyed at the end of the Late Bronze Age but were then resettled during Iron I.

Evidence for Iron Age IIA Resettlement of Azekah began in Iron IIA. The survey results did not suggest the existence of a large settlement during this period. Materials were found only in Areas T1 and T2, on the top of the tell, and their nature is still unclear. In Area T1, the finds were exposed just below a modern trench that cleared all later elements. The finds include a destruction layer with burned material, construction stones, slingshots, and intact vessels dating to Iron IIA. In Area T2, the finds include a domestic building, possibly of the four-room type, and a garbage pit nearby, filled

14

Oded Lipschits, Yuval Gadot, and Manfred Oeming

with a large amount of restorable pottery. A lamp and bowl deposit was found under the building’s floor, reflecting a well-known Canaanite habit that may suggest that, although the site was not settled for a long time, it was still occupied by descendants of the local population. Further investigation of this phase is needed before we can conclude that these finds are from the early or late Iron IIA; but in any case, these finds date prior to the destruction of Gath.

The Iron IIB and C Fragmented remains dating to the Iron IIB were found in most of our excavation areas. The spread of the finds fits the survey results, which collected Iron Age pottery from all of the survey’s fields. On the other hand, we expected to find the Iron IIB city in complete ruins, as was the case at Lachish and other sites in the Shephelah. But a clear destruction layer was found at this stage in two areas only. Because these remains were damaged by building activities dating to the Persian and Early Hellenistic periods, it is difficult to interpret their nature. Domestic architecture was found in Areas T1, T2, W1, and S1. The structure that was built during the Iron IIA in Area T2 was rebuilt along slightly different lines. In Area T1, we uncovered a small structure built of mud-brick walls, with a destruction layer on its floor. The best-preserved remains were found in Area S1, where parts of a building were unearthed. One of the building’s rooms was used for weaving, and a perfectly preserved loom structure was found, burned and fallen in place. The loom’s burned wooden beams were found beautifully articulated, together with 34 clay loom-weights, arranged in 2–3 rows; both of these discoveries marked the alignment of the loom. Pottery vessels found smashed on the floor, including among other vessels a torpedo jar and a lmlk jar, help date the destruction to the end of the 8th century BCE. Remains of more public architecture were found in Areas S2 and E3. In Area S2, the lower city, habitation was renewed after a long break and included an open public space. The exact nature of this area will only be understood after further excavation. Area E3, in the southeastern corner of the site, may possibly be the location of the Assyrian siege ramp, described in the “Azekah Inscription” (above, p.  2). Aerial photographs and measurements enabled us to compare the Azekah ramp to the one at neighboring Lachish, which was constructed with the same topographic and military logic in the same location, in Lachish’s southeastern corner. With support from the German–Israeli Foundation for Scientific Research and Development (GIF, Grant number 1238), we began a three-year project in 2015. Two massive walls were exposed at the base of the slope (they were thus not likely part of the city’s fortifications). The first was ca. 3 m wide and oriented north–south; we traced it for 15 m. The second wall is ca 2 m wide and had been built perpendicular to the slope. Its northwestern face abuts the first wall. Due to their similar size and alignment, we suggest that these two walls are part of the same system or a large

Four Seasons of Excavations at Tel Azekah

15

Fig. 4.  A building from the Iron IIB and C in Area S1.

structure that might be connected to the site’s lower fortifications or to activities carried on by a besieging army. Only continued excavation at the site will enable us to learn more about the architecture of this area and to understand the function of the walls and the period during which they were built. The importance of Azekah before the 701 BCE Assyrian campaign, and probably the settlement gap at the site in the first half of the 7th century BCE, can be confirmed by the discovery of the lmlk stamp impressions found there; all of the recognizable stamped handles are early types, and none are late types (Lipschits, Sergi, and Koch 2011: 30 and p. 33 n. 19). 8 Iron IIC finds at Azekah are very rare and are usually found in later mixed fills. We did not expose any structures that can be dated to this period, and the only place that we have a clear pottery assemblage is from the water cistern in Area S2. 8. Bliss (1899a, 1899b, 1900a) reported finding 17 lmlk stamp impressions at Azekah. Only 11 of these were drawn (Bliss 1899a: 104, pl. V:1–9; 1900a: 13). Eight have four-winged emblems (and thus belong to the early types). Three additional stamps bear two-winged emblems, and at least one has an undivided place-name and thus belongs to the early type as well (Bliss 1899b: pl. V:9). In their final report, Bliss and Macalister (1902: 107) counted a total of 13 lmlk stamp impressions bearing four-winged emblems. We therefore conclude that out of the 17 lmlk stamp impressions found at Azekah, at least 14 should be considered as early types and the rest as unidentified.

16

Oded Lipschits, Yuval Gadot, and Manfred Oeming

Pottery vessels dating to this period were found in the fill inside the cistern, and they indicate the last time the cistern was in use for storing water. However, even without many Iron IIC finds, it is clear that Azekah was a regional center during that period. To date, ten rosette stamped handles have been found at the site, and though this cannot be considered a massive number of handles, it is still the second most numerous quantity found at sites in the Judean Shephelah. 9 The current understanding of the site at the end of the Iron Age is that, just like Lachish, the size of settlement was very small, probably restricted to the fortress, with an access to the cistern in the lower area.

Unexpected Archaeological Results The Middle Bronze Age One of the major discoveries at Azekah is the unearthing of the city’s Middle Bronze Age western fortifications. Although the topography of the tell’s western slope indicates that there is a fortification wall buried below the surface (see already Dagan 2000: 200), the actual date and nature of the fortifications were unknown and the very few finds dating to the Middle Bronze Age in the initial survey did not suggest that the site was large enough to be fortified. The city wall had already been exposed in our first season of excavations in Area W1. We then continued to trace the wall farther to the south (Areas W2 and W3) and farther to the north (Area N1). At this stage, it is unclear how the wall continues southward and northward. In both cases, our assumption is that the current southern and eastern slopes of the tell do not represent its Middle Bronze Age contour and that the site was probably smaller at that time. The survey already indicated that the Middle Bronze site, like the earlier Early Bronze settlement, did not extend all the way to the eastern slope. This is now supported: we did not discover any eastern fortification line in Area E1 that can be understood as completing the eastern fortifications. If the wall is not completely eroded, then it must be located farther west than the top of E3, in an unexcavated part of the upper surface of the tell―in what we surmise is very close to the eastern squares of Area T2. The size and complexity of the fortifications makes their exposure very difficult and demanding. In Area W2, for example, we exposed two parallel wall lines but it is unclear if the walls are contemporary (and, furthermore, whether one of the walls is a structural element) or that they date to two different periods. Clearly, 9. A further indication of Azekah’s importance is the exceptional find of four unique stamped handles: sub-types IA5 (Bliss/Macalister 1902: pl. 56, no. 43z), IC7 (no. 35z), IC11 (no. 37z), and IIIB10 (no. 39z), according to the typology of Koch and Lipschits (2013). All the Shephelah sites yielded 51 rosette stamped handles, which is about 21.5% of the corpus. Of these, 24 were retrieved at Lachish (10.5% of the corpus). For the finds at Azekah, see Bliss and Macalister 1902: pl. 56: 35z–44z. A tenth rosette stamped handle was found at Azekah during a survey that was conducted by S. Emmanuilov in June 2009; see Koch and Lipschits 2013: 59.

Four Seasons of Excavations at Tel Azekah

17

Fig. 5.  Middle Bronze mud-brick fortification wall in Area N.

many more years of excavation and research are needed to reach firm conclusions in regard to this specific question and in regard to the composition, technique, and date of the fortifications in general. It seems, however, that at this stage some conclusions can be carefully proposed. To date, 10 meters of the wall in Area N1, 8.5 meters in Area W1, 13 meters in Area W2, and 6.5 meters in Area W3 have been exposed. As is demonstrated by the long section of the wall exposed in Area W2 the wall was not built in a straight line and includes insets and offsets. The wall reaches a width of 3 m and is built of a stone foundation and a mud-brick superstructure. Its outer face is made of wellworked stones, while the inner fill is made of field-stones. Approaching the base of the wall from the west is a glacis. It consists of a compacted layer of crushed lime rock sealing a thick layer of chalk stones from above. Further below the earth fill, pottery sherds dating to the EB III were found. Rectangular mud bricks were placed on top of the stone foundation. These were exposed in all areas. The best preserved mud bricks were found in Area N1, where they are in a standard size of 0.4 × 0.4 × 0.12 m, set in regular courses. On the outer surface of the mud-brick wall, a line of whitish material was visible; it was probably a coating to prevent damage to the edges from rain water. A sample of this material

18

Oded Lipschits, Yuval Gadot, and Manfred Oeming

was sent for examination, and the results suggest that it is made of high-quality lime, similar to the substance that was found in the Middle Bronze Age mud-brick gate at Ashkelon. 10 A basic and preliminary calculation of the number of mud bricks used to build the city-wall shows that one running meter of the wall (3 m wide and with the assumption of 3 m height) required 900 bricks. The minimal length of the wall was about 700 m. This means that 630,000 bricks were needed for the wall alone (without the towers and a gate). Every mud brick weighs about 5 kg; thus, 3,150 metric tons of mud had to be brought to the site, probably from the Elah Valley below and to the east of the site. Add to this the enormous number of mud bricks that were needed for houses and structures on the site and there is a ready explanation for the raising of the level of the site by more than 2 m between the Middle Bronze Age and Iron IIB. To the south, the city wall bonds with “Tower 3,” excavated in the past by Bliss and Macalister and erroneously dated to the Roman period (see above). The tower is made of three massive walls built of large worked stones and set deep into the slope and earlier layers. Eight courses of the wall have been exposed thus far. The tower might have served as a buttress, projecting westward from the line of the wall and protecting an ascent leading up toward a presumed gate. It is still unclear how the wall continued south of Tower 3 and how it is connected, if at all, with Towers 1 and 2. The dating of the city wall is still undetermined. The best method for dating a city wall would be to expose its inner side and its relation to architectural units built in relation to the wall or units being cut by it. Currently, we have managed to expose the inner face of the wall only in one area (W1) and, unfortunately, no architecture was found approaching the wall from the inside. In Areas N1, W2, and W3, the inner faces have not yet been exposed. We do, however, have supportive evidence from Area W2 that the fortifications date to the Middle Bronze Age. The first indication is a jar burial of an infant, found cutting into the mud bricks of the city wall. A cylindrical juglet that served as a burial gift was placed in the jar next to the infant and dates to the end of the Middle Bronze or the beginning of the Late Bronze Age. This proves that the wall was already standing at the time of the burial. More supportive evidence was found in a section that we cut into the glacis, where we found pottery that is mostly dated to the Early Bronze Age but some of which also dates to the Middle Bronze Age. Based on these finds, the building of the city wall should be dated to the MB IIa or IIb. It is possible that it was built along the lines of an earlier wall (from the EB III), but this remains to be determined.

10.  We wish to thank Mrs. Aliza van Zweden of the Israel Antiquities Authority for examining the mud bricks and sharing her knowledge with us.

Four Seasons of Excavations at Tel Azekah

19

The Late Persian / Early Hellenistic Period Only 1.7% of the sherds collected during the preliminary survey date to the Persian–Hellenistic period, and so our expectations were that the tell was only sporadically inhabited during this era. Thus, we were quite astonished to find that in most areas the first upper layer is dated to the Persian–Hellenistic period. Our excavations exposed a relatively large village or town dating to the 4th–3rd centuries BCE. Remains of this settlement were found in Areas N1, W1, W2, and S1. Three identifiable buildings have been exposed thus far: one in Area W1, one in Area N1, and one farther to the south in Area S1. These buildings each seem to consist of a central wide courtyard that was surrounded by built wings. The courtyard of the house in Area W1 includes at least five stone-lined silos, and we therefore named the structure “Granary Building.” The courtyard of the building in Area N1 includes an oven and a kiln. The nature of the building exposed in Area S1 is less clear, because we have only found its southeastern corner. The three buildings are located along the perimeter of the tell. Farther east and into the central areas of the tell we found only open grounds and garbage pits; it seems that it was an open space during the late Persian–early Hellenistic Periods. More finds from the late Persian and early Hellenistic periods were found in Area S2, in the lower terrace of the site. The water reservoir that was already in use in the Late Bronze period and continued in use in the Iron IIB was reused in the late Persian–early Hellenistic periods (Stage S1–2) as a burial area: remains of at least 16 people, some of them children and infants, were laid there, together with burial offerings. The use of this area as a burial place implies that the site at this period was concentrated only in the upper tell and that the lower terrace was outside the town limits. Chronologically, the buildings in Areas W1 and S1 were built into earlier Iron Age architecture, destroying the remains of what preceded them. This was probably also the case in Area N1. Since most of the Persian-period pottery that was found at Azekah was dated to the Late Persian period (Shatil forthcoming), it appears that there was a settlement gap at the site in the 6th century BCE and that the site was small and limited in size and population during the Early Persian period. The buildings were finally abandoned during the 3rd century BCE (Shatil forthcoming). The inhabitants left behind mostly large storage jars that they could not carry, similar to the abandonment process that took place at nearby Kh. Qeiyafa during this same period (Sandhaus and Kreimerman 2015: 254, 263–66).

Late Hellenistic / Late Roman Periods Following the abandonment of the upper tell during the 3rd century BCE, and possibly after a short gap, settlement at the site was renewed but this time only on its eastern slopes. Finds from Area E3 show that the slopes were used for the construction of a domestic quarter that existed from the 2nd century BCE and until the

20

Oded Lipschits, Yuval Gadot, and Manfred Oeming

Bar Kokhba Revolt in the second century CE. The upper tell was mostly abandoned, except for the fortress that crowned the highest point on the tell.

The Expected and the Unexpected Archaeological Results: Some Final Thoughts Based on the excavations at Azekah thus far, we can summarize some general conclusions regarding expected and unexpected results in the light of large-scale excavations at major sites. First, we can demonstrate that there are actual archaeological materials from periods that, based on clear historical data, we expected to find. This is especially true for the Iron II. It might be that, based on the historical sources (as in the case of the Iron IIB), the expectation of finding a large city will not materialize and that the archaeological results will confirm that only a small site (or possibly only a citadel) existed in this period; but in this case, it is probably a question of reconsidering the historical sources, interpreting them, and building the right expectations based on a revised interpretation. The absence of historical sources, however, means nothing, and the finds from the Middle Bronze and Late Bronze Ages, as well as from the Late Persian and Early Hellenistic periods, are good examples. As we learned from our experience at Ramat Rahel, even in periods that are very well documented, the absence of information about a site should not be used to prove anything about the site. Furthermore, we can demonstrate that there are actual archaeological discoveries from periods that we expected to find based on archaeological surveys and the careful study of previous old excavations, especially from periods for which we have very strong and clear indications, like the Early Bronze, the Late Bronze, and Iron IIB. Even the size of the settlements from these periods could be deduced, based on the location of the finds in the survey’s fields. However, the absence of finds in surveys means nothing, and the evidence from the Middle Bronze Age and from the Late Persian–Early Hellenistic periods provide good examples of this fact. The combination of historical research, finds from surveys, and analyses of old excavations helps to close some of the gaps in our knowledge regarding expected and unexpected finds. But we are especially cognizant that, in the case of Azekah, the absence of evidence of all sorts (limited historical sources, building techniques, nature of fortifications, etc.) means nothing, especially when it comes to the Middle Bronze Age and the Late Persian–Early Hellenistic periods. It seems that these two periods, each with its unique nature, deserves special attention and greater theoretical exploration and inquiry in order to understand the reasons for the gap that exists between archaeological surveys and historical documents on the one hand and the results of excavations on the other. To do this, however, is already beyond the scope of this summary.

Four Seasons of Excavations at Tel Azekah

21

Acknowledgments The excavations at Tel Azekah are directed by the authors under the auspices of the Institute of Archaeology at Tel Aviv University (Israel), in cooperation with Heidelberg University (Germany). Participating consortium institutions include Macquarie University (Australia), Oldenburg University (Germany), Charles University in Prague (Czech Republic), and King University and the University of Iowa (United States of America). The research was supported by the German–Israeli Foundation for Scientific Research and Development (GIF, Grant number 1238).

References Aharoni, Y. 1979 The Land of the Bible: A Historical Geography. London: Burns and Oates. Aharoni, Y. 1987 Eretz Israel in Biblical Times. A Geographical History. Jerusalem: Yad Izhak Ben-Zvi (Hebrew). Ahlström, G. W. 1982 Royal Administration and National Religion in Ancient Palestine. Leiden: Brill. Albright, W. F. 1924 Researches of the School in Western Judaea. BASOR 15: 2–11. Albright, W. F. 1949 The Archaeology of Palestine. London: Penguin. Arnold E. R.; Hartman G.; Greenfield, H. J.; Shai, I.; Babcock, L. E.; and Maeir, A. M. 2016 Isotopic Evidence for Early Trade in Animals between Old Kingdom Egypt and Canaan. PLoS ONE 11(6): e0157650doi:10.1371/journal.pone.0157650. Avi-Yonah, M., and Yeivin, S. 1955 The Antiquities of Israel. Tel Aviv: 120–323 (Hebrew). Ben-Arieh, Y. 1972 The Geographical Exploration of the Holy Land. PEQ 104: 81–92. Bliss, F. J. 1899a First Report on the Excavations at Tell Zakariya. PEFQSt: 10–25. 1899b Second Report on the Excavations at Tell Zakariya. PEFQSt: 89–111. 1899c Third Report on the Excavations at Tell Zakariya. PEFQSt: 170–187. 1900 Fourth Report on the Excavations at Tell Zakariya. PEFQSt: 7–19. Bliss, F. J., and Macalister, R. A. S. 1902 Excavations in Palestine during the Years 1898–1900. London: PEF. Bocher, E., and Lipschits, O. 2013 The Corpus of yršlm Stamp Impressions—The Final Link. Tel Aviv 40: 99–116. Bonn, A.G.; Moyer, H.; and Notis, M. R. 1993 The Typology and Archaeometallurgy of the Copper-base Artifacts. Pp. 203–20 in F. W. James and P. E. McGovern, eds., The Late Bronze Egyptian Garrison at Beth Shan: A Study of Levels VII and VIII. Philadelphia: University Museum. Bunimovitz, S., and Lederman, Z. 2009 The Archaeology of Border Communities: Renewed Excavations at Tel BethShemesh, Part 1: The Iron Age. Near Eastern Archaeology 72: 114–42.

22

Oded Lipschits, Yuval Gadot, and Manfred Oeming

Bunimovitz, S. and Zimhoni, O. 1990 “Lamp and Bowl” Foundation Deposits from the End of the Late Bronze Age – Beginning of the Iron Age in Eretz-Israel. Eretz-Israel 21: 41–55. (Hebrew, with English summary) Conder, C. R., and Kitchener, R. E. 1881 The Survey of Western Palestine, Vol. I: Galilee. London: Palestine Exploration Fund. 1881–1883  The Survey of Western Palestine, Vol. III. London: Palestine Exploration Fund. Dagan, Y. 1992 The Judean Shephelah in the Monarchic Period in Light of Archaeological Surveys and Excavations. M.A. thesis, Tel Aviv University. Tel Aviv. (Hebrew) 2000 The Settlement in the Judean Shephelah in the Second and First Millennium B.C.E.: A Test-Case of Settlement Processes in a Geographical Region. Ph.D. dissertation, Tel Aviv University. 2011 Tel Azekah. New Look at the Site and Its “Judean Fortress.” Pp. 71–86 in I. Finkelstein and N. Naʾaman, eds., The Fire Signals of Lachish: Studies in the Archaeology and History of Israel in the Late Bronze Age, Iron Age, and Persian Period in Honor of David Ussishkin. Winona Lake: Eisenbrauns. Decker, W. 1984 Spiel. Pp. 1150–52 in W. Helck and E. Otto, eds., Lexikon der Ägyptologie V. Wiesbaden: Harrassowitz. Dever, W. G. 1986 Gezer IV: The 1969–71 Seasons in Field VI, the “Acropolis.” Annual of the Nelson Glueck School of Biblical Archaeology. Jerusalem: Nelson Glueck School of Biblical Archaeology. Dever, W. G.; Lance, H. D.; and Wright, G. E. 1970 Gezer I: Preliminary Report of the 1964–1966 Seasons. Annual of the Nelson Glueck School of Biblical Archaeology. Jerusalem: Nelson Glueck School of Biblical Archaeology. Dever, W. G., et al. 1974 Gezer II: Report of the 1967–70 Seasons in Fields I and II. Annual of the Nelson Glueck School of Biblical Archaeology. Jerusalem: Nelson Glueck School of Biblical Archaeology. Di Segni, L. 1999 The ‘Onomasticon’ of Eusebius and the Madaba Map. Pp. 115–20 in M. Piccirillo and E. Alliata, eds., The Madaba Map Centenary 1897–1997: Travelling Through the Byzantine Umayyad Period. Proceedings of the International Conference Held in Amman, 7–9 April, 1997. Jerusalem: Studium Biblicum Franciscanum. Emmanuilov, S. 2012 The History of the Settlement at Tel Azekah in light of Archaeological Survey. M.A. thesis, Tel Aviv University. Tel Aviv. (Hebrew with English Summary) Fowler, H. N. 1900 Archaeological News. AJA 4: 247–48. Geva, H. 1985 The “First Wall” of Jerusalem during the Second Temple Period: An ArchitecturalChronological Note. Eretz-Israel 18: 21–39 (and pls. 7–8) (Hebrew).

Four Seasons of Excavations at Tel Azekah

23

Gitler, H. and Tal, O. 2006 The Coinage of Philistia of the Fifth and Fourth Centuries BC: A Study of the Earliest Coins of Palestine. Collezioni Numismatiche 6. Milan: Edizione Ennere / New York: Amphora Books / B& H Kreindler. Greenfield, H. J.; Shai, I.; and Maeir, A. 2016 Understanding Early Bronze Age Urban Patterns from the Perspective of NonElite Neighborhood: The Excavations at Tell eṣ-Ṣafi/Gath, Israel. Pp. 475–89 in volume 3 of R. A. Stucky, O. Kaelin, and H. P. Mathys, eds., Proceedings of the 9th International Congress on the Archaeology of the Ancient Near East, 9–13 June 2014, Basel. Wiesbaden: Harrassowitz. Hallo, W. H. 1993 Games in the Biblical World. Eretz-Israel 24: 83–88. Hallote, R. S. 2006 Bible, Map and Spade: the American Palestine Exploration Society, Frederick Jones Bliss and the Forgotten Story of Early American Biblical Archaeology. Piscataway, NJ: Gorgias. Keel, O., and Küchler, M. 1982 Orte und Landschaften der Bibel 2. Göttingen: Vandenhoeck & Ruprecht. Kloner, A., and Zissu, B. 2003 Hiding Complexes in Judah: An Archaeological and Geographical Update on the Area of the Bar Kokhba Revolt. Pp. 181–216 in P. Schäfer, ed., The Bar Kokhba War Reconsidered: New Perspectives on the Second Jewish Revolt against Rome. Tübingen: Mohr Siebeck. Koch, I., and Lipschits, O. 2013 The Rosette Stamped Jar Handle System and the Kingdom of Judah at the End of the First Temple Period. ZDPV 129: 55–78. Koch, I.; Metzer, S.; Gadot, Y.; Oeming, M.; and Lipschits, O. forthcoming  Agyptiaca from Tel Azekah in a Canaanite Cultural Context. Egypt and the Levant. Lederman, Z., and Bunimovitz, S. 2014 Canaanites, “Shephelites” and Those Who Will Become Judahites. Pp. 61–72 in G.  Stiebel et al., eds., New Studies in the Archaeology of Jerusalem and Its Region (VIII). Lipschits, O.; Sergi, O.; and Koch, I. 2010 The Date of the lmlk and ‘Private’ Stamp Impressions: A Fresh Look. Tel Aviv 37: 3–32. 2011 Judahite Stamped and Incised Jar Handles: A Tool for the Study of the History of Late Monarchic Judah. Tel Aviv 38: 5–41. Lipschits, O.; Gadot, Y.; and Oeming, M. 2012 Tel Azekah 113 Years After: Preliminary Evaluation of the Renewed Excavations at the Site. Near Eastern Archaeology 75: 196–206. Macalister, R. A. S. 1899 The Rock-Cuttings of Tell Zakariya. PEFQSt: 39–53. 1900 Further Notes on the Rock-Cuttings of Tell Zakariya. PEFQSt 39–53. 1912a The Excavations at Gezer. 1902–1905 and 1907–1909, Vol. II. London. 1912b The Excavations at Gezer. 1902–1905 and 1907–1909. Vol. III. London. Mazar, A. 1990 Iron Age I and II Towers at Giloh and the Israelite Settlement. IEJ 40: 77–101. 1997 Iron Age Chronology: A Reply to I. Finkelstein. Levant 29: 157–167.

24

Oded Lipschits, Yuval Gadot, and Manfred Oeming

Metzer, S. 2015 On the Eve of Destruction: Analyzing the Chronology, Function and Distribution pattern of a Late Bronze Pottery Assemblage from Tel Azekah. M.A. thesis, Tel Aviv University. Tel Aviv. Metzer, S.; Gadot, Y.; and Lipschits, O. forthcoming  A Snapshot of the Destruction Layer of Tel Azekah Seen against the Backdrop of the Final Days of the Late Bronze Age. ZDPV. Miroschedji, P. de 2006 At the Dawn of History: Sociopolitical Developments in Southwestern Canaan in Early Bronze Age III. Pp. 55–78 in A. M. Maeir and P. de Miroschedji, eds., ‘I Will Speak the Riddles of Ancient Times’: Archaeological and Historical Studies in Honor of Amihai Mazar on the Occasion of His Sixtieth Birthday. Winona Lake, IN: Eisenbrauns. Naʾaman, N. 1974 Sennacherib’s “Letter to God” on His Campaign to Judah. BASOR 214: 25–39. Napchan-Lavon, S. 2014 Bliss and Macalister’s Excavations at Tell Zakariya (1898–1899) in Light of Modern Research. M.A. thesis, Tel Aviv University. Tel Aviv. Napchan-Lavon, S.; Gadot, Y.; and Lipschits, O. 2014 Bliss and Macalister’s Excavations at Tell Zakariya (Tel Azekah) in Light of Published and Previously Unpublished Material. Pp. 74–95 in S. Wolff, ed., Villain or Visionary? R.A.S. Macalister and the Archaeology of Palestine. Palestine Exploration Fund Annual 12. Wakefield: Maney. Negev, A., and Gibson, S. 2001 Archaeological Encyclopedia of the Holy Land. New York: Continuum. Ortiz, S., and Wolf, S. 2012 Guarding the Border to Jerusalem: The Iron Age City of Gezer. NEA 75: 4–19. Peterson, B. J. E. 1975 ‘Brettspiel.’ Pp. 853–55 in W. Helck and E. Otto, ed., Lexikon der Ägyptologie I. Wiesbaden: Harrassowitz. Rainey, A. F. 1983 The Biblical Shephelah of Judah. BASOR 251: 1–22. Reich, R. 1990 Miqwaʾot (Jewish Ritual Immersion Baths) in Eretz-Israel in the Second Temple and the Mishnah and Talmud Periods. Ph.D. dissertation, The Hebrew University of Jerusalem, 2 vols. Jerusalem (Hebrew). Robinson, E., and Smith, E. 1856 Biblical Researches in Palestine, and in the Adjacent Regions. Vol. III. Boston: Crocker and Brewster. Sandhaus, D., and Kreimerman, I. 2015 The Late 4th/3rd Century BCE Transition in the Judean Hinterland in Light of the Pottery of Khirbet Qeiyafa. Tel Aviv 42: 251–71. Sapir-Hen, L.; Gadot, Y.; and Lipschits, O. forthcoming  Ceremonial Donkey Burial, Social Status and Settlement Hierarchy in the Early Bronze III: The Case of Tel Azekah. In: Lev-Tov, J., Wapnish, P., and Gilbert, A., eds., Taking A Broad View of Zooarchaeology in Honor of Brian Hesse’s Contributions to the Field. Atlanta: Lockwood Press. Schwarz, J. 1850 Descriptive Geography and Brief Historical Sketch of Palestine. Philadelphia.

Four Seasons of Excavations at Tel Azekah

25

Sebbane, M. 2000 Two Game Boards. Pp. 226–31 in Y. Hirschfeld, Y., ed., Ramat Hanadiv Excavations: Final Report of the 1984–1998 Seasons. Jerusalem: Israel Exploration Society: 226–231. 2001 Board Games from Canaan in the Early and Intermediate Bronze Ages and the Origin of the Egyptian Senet Game. Tel Aviv 18: 213–30. Seger, J. D. 1997 Azekah. P. 243 in E. M. Meyers, ed., The Oxford Encyclopedia of Archaeology in the Near East. Vol. 1. Oxford: Oxford University Press. Sharon, I. 1987 Phoenician and Greek Ashlar Construction Techniques at Tel Dor, Israel. BASOR 267: 21–42. Shatil, N. forthcoming  Pottery from the Persian and the early Hellenistic Periods from Tel Azekah. M.A. thesis, Tel Aviv University. Tel Aviv. (Hebrew with English Summary) Stern, E. 1971 Azekah. Pp. 133–37 in H. Tadmor H., ed., Encyclopedia Biblica, vol. 6. Jerusalem: Bialik Institute: Jerusalem. (Hebrew) 1993 Azekah. Pp. 123–24 in E. Stern, ed., NEAEHL I. Jerusalem: Israel Exploration Society / Washington, DC: Biblical Archaeology Society. Tadmor, H. 1958 The Campaigns of Sargon II of Assur: A Chronological-Historical Study. JCS 12: 77–100. Tepper, Y., and Shahar, Y. 1987 Hiding Complexes at Azekah. Pp. 171–85 in A. Kloner and Y. Tepper, Y., eds., The Hiding Complexes in the Judean Shephelah. Tel Aviv: Hakibbutz Hameuchad. (Hebrew) Torczyner, H. 1938 Lachish I: The Lachish Letters. London: Oxford University Press. Ussishkin, D. 2004 The Renewed Archaeological Excavations at Tel Lachish (1973–1994). Monograph Series of the Institute of Archaeology of Tel Aviv University. Tel Aviv: Institute of Archaeology of Tel Aviv University. Votaw, C. W. 1899 Present Excavations in Palestine. The Biblical World 14/6 (Dec.): 434–43. Watson, C.M. 1915 Palestine Exploration Fund: Fifty Years’ Work in the Holy Land, a Record and a Summary 1865–1915. London: PEF. Weippert, H. 1977 Spielgerät. Pp. 310–11 in K. Galling, ed., Biblisches Reallexikon. Tübingen: Mohr Siebeck. Yadin, Y.; Aharoni, Y.; Amiran, R.; Dothan, T.; Dunayevsky, I.; and Perrot, J. 1960 Hazor II. An Account of the Second Season of Excavations, 1956. Jerusalem: Magnes. Yasur-Landau, A.; Gross, B.; Gadot, Y.; Oeming, M.; and Lipschits, O. 2014 A Rare Cypriot Krater of the White Slip II Style from Azekah. IEJ 64: 1–8. Zissu, B. 2006 “A City Whose House-Roofs Form Its City Wall,” and a City that was not Encompassed by a Wall in the Days of Joshua Bin Nun – According to Archaeological Finds from Judea. Pp. 85–100 in Y. Eshel, ed., Judea and Samaria Research Studies 15. Kedumim—Ariel. (Hebrew)

This page left intentionally blank.

Swinging on the “Sorek Seesaw”: Tel Beth-Shemesh and the Sorek Valley in the Iron Age Shlomo Bunimovitz and Zvi Lederman

Introduction The biography of ancient Beth-Shemesh in the Iron Age is seemingly the story of a border site whose fate was determined by powerful external forces. Yet, a closer look brings into relief the important role played by the site’s inhabitants as active social agents taking part in shaping their own destiny within the changing political and cultural environment that surrounded them. Tel Beth-Shemesh is strategically situated on the southern bank of the Sorek Valley, at the 200 m topographical contour line that separates the Shephelah from the inner Coastal Plain (fig. 1). Overlooking the corridor leading from the coast inland via the Sorek brook, it guarded the entrance to the Judean hill country. Throughout the Iron Age, Beth-Shemesh was a border settlement due to its location at the juncture of Canaanite, Philistine, and Israelite territories and cultural spheres. The boundary between these political and cultural entities was not a fixed and rigid line but fluctuated and shifted, in resonance with the checkered history of the northern Shephelah. Little doubt is left by the historical sources of the time about the changing fortunes of the Sorek Valley sites and their complicated political history (Rainey 1983; Mazar 1994; Rainey and Notley 2006). The question posed by this intriguing situation is twofold. Can archaeology help in reconstructing the political history of the region during the Iron Age? And, more importantly, can it portray and explain the social and cultural history of the valley at that period? The Sorek Valley and its environs have been the focus of intensive archaeological investigation since the beginning of the 20th century. Two early cycles of excavations were conducted at Tel Beth-Shemesh in 1911–1912 by Duncan Mackenzie on behalf of the Palestine Exploration Fund (1911; 1913; Mackenzie et al. 2016) and in 1928–1933 by Elihu Grant from Haverford College, Pennsylvania (Grant and Wright 1939). More recently, Tel Batash-Timnah was excavated by Amiahi Mazar (1997; Mazar and Panitz-Cohen 2001; Panitz-Cohen and Mazar 2006) and Tel Miqne-Ekron by Trude Dothan and Seymor Gitin (2008). Our renewed excavations at Tel Beth-Shemesh since 1990 complement the investigation of the main sites in the Sorek Valley and shed new light on Iron Age inter-site relations in the valley. 27

28

Shlomo Bunimovitz and Zvi Lederman

Fig. 1.  The Sorek Valley and adjacent regions.

Moreover, they provide new insights into social and cultural processes relevant not only on a local scale but for the Shephelah as a whole and even beyond (Bunimovitz and Lederman 2009; 2014; 2016). A longue durée perspective of the Iron Age in the Sorek Valley enables us to identify an intriguing pattern—the “Sorek Valley Seesaw.” As will be elaborated below, this pattern indicates that Philistia and the Kingdom of Judah gained power at their border zone alternately, depending on the relative strength of the antagonist that each saw on the other side of the border. The “seesaw” describes not only political and cultural changes at the border zone but also subsequent significant shifts in the border itself.

Setting the Seesaw: Philistine Settlement and Canaanite Resistance The arrival and settlement of the Philistines in southern Canaan marks a watershed in the geopolitics and cultural history of the region. Insight into the effect of the Philistines on Beth-Shemesh and its environs—the establishment of the “Sorek Seesaw”—is gained both by a look at Late Bronze Age/Iron Age I settlement patterns in the Shephelah and a comparative study of the Iron Age I cultural assemblages of the main sites along the Sorek Valley. At the end of the Late Bronze Age, the southern coastal plain of Canaan and the adjacent Shephelah were dotted with dozens of sites, undoubtedly the most

Tel Beth-Shemesh and the Sorek Valley in the Iron Age

29

Fig. 2.  Pottery of Level 6 at Tel BethShemesh reflecting strong Canaanite traditions.

densely populated area in the entire country at that time. In the Iron Age I, however, the number of sites dropped drastically (Finkelstein 1996; 2000; Faust 2013a: 204, 206–8). Further inquiry reveals a two-fold change in the settlement pattern of the southern part of the region, between Lachish and Tell eṣ-Ṣafi/Gath: on the one hand, there was an almost complete abandonment of the countryside and, on the other, there was a great expansion of urban life. In sheer contrast, settlement in the northern Shephelah, from the Sorek Valley to the region of Gezer, continued almost unchanged (Shavit 2000: 215–17; Singer 1985:116–18). This pattern was interpreted by Bunimovitz (1998: 107–8; see also Shavit 2008:154–60) as reflecting Philistine forced synoecism—a purposeful displacement of the Canaanite rural population from their own territory, relocating them in the main Philistine centers. Canaanite settlements at the periphery of heartland Philistia escaped this hostile takeover, at least for the initial stage of Philistine settlement (marked by the appearance of locally produced Mycenaean IIIC1 pottery; see also Faust and Katz 2011). In the 12th century BCE, Ekron, one of the Philistine pentapolis, was established on top of the small Canaanite settlement at Tel Miqne, on the western outskirts of the Sorek Valley. The unequivocal new elements in the material culture of Strata VII—IV, such as high percentages of Aegean-style pottery of the Mycenaean IIIC1 and Bichrome phases (41%, Dothan and Zukerman 2004: table 1); considerable

30

Shlomo Bunimovitz and Zvi Lederman

pork consumption (26%, Lev-Tov 2006: table 6.3, chart 6.1; for Philistine foodways, see Yasur-Landau 2010: 228–40, 295–300; Maeir 2015); hearths; cylindrical loom-weights; figurines (Ben-Shlomo 2010) and more, indicate that Philistine immigrants inhabited and presumably governed the large Iron I city (for the current debate concerning the contents, origins, and meaning of Philistine material culture see, for example, Faust and Lev-Tov 2011; 2014; Maeir, Hitchcock, and Kolska Horwitz 2013; Hitchcock and Maeir 2013; with additional literature). Close to Tel Miqne-Ekron, the Canaanite town of Tel Batash-Timnah was soon annexed to the territory and cultural sphere of its mighty Philistine neighbor. The combination of distinctive changes in the overall architectural planning and building traditions of Tel Batash Stratum V, a great amount (34%) of Bichrome Philistine pottery, pig consumption (8%), and a variety of Philistine affiliated items all seem to indicate that in the Iron Age I this site became a “daughter” settlement of Philistine Ekron (Mazar 1994: 251; Mazar and Panitz-Cohen 2001: 277; 2006: 137). Just 7 km east of Tel Batash, Tel Beth-Shemesh presents a drastically different picture. In a sequence of four successive Iron Age I occupation phases (Levels 7–4, 12th–10th centuries BCE), Canaanite cultural traditions appear to be dominant (in architecture, pottery, bronze production, and more—Bunimovitz and Lederman 2009: 121–23; 2016: 159–245; Ashkenazi, Bunimovitz, and Stern 2016; fig. 2). Locally produced Mycenaean IIIC1 pottery is completely missing, and only a meager amount of Aegean-style pottery (decorated and undecorated) of the Bichrome phase (about 5% of the total amount of pottery recovered in Levels 6–5) reached the site. Other items of Philistine type or affiliation are also missing. Furthermore, pork consumption was completely avoided at Beth-Shemesh, in contrast to its heavy usage in the cuisine of adjacent Philistine sites (Bunimovitz and Lederman 2011a: fig. 8). This conspicuous difference between closely neighboring sites cannot be related to economic and ecological factors and must be interpreted within a cultural perspective. In line with anthropological theory as well as ethnographic and ethno­archaeological case studies, we have therefore argued in a number of recent publications that the settlement of the Philistines in southern Canaan and their later expansion out of their core area created new social and cultural boundaries in the region due to competition for agricultural land and resources (Bunimovitz and Lederman 2008; 2009:120–24; 2011a; 2014; 2016: 227–31). The Canaanite population inhabiting the territories occupied by the Philistines was thoroughly exposed now to the idiosyncratic cultural habits of its foreign masters and may have undergone a process of hybridization. Other Canaanites, located at the periphery of Philistia, for example, in the Sorek Valley, had the freedom of choice either to join their new aggressive neighbors or to resist them by maintaining their own identity. One conspicuous way to emphasize their “otherness” vis-à-vis Philistine identity– apparent at Beth-Shemesh—would have been to avoid pork consumption, in contrast to Philistine dietary habits. Another cultural distinction could be achieved by rejecting Philistine pottery that was also considered as an ethnic marker due to its

Tel Beth-Shemesh and the Sorek Valley in the Iron Age

31

symbolic association with Philistine drinking habits. The same rationale was applied to other items affiliated with the Philistine cultural sphere. Philistine settlement in the western part of the Sorek Valley upset the region’s equilibrium and seemingly pushed the Sorek seesaw to its first swing—Philistines “up,” Canaanites “down.” Nonetheless, our analysis has clarified that matters were more complicated than suggested by this simplistic metaphor. When political and social conditions in the Shephelah changed due to the coming of the Philistines, the indigenous Canaanite population of Beth-Shemesh living at the periphery of Philistine territory showed remarkable cultural persistence and resistance. By denying Philistine ethnic markers and behavior, they culturally identified themselves as “non-Philistines.” A similar phenomenon is apparent also at other sites in the eastern Shephelah (Faust and Katz 2011; Faust 2012; 2013a) and culminated with the establishment of the fortified site of Khirbet Qeiyafa by the local Canaanite population (for the similar “Shephelite” cultural affinities of the contemporaneous settlements of Khirbet Qeiyafa and Beth-Shemesh Level 4, see Lederman and Bunimovitz 2014; also Naʾaman 2010; for the excavators’ different opinion, see Garfinkel, Kreimerman, and Zilberg 2016, with previous literature). The continuous pressure exercised by the Philistines on the resisting Canaanites at their eastern periphery reached its apogee at the end of Iron Age I with the violent destruction of both Beth-Shemesh (Level 4) and Khirbet Qeiyafa. Apparently, this unfortunate situation made the Canaanites of the eastern Shephelah natural allies of the emerging Judahite kingdom during the first swing of the Sorek seesaw.

First Swing: The Early Judahite State at the Philistine Border The 10th century BCE witnessed a major change in power relations along the Sorek Valley. First and foremost, the huge Philistine city of Stratum IV at Tel MiqneEkron shrank dramatically in size and decreased in importance for the next two centuries, until the late 8th century BCE. This surprising phenomenon has been interpreted variously: Dothan (1989: 9–12) and Finkelstein (1995: 232; 2002: 116) relate Ekron’s decline to the renewed Egyptian interest in southern Canaan, which culminated in Shishak’s campaign. Gitin (1989: 41), Mazar (1994: 253–54; see also Mazar and Panitz-Cohen 2001: 278), and Faust (2013b), however, envisage the rise of the United Monarchy as the main source of negative effect on Ekron’s status and prosperity. The decline of Ekron may explain the abandonment of the 10th century BCE town at Tel Batash-Timnah (Stratum IV) and the ensuing gap of occupation at this site until the 8th century BCE. Mazar (1994: 253–56; 1997: 255; see also Mazar and Panitz-Cohen 2001: 277–79) considers the settlement of Tel Batash Stratum IV to be Israelite, belonging to the emerging Israelite state. This conclusion is based on a correlation between a new technique of pottery decoration that appears in Stratum IV—red slip and hand-burnish—and the rise of the kingdom of David and

32

Shlomo Bunimovitz and Zvi Lederman

Solomon. Furthermore, the construction of Tel Batash Stratum IV is thought to reflect Israelite expansion into the territory of the declining Philistine city-state of Ekron. We have suggested a different interpretation for the modest and short-lived 10th century BCE settlement at Tel Batash-Timnah (Bunimovitz and Lederman 2006). In our opinion, red slip and burnish—shown by Mazar (1998) to appear first on the northern coast of Philistia—cannot be used as ethnic markers. Rather than Israelite, the population of Tel Batash Stratum IV probably continued to be comprised of a mixture of Canaanites and Philistines. A comparative study of the early Iron II pottery assemblages from the Sorek sites reveals that Tel Batash was deserted a short time after the destruction of the large Iron I city at Ekron (Bunimovitz and Lederman 2006). The long abandonment and revival in the 8th century BCE of Tel Batash-Timnah seems also to be in harmony with the decline and recovery of its mighty neighbor Ekron. The correlation between the fall and rise of both sites cannot be incidental and should be related to the same historical processes. In contrast with its declining neighbors, Beth-Shemesh thrived during the 10th–8th centuries BCE. Grant and Wright (1939: 15–16, 67–71; Wright 1976: 252) were of the opinion that, in the days of David and Solomon, Beth-Shemesh (Stratum IIa) was replanned and fortified. The appearance of public buildings such as a “Governor’s Residence,” a storehouse, and a large silo left no doubt in their minds that Beth-Shemesh became the administrative center of Solomon’s second district, as implied in 1 Kgs 4. Indeed, our renewed excavations confirm the idea that in early Iron Age II the village of Beth-Shemesh was transformed into a city, with all of the symbols of central political government. In previous publications, we already presented in detail the main finds from our Level 3 (= Grant and Wright’s Strata IIa and IIb), which included fortifications, public buildings, redistributive economic center, iron workshop, and an impressive underground water reservoir. Furthermore, we have discussed their important implications for the debate about the early monarchy in Judah, suggesting a new perspective for the problem—“a view from the border” (Bunimovitz and Lederman 2001; 2006; 2009: 124–36; 2016: 281–82, 370–79). It is therefore sufficient here to emphasize that, in the three cycles of excavations at Beth-Shemesh, a large enough area has been exposed to afford a rare view of a city from the early days of the Monarchy, with all of its functional details. A plan of the site integrating our finds with the fragmented information left by our predecessors clearly reflects the major transformation experienced Beth-Shemesh (fig. 3). The plan highlights the markers of central government that suddenly appeared at the site: a series of public buildings that, according to our judgment, were established during the second half of the 10th century and the beginning of the 9th century BCE. What were the circumstances for the profound organizational change that transformed Beth-Shemesh from an unfenced village into a fortified urban center on Judah’s border with Philistia? An obvious answer to this question would be

Tel Beth-Shemesh and the Sorek Valley in the Iron Age

33

Fig. 3.  Integrated map of Iron Age IIA public buildings at Tel Beth-Shemesh: “Residency” and large silo (1928–1930); storehouse (1930); public building (Area B); water reservoir and fortifications (Area C); “commercial area” and earlier iron-smithing workshop (Area E).

that the young Israelite monarchy took advantage of the decline of Ekron and its “daughters” to establish its presence in the Shephelah. Nonetheless, the lack of clear archaeological evidence for Israelite expansion into the territory of Ekron renders this conclusion too simplistic. As will be remembered, in Iron Age I, Philistine pressure created a cultural boundary within the Sorek Valley. The decline of Tel Miqne-Ekron and Tel Batash-Timnah in the 10th century BCE must have diminished the tension between the different groups living along the valley. New opportunities were opened for intercultural contact, as well as for renewed ethnic ambiguity and boundary crossing. Paradoxically, it is in this very hour, when direct Philistine threat in the valley had almost disappeared, that the young monarchy emerging in the hill-country region had to maintain a close eye on its periphery. Now was the time to define its territory, to consolidate its hold on border

34

Shlomo Bunimovitz and Zvi Lederman

communities that might slip away, and to politicize the ethnic entity that would become a nation. Thus, the village of Beth-Shemesh was turned into a border town in the Sorek Valley, with all of the symbols of centralized political power. The foundation of monumental buildings served both as propaganda as well as practical functions. The image of the site was completely changed, and the political loyalty of its inhabitants was assured.

Second Swing: Israel–Judah Rivalry and Philistine Recovery From the end of the 10th to the 8th centuries BCE, Beth-Shemesh was the central city in the Sorek Valley. Naʾaman (2003a: 54) has equated its important status as a border town in Judah with that of Gezer in the kingdom of Israel. However, in the first half of the 8th century BCE, the geopolitical status in the Sorek Valley changed dramatically once again. A massive destruction affected the flourishing administrative city of Level 3 at Beth-Shemesh, leaving its public buildings burned to the ground. Grant and Wright (1939: 14) already had observed this destruction as terminating their Stratum IIb, but they were hesitant to identify the agent. In addition to the clash between Joash king of Israel and Amaziah king of Judah, which ended with the latter’s defeat (2 Kgs 14), their list of possible candidates includes the campaigns of the Assyrian kings Tiglath-pileser III and Sargon II, as well as the Philistine attack in the Shephelah recorded in 2 Chr 28:18. The Assyrian campaigns that took place late in the 8th century BCE do not coincide with the pottery chronology of Level 3 at Beth-Shemesh, and the narrative about a Philistine attack on the Shephelah in the days of King Ahaz has been reconsidered by modern scholarship and pegged as an ideological fiction (Naʾaman 2003a, with earlier literature) Another agent that may have been responsible for the destruction of Level 3 at Beth-Shemesh, ignored at the time by Grant and Wright but highly popular today among students of the Iron Age IIA–IIB transition, is the earthquake that occurred during the days of Uzziah, ca. 760 BCE (Amos 1:1; Zech 14:5; Austin, Franz, and Frost 2000; see, e.g., Ussishkin 2004: 83; Herzog and Singer-Avitz 2004: 230). However, we think that the idea that a fierce earthquake destroyed major sites in Judah is nothing more than a factoid constructed by current archaeological scholarship, bearing no correspondence to any reality in the past; instead, we have argued that the massive destruction of Level 3 at Beth-Shemesh shows clear signs of human agency (Bunimovitz and Lederman 2011b: 43–45; 2016: 363–65; see also Faust 2005: 106–7; Fantalkin and Finkelstein 2006: 22–24). We are therefore inclined to assign this destruction to the clash between Jeohash and Amaziah that took place at Beth-Shemesh at the beginning of the 8th century, ca. 790 BCE. It seems that the eclipse of Beth-Shemesh as Judah’s main border town in the Sorek basin paved the way for the resurrection of Philistine Ekron-Tel Miqne.

Tel Beth-Shemesh and the Sorek Valley in the Iron Age

35

Fig. 4.  Remains of a lever-and-weights olive oil press installation from Level 2 at Tel Beth-Shemesh.

Relying on Assyrian sources, Schniedewind (1998: 75) and Naʾaman (2003b: 84–85) have suggested that Ekron’s recovery and renewed growth (Stratum Ic), following the long decline that began in the 10th century BCE, began as early as the second half of the 8th century BCE. “New Ekron” probably stood behind the reestablishment of its traditional daughter-settlement, Timnah-Tel Batash (Stratum III), and not King Uzziah, as suggested by Mazar (1994: 256–57; also Mazar and PanitzCohen 2001: 279). Yet another attractive possibility is that the new stronghold of Tel Batash III was established by King Hezekiah, and if so, this event might be related to the third major geopolitical change in the Sorek Valley.

Third Swing: Hezekiah’s Rebellion and Western Expansion Following the destruction of Level 3 at Beth-Shemesh, a large and unfenced olive-oil producing settlement (Level 2) was established at the site. At least 18 oil presses of the lever-and-weights type and numerous simple oil-production installations were found at Tel Beth-Shemesh by the three expeditions excavating at the site (fig. 4; for further details and references, see Mackenzie et al. 2016: 134– 39; Bunimovitz and Lederman 2016: 419–69). As evident from other sites in the

36

Shlomo Bunimovitz and Zvi Lederman

Fig. 5.  lmlk (1) and official (2) seal impressions from Level 2 at Tel Beth-Shemesh.

Shephelah, such as Tell Beit-Mirsim (Albright 1943: 56), oil production was economically important to the kingdom of Judah, especially in the later part of the 8th century BCE. The archaeological evidence indicates that the Judahite oil industry at that time was based on a semi-specialized cottage industry and was incorporated within the social framework of the peasant communities in the Shephelah (Bunimovitz and Lederman 2011b: 46–47; 2016: 464–65). The change in the character of Beth-Shemesh from an administrative center on the border of Judah and Philistia (Level 3) to an unfortified agricultural/industrial town (Level 2) seems to be puzzling; once again, however, the reasons behind this profound change lie in the Sorek Valley’s border geopolitics. Both biblical and Assyrian sources inform us that, as part of King Hezekiah’s revolt against the Assyrian emperor Sennacherib, he took measures to secure the approaches to the Judahite hills and expanded Judah’s western border in the Sorek Valley and its surroundings (for discussion and references, see Bunimovitz and Lederman 2016: 51–52). His interference in Ekron’s political affairs, reported in Sennacherib’s annals, probably led to a take-over of Timnah-Tel Batash III (Mazar 1994: 258–59; Mazar and Panitz-Cohen 2001: 280). In fact, it is reasonable to imagine Hezekiah as the builder of the fortified settlement at the place, securing it with a garrison provisioned with tenths of lmlk jars. Further information concerning Hezekiah’s western expansion may be gleaned from the distribution of lmlk seal impressions, which are now unanimously associated with his administration.

Tel Beth-Shemesh and the Sorek Valley in the Iron Age

37

In the northern Shephelah, seal impressions of this type have been found, as well as at Beth-Shemesh and Azekah, at Tell es-Safi/Gath, Tel Miqne-Ekron, Tel BatashTimnah, and Gezer. It thus appears that the latter cities were incorporated into Hezekiah’s kingdom (Naʾaman 1979: 75–77; Vaughn 1999: 109–10, 185–97). The numerous lmlk and official stamp impressions found at Beth-Shemesh indicate that, despite its transformation, Beth-Shemesh must have played a significant role in the economic infrastructure of Hezekiah’s kingdom (Vaughn 2016; fig. 5 here). According to Vaughn (1999: 81–167), the royal jars had already circulated among both fortified and unfortified sites in Judah over a number of years prior to Sennacherib’s attack, reflecting a broad effort to strengthen the kingdom against the anticipated military response to Hezekiah’s rebellion. Beth-Shemesh, with its high number and great variety of lmlk and official impressions, seems, therefore, to have functioned as one of the major storage depots and/or distribution centers in this effort. It is certainly possible that the lmlk jars at Beth-Shemesh had to do with the intensive olive-oil production at the site and shipment of the product to other sites around the kingdom (cf. Vaughn 1999: 155).

Fourth Swing: Sennacherib’s Campaign and Its Aftermath Sennacherib’s campaign against Judah in 701 BCE brought with it the fourth major geopolitical turn in the Sorek Valley. Biblical sources, Assyrian inscriptions, and the archaeological evidence all indicate that it wreaked heavy destruction on the Shephelah of Judah. Many sites, including Beth-Shemesh, Level 2, were destroyed and abandoned (Bunimovitz and Lederman 2016: 466–67; Dagan 2004: 2681–82; Finkelstein and Naʾaman 2004; Faust 2008; 2013a: 214–15). Sennacherib boasted in his annals that he conquered 46 fortified cities in Judah and countless rural villages, deporting more than 200,000 people. Moreover, he recounts diminishing the kingdom of Hezekiah by tearing off parts of his state—presumably in the Shephelah—and giving them to the Philistine city-states Ekron, Ashdod, and Gaza (Oppenheim 1969: 288) Thus, in the early 7th century BCE, the border zone between Judah and Philistia changed radically as it was emptied of its population. Across the border, in Philistia, the situation in the wake of Sennacherib’s campaign was very different. As shown by archaeological research, the renewed Assyrian domination and rule generated a burst of immense economic growth. Ekron became one of the largest olive-oil-producing centers in the Levant (Gitin 1989; 1995; 1997). In order to produce thousands of tons of oil and to feed the masses of production workers, the Assyrians and their Philistine vassals needed the olive trees and fertile fields of the Shephelah of Judah. It is likely that the olive groves around Beth-Shemesh and its wheat fields in the Sorek Valley were controlled by Ekron by means of its daughter-settlement Timnah, which was rehabilitated (Stratum II) and also produced olive oil (Mazar 1994: 260–61; Mazar and Panitz-Cohen 2001: 281). For decades, the Judahite residents of the Shephelah were unable to return to their

38

Shlomo Bunimovitz and Zvi Lederman

Fig. 6.  Pottery vessels found on a hewn plastered bench at the entrance to the underground water reservoir of Beth-Shemesh. They represent the very last use of the reservoir prior to its brutal blockage.

fertile fields and groves, which were cultivated by others. This unfortunate situation is echoed in the lament of the prophet Isaiah who witnessed the destruction carried out by Sennacherib and its dire consequences: “Your country lies desolate, your cities are burned with fire; in your very presence aliens devour your land” (Isa 1:7). Unlike the peripheral regions of the Judahite kingdom (e.g., the Beersheba Valley and the Judean Desert), which witnessed accelerated growth and economic revival in the 7th century BCE, Judahite settlement in the Shephelah seems to have only partially recovered late in that century (for further discussion, see Bunimovitz and Lederman 2016: 54, 152–53; Faust 2008). Our discovery and excavation of the main underground water reservoir of Beth-Shemesh revealed a previously unknown chapter in the history of the site, shedding new light on the fate of Judahite presence the Shephelah in the aftermath of Sennecharib’s campaign (Bunimovitz and Lederman 2003). Finds within the reservoir and at its entrance structure indicate that a few Judahite families returned to Beth-Shemesh sometime in the mid7th century BCE and reactivated the water reservoir (fig. 6). However, this modest trial to resettle Beth-Shemesh was soon suppressed with an iron fist. Presumably, the Philistine neighbors of Beth-Shemesh—Ekron and its daughter settlement, Tel Batash—now incorporating the olive groves and grain fields of Beth-Shemesh into

Tel Beth-Shemesh and the Sorek Valley in the Iron Age

39

their thriving olive oil industry, deliberately sealed the water reservoir, the site’s source of life. Beth-Shemesh was finally deserted and never recovered.

Further Iron Age “Seesaws” in the Shephelah? The above longue durée perspective of the Sorek Valley during the Iron Age brings into relief an intriguing pattern of power relations: when the Judahite sites were “up,” the Philistine sites were “down,” and vice versa. This “seesaw” pattern of alternating prosperity and decline of the two rival polities sharing the valley seems to repeat at other points on Judah’s border with Philistia. Thus, for example, Lachish and Gath–Tell es-Safi presumably had their own seesaw to swing on, but the details of the power exchanges between the two sites and its schedule are not the same as in the Sorek Valley (Bunimovitz and Lederman 2011b). Our seesaw model was developed as an interpretive metaphor for the archaeological data related to the Iron Age in the Sorek Valley. Yet, we have no doubt that the intensive current archaeological research in other parts of the Shephelah (summarized in this book) will reveal more “seesaws” that will help to clarify the spatially and diachronically checkered picture of power relations along the shifting border between Judah and Philistia.

Acknowledgments The excavations at Tel Beth-Shemesh are directed by the authors under the auspices of the Institute of Archaeology of Tel Aviv University. Participating consortium institutions include Harding University, Arkansas, USA and the University of Lethbridge, Alberta, Canada. The research was supported by the Israel Science foundation (ISF) (grant nos. 898/99; 980/03; 1068/11), the Memorial Foundation for Jewish Culture, and by an Early Israel grant (New Horizons project), Tel Aviv University.

References Albright, W. F. 1943 The Excavation of Tell Beit Mirsim, Vol. III: The Iron Age. Annual of the American Schools of Oriental Research 21–22. Boston. Ashkenazi, D., Bunimovitz, S., and Stern, A. 2016 Archaeometallurgical Investigation of Thirteenth- Twelfth Centuries B.C.E. Bronze Objects from Tel Beth-Shemesh, Israel. Journal of Archaeological Science: Reports 6: 170–81. Austin, S. A., Franz, G. W., and Frost, E. G. 2000 Amos’s Earthquake: An Extraordinary Middle East Seismic Event of 750 B.C. International Geology Review 12: 657–71. Ben-Shlomo, D. 2010 Philistine Iconography: A Wealth of Style and Symbolism. Orbis Biblicus et Orientalis 241. Fribourg: Academic Press.

40

Shlomo Bunimovitz and Zvi Lederman

Bunimovitz, S. 1998 Sea Peoples in Cyprus and Israel: A Comparative Study of Immigration Processes. Pp. 103–13 in Mediterranean Peoples in Transition: Thirteenth to Early Tenth Centuries B.C.E.: In Honor of Professor Trude Dothan, eds. S. Gitin, A. Mazar, and E. Stern. Jerusalem: Israel Exploration Society. Bunimovitz, S., and Lederman, Z. 2001 The Iron Age Fortifications of Tel Beth-Shemesh: A 1990–2000 Perspective. Israel Exploration Journal 51: 121–48. 2003 The Final Destruction of Beth-Shemesh and the Pax Assyriaca in the Judahite Shephelah. Tel Aviv 30: 3–26. 2006. The Early Israelite Monarchy in the Sorek Valley: Tel Beth-Shemesh and Tel Batash in the 10th and 9th Centuries B.C.E. Pp. 407–27 in I Will Speak the Riddles of Ancient Times: Archeological and Historical Studies in Honor of Amihai Mazar on the Occasion of his Sixtieth Birthday, eds. A. M. Maeir and P. de Miroschedji. Winona Lake, IN: Eisenbrauns. 2008 A Border Case: Beth-Shemesh and the Rise of Ancient Israel. Pp. 21–31 in Israel in Transition: From the Late Bronze II to Iron IIa (c. 1250–850 B.C.E), Vol. I: The Archaeology, ed. L. L. Grabbe. New York: T. & T. Clark. 2009 The Archaeology of Border Communities: Tel Beth-Shemesh Renewed Excavations, Part 1: The Iron Age. Near Eastern Archaeology 72: 116–44. 2011a Canaanite Resistance: The Philistines and Beth-Shemesh—A Case Study from Iron Age I. Bulletin of the American Schools of Oriental Research 364: 37–51. 2011b Close yet Apart: Diverse Cultural Dynamics at Iron Age Beth-Shemesh and Lachish. Pp. 33–53 in The Fire Signals of Lachish: Studies in the Archaeology and History of Israel in the Late Bronze Age, Iron Age and Persian Period in Honor of David Ussishkin, ed. I. Finkelstein and N. Naʾaman. Winona Lake, IN: Eisenbrauns. 2014 Migration, Hybridization and Resistance: Identity Dynamics in Early Iron Age Southern Levant. Pp. 252–65 in The Cambridge Prehistory of the Bronze and Iron Age Mediterranean, ed. A. B. Knapp and P. van Dommelen. Cambridge: Cambridge University Press. 2016 Tel Beth-Shemesh: A Border Community in Judah: Renewed Excavations 1990–2000: The Iron Age. Institute of Archaeology Monograph Series 34. Tel Aviv: Institute of Archaeology, Tel Aviv University / Winona Lake, IN: Eisenbrauns. Dagan, Y. 2004 Results of the Survey: Settlement Patterns in the Lachish Region. Pp. 2672–90 in vol. 5 of The Renewed Archaeological Excavations at Lachish (1973–1994), ed. D. Ussishkin. Monograph Series 22. Tel Aviv: Institute of Archaeology, Tel Aviv University. Dothan, T. 1989 The Arrival of the Sea Peoples: Cultural Diversity in Early Age Canaan. Pp. 1–59 in Recent Excavations in Israel: Studies in Iron Age Archaeology, eds. S. Gitin and W. G. Dever. Annual of the American Schools of Oriental Research 49. Winona Lake, IN: Eisenbrauns. Dothan, T., and Gitin, S. 2008 Miqne, Tel (Ekron). Pp. 1952–58 in The New Encyclopedia of Archaeological Excavations in the Holy Land, 5 (Supplementary Volume), ed. E. Stern. Jerusalem: Israel Exploration Society/Biblical Archaeology Society Dothan, T., and Zukerman, A. 2004 A Preliminary Study of the Myecenaean IIIC:1 Pottery Assemblages from Tel MiqneEkron and Ashdod. Bulletin of the American Schools of Oriental Research 333: 1–54

Tel Beth-Shemesh and the Sorek Valley in the Iron Age

41

Fantalkin, A., and Finkelstein, I. 2006 The Sheshonq I Campaign and the 8th B.C.E. Earthquake:―More on the Archaeology and History of the South in the Iron I–II. Tel Aviv 33: 18–42. Faust, A. 2005 The Settlement on Jerusalem’s Western Hill and the City’s Status in the Iron Age II Revisited. Zeitschrift des deutschen Palästina-Vereins 121: 97–118. 2008 Settlement and Demography in Seventh Century Judah and the Extent and Intensity of Sennacherib’s Campaign. Palestine Exploration Quarterly 40: 168–94. 2012 Between Israel and Philistia: Ethnic Negotiations in the Iron Age I. Pp. 121–35 in The Ancient Near East in the 12th–10th Centuries B.C.E.: Culture and History. Proceedings of the International Conference Held at the University of Haifa, 2–5 May, 2010, ed. G. Galil, A. Gilboa, A. M. Maeir, and D. Kahn. Alter Orient und Altes Testament 392. Münster: Ugarit-Verlag. 2013a The Shephelah in the Iron Age: A New Look at the Settlement of Judah. Palestine Exploration Quarterly 145: 203–19. 2013b From Regional Power to Peaceful Neighbour: Philistia in the Iron I–II Transition. Israel Exploration Journal 63: 174–204. Faust, A., and Katz, H. 2011 Philistines, Israelites and Canaanites in the Southern Trough Valley during the Iron Age I. Ägypten und Levante 21: 231–47. Faust, A., and Lev-Tov, J. 2011 The Constitution of Philistine Identity: Ethnic Dynamics in Twelfth to Tenth Century Philistia. Oxford Journal of Archaeology 30:13–31. 2014 Philistia and the Philistines in the Iron Age I: Interaction, Ethnic Dynamics and Boundary Maintenance. HIPHIL Novum 1: 1–24. Finkelstein, I. 1995 The Date of the Settlement of the Philistines in Canaan. Tel Aviv 22: 213–39. 1996 The Philistine Countryside. Israel Exploration Journal 46: 225–42. 2000 The Philistine Settlements: When, Where and How Many? Pp. 159–80 in The Sea Peoples and Their World: A Reassessment, ed. E. D. Oren. University Museum Monograph 208; University Museum Symposium Series 11. Philadelphia: University Museum, University of Philadelphia. 2002 The Campaign of Shoshenq I to Palestine. A Guide to the 10th Century Polity. Zeitschrift des Deutschen Palästina Vereins 118: 109–35. Finkelstein, I., and Naʾaman, N. 2004 The Judahite Shephelah in the Late 8th and Early 7th Centuries B.C.E. Tel Aviv 31: 60–79. Gitin, S. 1989 Tel Miqne-Ekron: A Type-Site for the Inner Coastal Plain in the Iron Age II Period. Pp. 23–58 in Recent Excavations in Israel: Studies in Iron Age Archaeology, ed. S. Gitin and W. G. Dever. Annual of the American Schools of Oriental Research 49. Winona Lake, IN: Eisenbrauns. Gitin, S. 1995 Tel Miqne-Ekron in the 7th Century B.C.E.: The Impact of Economic Innovation and Foreign Cultural Influences on a Neo-Assyrian Vassal City-State. Pp. 61–79 in Recent Excavations in Israel: A View to the West, ed. S. Gitin. Archaeological Institute of America, Colloquia & Conference papers 1. Dubuque, IA: Kendall/Hunt.

42

Shlomo Bunimovitz and Zvi Lederman

1997 The Neo-Assyrian Empire and its Western Periphery: The Levant, with a Focus on Philistine Ekron. Pp. 77–103 in Assyria 1995, ed. S. Parpola and R. M.Whiting. Helsinki: Neo-Assyrian Text Corpus Project. Grant, E., and Wright, G. E. 1939 Ain Shems Excavations (Palestine). Part V (Text). Haverford: Haverford College. Herzog, Z., and Singer-Avitz, L. 2004 Redefining the Center: The Emergence of State in Judah. Tel Aviv 31: 209–44. Hitchcock, L. A., and A. M. Maeir. 2013 Beyond Creolization and Hybridity: Entangled and Transcultural Identities in Philistia. Archaeological Review from Cambridge 28: 43–65. Lederman, Z., and Bunimovitz, S. 2014 Canaanites, “Shephelites” and Those Who Will Become Judahites. Pp. 61–71 in New Studies in the Archaeology of Jerusalem, ed. G. D. Stiebel, A. Peleg-Barkat, D. Ben-Ami, and Y. Gadot. Collected Papers 8. Jerusalem: Tel Aviv University/Israel Antiquities Authority/The Hebrew University (Hebrew). Lev-Tov, J. 2006 The Faunal Remains: Animal Economy in the Iron Age I. Pp. 207–33 in Tel MiqneEkron Excavations 1995–1996: Field INE East Slope Iron Age I (Early Philistine Period), ed. M. W. Meehl, T. Dothan, and S. Gitin. Jerusalem: Albright Institute of Archaeological Research Mackenzie, D. 1911 Excavations at Ain Shems (Beth-Shemesh). Palestine Exploration Fund Annual 1: 41–94. 1913 Excavations at Ain Shems (Beth-Shemesh). Palestine Exploration Fund Annual 2. Mackenzie, D., Bunimovitz, S., Lederman, Z., and Momigliano, N. 2016 The Excavations of Beth-Shemmesh, November–December 1912. PEF Annual 13. London: Routledge. Maeir, A. M. 2015 Philistine Foodways. Pp.  402–04 in Archaeology of Food: An Encyclopedia, ed. K. B. Metheny and M. C. Beaudry. Lanham: Rowman & Littlefield. Maeir, A. M., Hitchcock, L. A., and Kolska Horwitz, L. 2013 On the Constitution and Transformation of Philistine Identity. Oxford Journal of Archaeology 32: 1–38. Mazar, A. 1980 Excavations at Tell Qasile, Part One: The Philistine Sanctuary: Architecture and Cult Objects. Qedem 12. Jerusalem: The Hebrew University of Jerusalem. 1994 The Northern Shephelah in the Iron Age: Some Issues in Biblical History and Archaeology. Pp.  247–67 in Scripture and Other Artifacts: Essays on the Bible and Archaeology in Honor of Philip. J. King, ed. M. D. Coogan, J. C. Exum, and L. E. Stager. Louisville: Westminster John Knox. 1997 Timnah (Tel Batash) I: Stratigraphy and Architecture, Text. Qedem 37. Jerusalem: The Hebrew University of Jerusalem. 1998 On the Appearance of Red Slip in the Iron Age I period in Israel. Pp. 368–78 in Mediterranean Peoples in Transition: Thirteenth to Early Tenth Centuries B.C.E.: In Honor of Professor Trude Dothan, ed. S. Gitin, A. Mazar, and E. Stern. Jerusalem: Israel Exploration Society. Mazar, A., and Panitz-Cohen, N. 2001 Timnah (Tel Batash) II: The Finds from the First Millennium B.C.E. Qedem 42. Jerusalem: The Hebrew University of Jerusalem.

Tel Beth-Shemesh and the Sorek Valley in the Iron Age

43

Naʾaman, N. 2003a In Search of Reality behind the Account of the Philistine Assault on Ahaz in the Book of Chronicles. Transeuphratène 26: 47–63. 2003b Ekron under the Assyrian and Egyptian Empires. Bulletin of the American Schools of Oriental Research 331: 81–91. 2010 Khirbet Qeiyafa in Context. Ugarit-Forschungen 42: 497–526. Oppenheim, A. L. 1969 Babylonian and Assyrian Historical Texts. Pp. 265–317 in Ancient Near Eastern Texts Relating to the Old Testament. 3rd ed. ed. J. B. Pritchard. Princeton: Princeton University Press. Panitz-Cohen, N. 2006 The Pottery of Strata XII–V at Tel Batash. Pp. 9–150 in Timnah (Tel Batash) III: The Finds from the Second Millennium B.C.E., ed. N. Panitz-Cohen and A. Mazar. Qedem 45. Jerusalem: The Hebrew University of Jerusalem. Panitz-Cohen, N. and Mazar, A., eds. 2006 Timnah (Tel Batash) III: The Finds from the Second Millennium B.C.E. Qedem 45. Jerusalem: The Hebrew University of Jerusalem. Rainey, A. F. 1983 The Biblical Shephelah of Judah. Bulletin of the American Schools of Oriental Research 251: 1–22. Rainey, A. F., and Notley, R. S. 2006 The Sacred Bridge: Carta’s Atlas of the Biblical World. Jerusalem: Carta. Schniedewind, W. M. 1998 The Geopolitical History of Philistine Gath. Bulletin of the American Schools of Oriental Research 309: 69–77. Shavit, A. 2000 Settlement Patterns in the Ayalon Valley in the Bronze and Iron Ages. Tel Aviv 27: 189–230. 2008 Settlement Patterns of Philistine City-States. Pp.  135–64 in Bene Israel: Studies in Archaeology of Israel and the Levant during the Bronze and Iron Ages in Honor of Israel Finkelstein, ed. A. Fantalkin and A. Yasur-Landau. Leiden: Brill. Singer, I. 1985 The Beginning of Philistine Settlement in Canaan and the Northern Boundary of Philistia. Tel Aviv 12: 109–22. Ussishkin, D. 2004 The Renewed Archaeological Excavations at Lachish (1973–1994). Monograph Series of the Institute of Archaeology 22. 5 volumes. Tel Aviv: Institute of Archaeology, Tel Aviv University. Vaughn, A. 1999 Theology, History, and Archaeology in the Chronicler’s Account of Hezekiah. Archaeology and Biblical Studies 4. Atlanta, GA: Scholars Press. 2016 lmlk and Official Seal Impressions. Pp. 480–501 in Tel Beth-Shemesh: A Border Community in Judah. Renewed Excavations 1990–2000: The Iron Age. Monograph Series 24. Tel Aviv: Institute of Archaeology, Tel Aviv University. Wright, G. E. 1976 Beth-Shemesh. Pp. 248–53 in Encyclopedia of Archaeological Excavations in the Holy Land I, ed. M. Avi-Yonah, Jerusalem: Israel Exploration Society and Massada. Yasur-Landau, A. 2010 The Philistines and Aegean Migration at the End of the Late Bronze Age. Cambridge: Cambridge University Press.

This page left intentionally blank.

Tel Burna: A Judahite Fortified Town in the Shephelah Itzhaq Shai

Tel Burna, located along the northern banks of Wadi Guvrin (fig. 1), is in the heart of the Judean Shephelah, one of Israel’s most intensively researched regions. Yet, until our current project began, the site was one of the few multi-period settlements in the region that remained unexcavated. As attested by a wide range of evidence (Egyptian, Assyrian, and Babylonian texts; biblical passages), this region served as a borderland in the Bronze and, especially, in the Iron Age, when Judahites and Philistines settled on opposite sides of the border. On account of this, when the current project at Tel Burna was first initiated, 1 one of the main research goals was to study border sites in ancient periods. After six seasons of field work, the preliminary results on the Iron Age remains at the site are presented here, discussing the material culture of the inhabitants and examining how it reflects the role of Tel Burna as a border site.

Biblical Identification 2 Whenever one investigates a site in this region, the site’s ancient identification cannot and should not be avoided. While more often than not we cannot be certain of a site’s ancient identity, considering the question opens the possibility of linking a ruin with a site mentioned in the texts available to us. Regardless of the problematic nature of the texts, they offer us another source of information on the past. Some scholars (e.g., Rainey 1983: 3; Naʾaman 2013) have claimed that Tel Burna should be identified with biblical Libnah, while others such as Tappy argue that Libnah should be placed elsewhere; Tappy (2008: 386) identifies Libnah with the nearby site of Tel Zayit. The biblical passages refer to Libnah as a Canaanite town that was conquered by Joshua, who then allotted it to the tribe of Judah ( Josh 10:29–30; 15:42). The 1.  The Tel Burna Excavation Project was initiated and directed by Dr. Joe Uziel and the author. As of 2013, the project is directed solely by the author, while Uziel continues to contribute his ideas, thoughts, and views in the analysis and publications of the finds. Funding for the project was provided by Ariel University and the Samaria & Jordan Rif R&D Center. 2.  For an in-depth study of the identification of the site, see Suriano, Shai, and Uziel (in press; cf. also McKinny and Dagan 2013).

45

46

Itzhaq Shai

Fig. 1.  Map of the Shephelah, with the location of Tel Burna.

city was then chosen as one of the Levitical cities ( Josh 21:13), which may point to its role as a border site (Rainey and Notley 2006: 127). According to 2 Kgs 8:22 (and 2 Chr 21:10), Libnah rebelled against Jehoram king of Judah in the mid-ninth century BCE) and was besieged by Sennacherib during the reign of Hezekiah in the late eighth century BCE (2 Kgs 19:8; Isa 37:8). Later, a woman from Libnah (Ḥamutal) married King Josiah in the seventh century BCE (2 Kgs 23:31–32; 2 Kgs 24:17–18; Jer 22:11), which may indicate the importance of the site along the border and the attempt by this king to create a political bond through marriage between the capital and its frontier. While there are other candidates for the location of ancient Libnah (including nearby Tel Zayit), the results of the archaeological project at Tel Burna support the identification of the site with biblical Libnah (see below).

Iron Age I Currently, no Iron Age I remains have been uncovered in our excavations. However, sherds of this period (including Philistine pottery; see Uziel and Shai 2010a: fig. 13:5) were collected during the surface survey, and a few were discovered in the excavations, but without a stratified context. The estimated settled area of the site during this period is 2 hectares (Uziel and Shai 2010a: 234, fig. 12). Therefore, it appears that, although the history of the Iron II settlement of Tel Burna correlates

Tel Burna: A Judahite Fortified Town in the Shephelah

47

Fig. 2.  Aerial view of the summit, looking southeast, with the remains of the fortifications.

with the Iron II sequence at nearby Lachish, the Iron Age I history seems different— but this conclusion must remain tentative until the details are clarified through excavation. It is important to note that the western slopes of the site were not covered with Iron Age I pottery, which may lead us to the conclusion that, while there was contact between the inhabitants of Tel Burna and the people of the coastal plain (i.e., the Philistines), as indicated by the presence of Philistine pottery, the inhabitants of Tel Burna preferred to live on the summit and the eastern and northern slopes. This preference may be related to their desire for security, since the western slopes faced the most powerful force of the time, the Philistines.

The Iron Age Casemate Wall The summit is defined by the distinct remains of fortification walls that have created a flat, almost square area of 70 × 70 meters (fig. 2). The fortification walls were already partially exposed along the perimeter of the upper tell, and the excavations thus far have revealed a segment of these walls in the northeast corner of the summit (fig. 3). The fortifications of Tel Burna were in use during the ninth and eighth centuries BCE. The discovery of a seventh century BCE silo that cuts the inner wall of the fortifications provides a terminus ante queme for the wall. Although the outer wall may have continued to function, the inner wall clearly went out of use by the seventh century BCE. The Iron Age II wall reflects the role of the site

48

Itzhaq Shai

Fig. 3.  Schematic plan of the Iron Age casemate wall.

during this period. 3 The material culture associated with the Iron Age II remains (see below) clearly indicates that this site was under Judah’s influence. Therefore, its location just between Gath (Tell es-Safi), the main Iron Age IIA city of the Philistines (e.g, Maeir 2012), and Lachish, the Judahite administration center in this region (Ussishkin 2004) explains why this casemate wall was constructed. The small summit that was enclosed by the wall was most probably used as a stronghold; on the one hand, it has a clear view toward the north (i.e., to Philistine Gath) and, on the other, it controlled the road running west–east through the Guvrin Valley. This supports the idea that the Kingdom of Judah was active in this region already in the Iron Age IIA (see also Naʾaman 2013: 254–55 regarding Libnah; Ussishkin 2014: 14–15)—not only after the destruction of Stratum A3 at Tell es-Safi/Gath (e.g., Fantalkin and Finkelstein 2006: 30–31; Koch 2012: 59–63).

Iron Age IIA To date, Iron IIA remains have been uncovered in both areas on the summit and its eastern slope (A2 and A1 respectively), exposing parts of a ninth-centuryBCE stratum (fig. 4). A surface (L21216) was exposed in Area A1 adjacent to a small installation (L21225) built of field-stones. On top of the installation, several loom weights were uncovered, alongside ninth-century-BCE pottery. Another surface was discovered east of the fortification wall; on it were smashed Iron Age IIA vessels, indicating that the settlement was not limited to the fortified summit during this phase. In Area A2, in the middle of the summit, the ninth-century-BCE remains were found in very small areas excavated beneath the eighth-century-BCE building (see below). Though not very much of this phase has been exposed yet, several points can be highlighted. 3. For a discussion of the casemate wall, its date, and function, see Shai et al. (2012: 141–57).

Tel Burna: A Judahite Fortified Town in the Shephelah

49

Fig. 4.  Plan of the Iron IIA remains in Area A1.

1. The summit was enclosed by the casemate fortification from at least the ninth century BCE (fig. 3; Shai et al. 2012: 154). 2. The settlement was not limited to the fortified summit but extended outside of the walls. This indicates that the site was not only a stronghold but was also settled beyond the fortified summit, which probably served as an administrative center, due to its location between Lachish, Azekah, and the nearby Philistine city of Gath. 3. Although the pottery assemblage is very limited, it seems much more like the pottery from Lachish Levels V–IV (Zimhoni 2004) than that from Tell es-Safi/Gath Stratum A3 (Shai and Maeir 2012), with certain types, more indicative of Philistine traditions, missing. For example, Late Philistine Decorated Ware (LPDW), which is considered to be a Philistine marker (see, e.g., Ben-Shlomo et al. 2004), is rare at Tel Burna. 4. The location of Tel Burna—midway between Gath, the dominant Philistine city in the Iron Age IIA, and Lachish, the main Judean city, monitoring the road along Nahal Guvrin, with visibility all the way to the coastal plain—would account for investment by the central authority of Judah in establishing a fortified town so close to the regional administrative center of Lachish.

Iron Age IIB Although the summit was fortified and enclosed by a casemate wall, the settlement was not limited to this plateau and was much larger, perhaps reaching its

50

Itzhaq Shai

peak in size during Iron IIB (ca. 8 hectares: Uziel and Shai 2010a: 238; see also Shai and Uziel 2014; Uziel et al. 2014). The main architectural phases dating to the Iron Age IIB (eighth century BCE) were discovered in both of the excavated areas on the summit, A1 and A2.

Area A1 As mentioned above, the casemate wall continued to be in use during this period (fig. 4). This is demonstrated by a surface (21210) that abuts the inner casemate wall (W12006). The pottery found on this surface includes wheel-burnished sherds and typical eighth-century-BCE forms (Shai et al. 2012: 148–49, fig. 10). Hence, it is clear that during the Iron Age IIB the summit was enclosed by a casemate wall.

Area A2 In the center of the tell (A2), a portion of a typical four-room house was uncovered (fig. 5). In several areas, the remains were damaged by later intrusions (mostly modern), but the plan of the building is clear. Its orientation is north–south/east– west and its size is ca. 12 × 15 m. Two massive monolithic pillars were discovered, and it is possible there were at least two more in the original plan of the building. The orientation of the three perpendicular rooms is east–west and the horizontal room is oriented north–south. Smashed vessels were discovered in the northwest corner of the northern room, buried under the accumulation of the destruction level. The vessels represent the ceramic horizon of the end of the eighth century BCE and, therefore, it is tempting to correlate these smashed vessels with Sennacherib’s campaign (701 BCE), although it is too early to determine if the destruction was due to local forces or something more elaborate. North of this building, there is a pavement (25404) made of flat field-stones in an area of 6 × 2 m. It is bordered by Building 32417 on the south and Wall 25406 on the north. The latter was built with a row of pillar bases and, most likely in a later phase, was filled with smaller field-stones in between the pillar bases. A door socket is located east of the wall leading to a beaten-earth floor (25405). Smashed complete vessels were found on the floor, as well as a concentration of about 30 loom-weights next to an installation made of 3 carved chalk stones placed on their narrow side and covered with plaster. This may have been part of a domestic installation used for textile production.

Iron Age IIC Above the eighth-century-BCE remains is the last Iron Age II stratum, which dates to the late Iron Age. This phase was found in Areas A1 and A2 and consists of silos and related architectural elements (fig. 6). The silos are lined with stone and in few a cases cut into the earlier remains, as is clearly demonstrated in Silo 15006, which used the earlier eighth-century-BCE pavement (25404) as its base. The silos

Tel Burna: A Judahite Fortified Town in the Shephelah

Fig. 5.  Plan of the Four-Room House.

51

52

Itzhaq Shai

Fig. 6.  Seventh-century-BCE silos.

yielded archaeobotanical remains that were recovered through the flotation of sediments. The diameter of most of the silos is ca. 1 m, yet one silo (32101) is exceptional: it is built of medium-sized field-stones, with a diameter of almost 2.5 m and a depth of 1.25 m. Archaeobotanical analysis of Silo 32101 yielded 16 different crop taxa and 32 wild plant taxa (Shai et al. 2014: 127–28, table 1). Fig seeds occur in the largest quantities. The second most abundant crop residue is barley, which is followed by linseed and wheat grains. The latter show characteristics both of free-threshing and emmer wheat grains but most probably represent a tetraploid wheat form. So far, chaff remains have not been discovered, except one rachis internode of barley. This implies that the silos were used for storage and that the residues found are not due to their use as refuse pits; in all likelihood, then, crop-processing did not take place in the immediate vicinity of the silo. The pottery assemblage includes types that are well attested in the late eighth century BCE, side by side with forms that are typical of the seventh century BCE (for a detailed description, see Shai et al. 2014). This is due to the nature of the sediment in the silo, which seems to have been backfilled after it ceased to be used at the end of the Iron Age or perhaps even later during the building activity that took place during the Persian period. Additionally, three stamped jar-handles were found in Silo 32101: one of the LMLK type, one of the Rosette type, and a private or official seal (see below and Shai et al. 2014).

Tel Burna: A Judahite Fortified Town in the Shephelah

53

Fig. 7.  Stamped handles.

The Finds Ceramic Assemblage The ceramic repertoire reflects the three main Iron Age strata: Iron Age IIA, IIB, and IIC. It should be noted that these three assemblages include the whole range of classes—storage, cooking, and serving vessels. The assemblage is typical of Iron Age II Judah, as implied by the LMLK and Rosette seal types, the cooking vessels, a lamp with raised disc-base, the decanter, the Judean folded-rim bowl, and so on. However, it is also important to note that a few of the cooking vessels are made of highly fired clay with silty and rounded sand quartz. The quartz grains look like beach sand found on the coast; nonetheless, this sand could have been added intentionally as temper in order to improve the quality of this vessel for use as a cooking pot; alternatively, perhaps the provenance of this vessel is coastal. The provenance study 4 revealed various groups of pottery representing sources in the Judean Shephelah but possibly also on the neighboring coastal plains and central hills.

Stamped Handles Several stamped handles of the common Judean style were found, including a LMLK type, 5 a Rosette type, and a private or official seal bearing the names ‫חגי‬/‫( לעזר‬fig. 7). The letters are consistent with the standard forms seen in the 4.  The petrographic analysis was conducted by Dr. David Ben-Shlomo and supported by the Samaria and the Jordan Rift Regional R&D Center. 5.  For a recent discussion and assessment of the function and dating of the LMLK seal impressions, see Lipschits et al. 2010; 2011; cf. also Ussishkin 2011; 2012; Lipschits 2012.

54

Itzhaq Shai

Fig. 8.  Judean Pillar Figurine.

eighth and seventh centuries Hebrew seal impressions. This set of names, ʿĒzer and Ḥaggî, appears on jar handles from two other sites in the Shephelah: Azekah and Gezer. While these other two may be stamped with the same seal, the stamped handle from Tel Burna is somewhat different. The fact that these seals come from the same area (the Shephelah) and time period (Iron Age II) suggests that they belong to a single person by the name of Ezer. The name Ezer occurs as a patronymic for a name that appears on various seals found at Beth-Shemesh and Lachish, which read “belonging to Ṣāpôn [son of] ʿĒzer” (Barkay and Vaughn 1996: 42–44). The occurrence of this name in multiple seals and as a patronymic on similar seals may indicate that this person held an important role within the kinship-based social network of Iron Age Judah. The presence of materials relating to the Judahite administrative system dating to both the Iron Age IIB and IIC demonstrates beyond any doubt that the site was controlled by the Kingdom of Judah during this period.

Judean Pillar Figurines Several Judean pillar figurines were found; the best-preserved (fig. 8) example has a face formed in a mold. This technique is common in the Iron Age II and can be found, for example, in Jerusalem (e.g., Darby 2011: 304–7; 2014; Ben Shlomo and Darby 2014). As noted by Kletter (1996), the Judean Pillar Figurines are typical of Judahite material culture, and their distribution clearly defines the border of the kingdom (see also Byrne 2004). It has also been shown that Judean Pillar Figurines usually were locally made, at each site (see, for example, Kletter 1999; Darby 2011: 308; Ben Shlomo and Darby 2014). Furthermore, the presence of several such

Tel Burna: A Judahite Fortified Town in the Shephelah

55

figurines at Tel Burna suggests that the inhabitants of the site shared ritual practices or beliefs with other Judahite centers.

Loom Weights As mentioned above, the remains of textile production facilities were exposed in two locations. The first was in Area A1 and is assigned to the Iron Age IIA. The second was discovered in Area A2 and is associated with Iron Age IIB. In addition to the loom-weight concentrations, these contexts also include the remains of installations. The earlier installation (L21225) was built of small field-stones that defined the area of the loom-weights. It is worth noting that installation L21225 abuts the inner casemate wall, which supports the conclusion that the fortification should be dated at least as early as the ninth century BCE and that domestic daily life activity took place in the area enclosed by the wall. The second (and later) installation (L42308) included about 30 loom-weights and 3 chalk stones covered with plaster. The purpose of the latter seems to have been to serve as a drain for liquid that may have been used in textile production. If this is the case, then we may have evidence of industrial or at least non-domestic textile production, but it is too early to be certain that this conclusion is correct.

Discussion Thus far, the excavations at Tel Burna have revealed evidence of an Iron Age II Judahite town. The town was established at least by the ninth century BCE, and it was during this period that the summit was enclosed by a casemate wall. It seems that the main reason behind this construction was the strategic location of the site between the main Iron IIA Philistine city in the region, Tell es-Safi/Gath, and the main Judahite administrative center in the Shephelah, Lachish. Accordingly, even though the settlement was limited to the summit, the town could have easily served as a stronghold with a clear view toward Philistine Gath and to the west over the coastal plain. 6 The large four-room house located on the summit is well-dated to the eighth century BCE and, along with the presence of the LMLK and “private” stamped handles, verifies that Tel Burna was at this time part of Judah both culturally and politically. Moreover, the stamped handle that bears the names ʿĒzer and Ḥaggî, which are also found at other sites in the Shepehlah (Gezer and Azekah), probably means that this person was an officer of the administration who was active in the region. The site continued to be settled in the last century of the Kingdom of Judah, as attested by the series of silos and some other architectural elements. This supports the suggested identification of the site with biblical Libnah, because, according

6.  For a suggestion about the line of Judahite fortresses including Tel Hesi and Tel Erani, see Master et al. 2014.

56

Itzhaq Shai

to the biblical testimony, Josiah’s wife came from Libnah (e.g., 2  Kgs 23:31–32; 24:17–18; Jer 22:11; see Suriano et al. in press). The Iron Age sequence at Tel Burna strengthens the idea that it was a border site located on the western frontier of the Kingdom of Judah. The long life of the casemate fortification wall (at least 200 years) supports this conclusion; it is significant for understanding the site’s role during the biblical period and for identifying the biblical name of the city. Stable carbon isotope analysis on ancient cereal grains has recently developed into an independent tool that permits the analysis of whether water stress played a role during the grain-filling period of the plant in antiquity (e.g., Araus et al. 1997; Fraser et al. 2013; Riehl and Shai 2015). Although low δ13C values in arid and semiarid environments indicate water stress, high δ13C values may be due to naturally available moisture, irrigation, or other unknown factors (Fraser et al. 2013). Stable carbon isotope data from Tel Burna indicate optimal soil moisture availability in the surroundings of the settlement, which actually increased in the Iron Age IIC when compared to the previous Iron Age IIB (Riehl and Shai 2015). These favorable conditions enabled profitable cultivation even of crop species with high water requirements, as indicated by the archaeobotanical remains from storage facilities. Assuming that the archaeobotanical remains at Tel Burna represent local crop production, they indicate relatively favorable growing conditions during the Iron Age II (Riehl and Shai 2015: 524–25). The typical Mediterranean set of crop species—olive, grape, and fig—is well-represented, accompanied by wheat, barley, pulses, and linseed (Shai et al. 2014; Riehl and Shai 2015). Cereals usually represent the dominant proportion of seed finds in agricultural sites not involved in trade. At Tel Burna, cereals account for only slightly more than 50% of the seeds. The crop species at Tel Burna occur in relatively equal amounts, although fig seeds may be somewhat overrepresented at the site due to their numerous seeds per distribution unit, indicating a broad-spectrum economy rather than highly specialized monocropping. We recognize neither a focus on a specific crop plant, which would be expected with surplus production conveyed to a regional or supra-regional trade network, nor do we see environmental fluctuation toward poorer conditions in the time from Iron Ages IIB to IIC, which could have led to an abandonment of individual crop species. Any indication of highly specialized agriculture or even a collapsing economy at Tel Burna is, therefore, lacking. In particular, the finds of linseed (Riehl and Shai 2015: 530–31) provide an illuminating insight into crop-production patterns. The comparatively large number of flax seeds and a ubiquity of 100% at Tel Burna make it clear that the plant was an important crop species at this location. The presence of oil-rich linseed may indicate that it was a preferred ancient product for consumption, for seed for sowing, its use as an oil crop, or for growing flax for linen production. The two varieties, flaxseed and linseed, differ slightly in their water requirements, but they are both drought-susceptible and require at least 400 mm of annual precipitation to thrive.

Tel Burna: A Judahite Fortified Town in the Shephelah

57

Seed flax requires approximately three times more available water than wheat or barley. As a general tendency, it can be noted that warmer and drier climate conditions stimulate seed growth in flax, whereas moister and cooler conditions foster length growth in the fibers. Flax is mentioned in the Bible on several occasions but always in relation to textile production and not as an oil crop, which would support the argument that agro-ecological conditions were sufficiently moist. The critical water requirement period for flax is from flowering to just prior to seed ripening, which takes approximately 3 months. In the course of the growing season, crop water use may be as high as 410 mm. Thus, flax cultivation at Tel Burna would have been possible without additional irrigation. Even if irrigation would have been necessary to buffer against inter-annual rainfall variability, the fact that linseed is ubiquitously represented at the site indicates that water availability should not have been a major problem during Iron Ages IIB and IIC (Riehl and Shai 2015: 532). When we set out to study the site, one of the primary research questions that interested us was the role of borders in the southern Levant during the Bronze and Iron Ages (Uziel and Shai 2010b). The results of the excavations thus far clearly point to Tel Burna being a Judahite site. This conclusion is supported by the architecture, as seen in both the four-room house (e.g., Bunimovitz and Faust 2003; Faust 2012: 213–29; and see Maeir’s [2013] critique) and the casemate fortifications (e.g., Shai et al. 2012 and additional references there), the pottery assemblage, the administrative system as reflected in the presence of the stamped jar handles, and the pillar figurines. This is particularly remarkable in light of the results of the excavations at Tell es-Safi/Gath, where the excavations yielded only a few pillar figurines in a large exposure of Iron Age IIB strata (e.g., Dagan 2014). Furthermore, stamped jar handles also were found at Tell es-Safi/Gath. Thus, it seems that it, too, site was part of the Judahite administration system in the late eighth century BCE (e.g., Maeir 2012). As such, while the Judahite affiliation of the inhabitants of Tel Burna in the eighth century BCE is clear, one may wonder who the inhabitants of Gath during the eighth century BCE were. Although it is possible that the presence of seventh-century-BCE remains at Tel Burna, in contrast with the lack of such remains at Gath (e.g., Maeir 2012; although note two Rosette handles found in the survey: Uziel and Maeir 2005), has contributed to our sense that Judahite cultural attributes are more common at Tel Burna. Nonetheless, further study and comparison of both assemblages, as well as expansion of the excavations of the Iron Age II strata from both sites, may lead to interesting assessments concerning the populations at both sites and the effects of political control and proximity to borders on local material culture.

58

Itzhaq Shai

References Araus, J. L., Febrero, A., Buxo, R., Camalich, M. D., Martin, D., Molina, F., Rodriguez-Ariza, M. O., Romagosa, I. 1997 Changes in Carbon Isotope Discrimination in Grain Cereals from Different Regions of the western Mediterranean Basin during the Past Seven Millenia: Palaeoenvironmental Evidence of a Differential Change in Aridity during the late Holocene. Global Change Biology 3: 107–18. Barkay, G., and A. G. Vaughn 1996 New Readings of Hezekian Official Seal Impressions. Bulletin of the American Schools of Oriental Research 304: 29–54. Ben-Shlomo, D., and E. D. Darby 2014 A Study of the Production of Iron Age Clay Figurines from Jerusalem. Tel Aviv 41: 180–204. Blakely, J. A., J. W. Hardin, and D. M. Master 2014 The Southwestern Border of Judah in the Ninth and Eighth Centuries B.C.E. Pp.  295–308 in Material Culture Matters: Essays on the Archaeology of the Southern Levant in Honor of Seymour Gitin, ed. J. R. Spencer, A. J. Brody, and R. A. Mullins. Winona Lake, IN: Eisenbrauns. Bunimovitz, S., and Faust, A. 2003 Building Identity: The Four-Room House and the Israelite Mind. Pp. 63–74 in Symbiosis, Symbolism, and the Power of the Past: Canaan, Ancient Israel, and Their Neighbors from the Late Bronze Age through Roman Palaestin, ed. W. G. Dever and S. Gitin. Winona Lake, IN: Eisenbrauns. Byrne, R. 2004 Lie Back and Think of Judah: The Reproductive Politics of Pillar Figurines. Near Eastern Archaeology 67: 137–51. Dagan, A. 2014 Between Judah and Philistia in the 8th Century BCE: The Material Culture of Tell es-Safi/ Gath as a Test Case for Political and Cultural Change. Unpublished Dissertation, Bar Ilan University, Ramat Gan. Darby, E. D. 2011 Interpreting Judean Pillar Figurines: Gender and Empire in Judean Apotropaic Ritual. Unpublished Ph.D. Dissertation, Duke, Durham. 2014 Interpreting Judean Pillar Figurines: Gender and Empire in Judean Apotropaic Ritual. Tübingen: Mohr Siebeck. Fantalkin, A. and Finkelstein, I. 2006 The Sheshonq I Campaign and the 8th Century BCE Earthquake: More on the Archaeology and History of the South in the Iron I–II. Tel Aviv 33:18–42. Fraser, R. A., Bogaard, A., Charles, M., Styring, A. K., Wallace, M., Jones, G., Ditchfield, P., Heaton, T. H. E. 2013 Assessing Natural Variation and the Effects of Charring, Burial and Pre-treatment on the Stable Carbon and Nitrogen Isotope Values of Archaeobotanical Cereals and Pulses. Journal of Archaeological Science 40: 4754–66. Kletter, R. 1996 The Judean Pillar-Figurines and the Archaeology of Asherah. British Archaeological Reports International Series 636. Oxford: Tempus Reparatum. 1999 Pots and Polities: Material Remains of Late Iron Age Judah in Relation to Its Political Borders. Bulletin of the American Schools of Oriental Research 314: 19–54.

Tel Burna: A Judahite Fortified Town in the Shephelah

59

Koch, I. 2012 The Geopolitical Organization of the Judean Shephelah during Iron Age I–IIA. Cathedra 143: 45–64 (in Hebrew). Lipschits, O. 2012 Archaeological Facts, Historical Speculations and the Date of the LMLK Storage Jars: A Rejoinder to David Ussishkin. Journal of Hebrew Scriptures 12. Lipschits, O., O. Sergi, and I. Koch 2010 Royal Judahite Jar Handles: Reconsidering the Chronology of the LMLK Stamp Impressions. Tel Aviv 37: 3–32. 2011 Judahite Stamped and Incised Jar Handles: A Tool for Studying the History of Late Monarchic Judah. Tel Aviv 38: 5–41. Maeir, A. M. 2012 The Tell es-Safi/Gath Archaeological Project 1996–2010: Introduction, Overview and Synopsis of Results. Pp. 1–88 in Tell es-Safi/Gath I: Report on the 1996–2005 Seasons, ed. A. M. Maeir. Ägypten und Altes Testament 69. Harrassowitz: Wiesbaden. 2013 Review of Avraham Faust The Archaeology of Israelite Society in Iron Age II. Review of Biblical Literature 9. McKinny, C., and A. Dagan 2013 The Explorations of Tel Burna. Palestine Exploration Quarterly 145: 294–305. Naʾaman, N. 2013 The Kingdom of Judah in the 9th Century BCE: Text Analysis versus Archaeological Research. Tel Aviv 40: 247–76. Rainey, A. F. 1983 The Biblical Shephelah of Judah. Bulletin of the American Schools of Oriental Research 251: 1–22. Rainey, A. F., and S. Notley 2006 The Sacred Bridge: Carta’s Atlas of the Biblical World. Jerusalem: Carta. Riehl, S., and Shai, I. 2014 Supra-regional Trade Networks and the Economic Potential of Iron Age II sites in the Southern Levant. Journal of Archaeological Science: Reports 3: 525–33. Shai, I., D. Cassuto, A. Dagan, and J. Uziel 2012 The Fortifications at Tel Burna: Date, Function and Meaning. Israel Exploration Journal 62: 141–57. Shai, I., A. Dagan, S. Riehl, A. Orendi, J. Uziel, and M. Suriano 2014 A Private Stamped Seal Handle from Tel Burna, Israel. Zeitschrift des Deutschen Palästina-Vereins 130: 121–37. Shai, I., and A. M. Maeir 2012 The Iron Age IIA Pottery Assemblage from Stratum A3. Pp.  313–63 in Tell es-Safi/ Gath I: Report on the 1996–2005 Seasons, ed. A. M. Maeir. Ägypten und Altes Testament 69. Wiesbaden: Harrassowitz. Shai, I., and J. Uziel 2014 Addressing Survey Methodology in the Southern Levant: Applying Different Methods for the Survey of Tel Burna, Israel. Israel Exploration Journal 64: 172–90. Suriano, M., I. Shai, and J. Uziel in press  In Search of Libnah. Journal of Near Eastern Society. Tappy, R. E. 2008 Historical and Geographical Notes on the“ Lowland Districts” of Judah in Joshua xv 33–47. Vetus Testamentum 58: 381–403.

60

Itzhaq Shai

Ussishkin, D. 2004 A Synopsis of the Stratigraphical, Chronological and Historical Issues. Pp. 50–119 in The Renewed Archaeological Excavations at Lachish (1973–1994), ed. D. Ussishkin. Monograph Series of the Institute of Archaeology of Tel Aviv University 22. Tel Aviv: Tel Aviv University. 2011 The Dating of the lmlk Storage Jars and Its Implications: Rejoinder to Lipschits, Sergi, and Koch. Tel Aviv 38: 220–40. 2012 LMLK Seal Impressions Once Again: A Second Rejoinder to Oded Lipschits. Antiguo Oriente 10: 1–13. 2014 Gath, Lachish and Jerusalem in the 9th Century BCE: An Archaeologist’s Perspective. Pp. 7–33 in New Studies on Jerusalem, vol. 20, ed. E. Baruch, A. Levy-Reifer, and A. Faust. Ramat-Gan: The Ingeborg Rennert Center for Jerusalem Studies, Bar-Ilan University. Uziel, J., and Maeir, A.M. 2005 Scratching the Surface at Gath: Implications of the Tell es-Safi/Gath Surface Survey. Tel Aviv 32: 50–75 Uziel, J., and I. Shai 2010a The Settlement History of Tel Burna: Results of the Surface Survey. Tel Aviv 37(2): 227–45. 2010b How to Choose a Site for Excavation: The Tel Burna Example. Biblical Archaeological Review 36: 28. Uziel, J., I. Shai, and D. Cassuto 2014 The Ups and Downs of Settlement Patterns: Why Sites Fluctuate. Pp.  295–308 in Material Culture Matters: Essays on the Archaeology of the Southern Levant in Honor of Seymour Gitin, ed. J. R. Spencer, R. A. Mullins, and A. Brody. Winona Lake, IN: Eisenbrauns. Zimhoni, O. 2004 The Pottery of Levels V and IV and its Archaeological and Chronological Implications. Pp. 1643–710 in The Renewed Archaeological Excavations at Lachish (1973– 1994), ed. by D. Ussishkin, Vol. 4. Monograph Series of the Institute of Archaeology of Tel Aviv University 22. Tel Aviv: Tel Aviv University.

Tel Gezer Excavations 2006–2015: The Transformation of a Border City Steven M. Ortiz and Samuel R. Wolff

Introduction The Tel Gezer Excavation project is a long-term joint project addressing chronological reevaluations, ethnic and social boundaries, and state formation in the southern Levant. As of 2015, the project has conducted eight summer field seasons. 1 The excavations are directed by Steven M. Ortiz of the Tandy Institute for Archaeology at Southwestern Baptist Theological Seminary and Samuel Wolff of the Israel Antiquities Authority. The excavations are sponsored by the Tandy Institute for Archaeology at Southwestern Baptist Theological Seminary. 2 The excavations are carried out within the Tel Gezer National Park and benefit from the cooperation of the National Parks Authority. The excavation project has an excellent relation with the local communities: Kibbutz Gezer and the Karmei Yosef Community Association. The Project is affiliated with the American Schools of Oriental Research.

Research Goals/Objectives The purpose of the project is to investigate state formation and regional boundaries in the northern Shephelah by investigating the Iron Age cultural horizon at Tel Gezer. These broad research trends in Iron Age archaeology are being addressed 1.  Gezer Survey: A survey project under the direction of Eric Mitchell was initiated the second season of the main project (2007). The goal of the survey is to survey comprehensively and systematically the region surrounding Tel Gezer and to locate and publish all archaeological features therein. The survey will complement the low intensity survey of the Aijalon Valley conducted by A. Shavit (2000). Gezer Water System: As part of the cultural heritage plan of the NPA and the hydrological research of Tsvika Tsuk, a joint project between the NPA and New Orleans Baptist Theological Seminary was initiated in 2010. Both of these two new projects are independent from the main excavations, although all three projects have a symbiotic working relationship and research designs. 2.  The project also received financial support from a consortium of institutions: Ashland Theological Seminary, College of the Ozarks, Emmaus Bible College, Lycoming College, Marian Eakins Archaeological Museum, and Lancaster Bible College and Graduate School. Other consortium members include Andrews University (2013), Clear Creek Baptist Bible College, Midwestern Baptist Theological Seminary (2006–2013).

61

62

Steven M. Ortiz and Samuel R. Wolff

by current research projects in the Shephelah and Southern Coastal Plain—specifically, ethnic and political boundaries in the Judean Hills and the Philistine coastal plain. Tel Gezer is an ideal site to address the regional geopolitical dynamic between Judah and Philistia during the Iron Age. Gezer was an important site in the history of ancient Palestine. It is located on one of the most important crossroads, and the ancient city is mentioned in several historical texts. Although previous excavations have revealed much of Gezer’s history, many questions that are key to the reconstruction of ancient Palestine remain unresolved.

Major Results Major results to date include (1) verification of the extension of the Middle Bronze glacis on the eastern slope of the western hill; (2) partial exposure of a Late Bronze Age building that was destroyed at the end of LB IIA (14th c. BCE); (3) discovery of an Iron Age I city wall, with a complex of several structures built up against the wall (11–10th c. BCE); (4) excavation of an Iron Age II Palace or Administrative Central Courtyard Building Complex; 3 (5) several 9th-century domestic units, destroyed in a violent conflagration tentatively associated with the campaigns of Hazael, built up against the reused casemate city wall; (6) an 8th-century administrative quarter that includes three large administrative and industrial public buildings; (7) a large four-room house (elite) destroyed by Tiglath-Pileser (734 BCE); and (8) additional units of a Hellenistic building complex that extends the Hellenistic exposure previously excavated by the Hebrew Union College Expedition. This article will focus on the Iron Age occupation levels. The complete stratigraphic profile and horizontal plans were only completed in the 2015 summer excavation season. Hence, this essay is a newer analysis of the excavations than the material presented at the 2014 World Jewish Congress. Because the ceramic reconstruction is currently being done in our lab, most of the discussion will focus on the stratigraphy and architecture.

History of Gezer Excavations The first intensive exploration of Tel Gezer was conducted by R. A. S. Macalister during the years 1902–1905 and 1907–1909, under the auspices of the Palestine Exploration Fund (PEF). 4 The results of these early excavations were published in three volumes (The Excavation of Gezer, 1912). 5 Macalister excavated nearly 60% of the tel. Unfortunately, the methods of excavation were very primitive: Macalister dug the site in strips and backfilled each trench. He distinguished eight levels of 3.  This building was partially exposed by the 1984 HUC excavations and called “Palace 10000” (Dever 1985). 4.  For a recent overview of the work of Macalister, see S. Wolff (ed.), Villain or Visionary? R.A.S. Macalister and the Archaeology of Palestine (PEF Annual 12, 2015). 5. Robert Alexander Stewart Macalister, The Excavation of Gezer 1902–1905 and 1907– 1909, Volumes I–III (London: John Murray, 1911, 1912).

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

63

Fig. 1.  Tandy Excavations on the southern slope between HUC Fields III and VII.

occupation. There were two smaller projects by Raymond Weill (1912–14, 1923–24) and Alan Rowe (1934). 6 The American Gezer Project began in 1964 under the auspices of the Hebrew Union College–Jewish Institute of Religion and the Harvard Semitic Museum, with Nelson Glueck and G. Ernest Wright as advisers. William Dever led the Phase I excavations (1964–1971), while Phase II was led by Joe D. Seger (1972–1974). These excavations distinguished 26 stratigraphic levels, from the Late Chalcolithic to the Roman period. The renewed project is not associated with the previous HUC excavations. Nevertheless, the project is cooperating and is developing a working 6.  Raymond-Charles Weill was known for his excavations in Jerusalem before and after World War I (1913–14 and 1923–24) under the patronage of Baron Rothschild. Sometime during the course of the Jerusalem excavations, Weill excavated lands on and around Tel Gezer that were acquired by Baron Rothschild. Not much was reported on these excavations until a publication by Aren Maeir (2004b). Recently, a previously unpublished monograph by Paule Zelwer-Silberberg was published posthumously on-line. In 1934, renewed excavations were conducted under the direction of Alan Rowe, under the auspices of the Palestine Exploration Society. This excavation was terminated after a short season. Only preliminary reports were produced, but the data from the excavation is available at the offices of the Palestine Exploration Fund.

64

Steven M. Ortiz and Samuel R. Wolff

Fig. 2.  Overview of Excavation Fields (2014).

relationship with the publication phase of the HUC excavations currently under the direction of Joe Seger of the Cobb Institute of Archaeology. Two smaller projects were conducted by Dever during 1984 and 1990 (Dever 1984, 1986, 1993) to address criticisms of various conclusions of the Hebrew Union College excavations, notably the dating of the outer wall. During the summer of 2005, the project co-directors met with Dever and examined the excavation notes and material culture remains located in the storerooms of the Nelson Glueck School of Biblical Archaeology of Hebrew Union College, Jerusalem. This initial review of the 1984 and 1990 excavation results provided the framework for the development of the research design.

Field Strategy The excavations are located on the south within the saddle between the western and eastern hills, west of the so-called “Solomonic” or six-chambered gate (Macalister’s “Maccabean Castle,” Hebrew Union College’s Field III) and east of HUC’s Fields VII and X. The excavations are designed to unite the Iron Age architectural elements and cultural horizons of Field VII and Field III of the HUC excavations with our renewed excavations, thus allowing for optimal reconstruction of the

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

65

growth and expansion of the Iron Age city as well as clear understanding of artifact distribution patterns.

Field E Field E encompasses an area west of the Iron Age Gate Complex (Field III of the HUC excavations). The goals of this area are to investigate the urbanization process of the Iron Age City. This field was our Field A from 2006–2009. The goal in this area is to investigate the urbanization process of the Iron Age City. This field includes an east–west section of squares from the Iron Age gate to the west exposing the city fortification system and its relation to building activity built up against the city wall and an area north of the fortification wall, where a series of large public buildings are located. In 2011, we excavated several squares containing the 8th-century Administrative Buildings (A–C). Beneath the buildings we excavated a complete 9th-century domestic unit as well as a large administrative building of the Late Iron Age IIA.

Field W Field W is located east of HUC Field VII and continues downslope to the south (our sondage). The goals of this area were to define the domestic area and provide a stratigraphic overview from the HUC Field VII domestic quarter to the southern slope encompassing the fortifications and line of the city wall. Indirectly, one goal also was to investigate the fortification line west of the Iron Age six-chambered gate. Field W evolved as a combination of Field B and the Fortification line of Field A. Field B was opened during the second season in 2007. In 2009, the field expanded south and connected with the casemate wall line of Field A. In 2010, Field B and the Field A Sondage became united as Field W. Work on the sondage has stopped because we have already established a sequence of fortifications from the Middle Bronze Age to the Iron Age IIB.

Excavation Results Middle Bronze Age The project design is focused on the Iron Age strata; nevertheless, in the process of determining the nature of the fortifications and the investigation of the “outer wall,” part of the Middle Bronze Age fortifications were excavated in the sondage. The Middle Bronze Age is well defined by the HUC excavations. 7 The contributions of the Tandy Gezer Excavations are: (1) adding to our knowledge of the MB fortification line on the southern slope of the tell; (2) determining that the Late Bronze Age buildings were built over the MB fortifications; (3) and, discovering that the builders of the Iron Age fortifications were aware of the MB glacis and incorporated their glacis and fortification system into the existing MB fortifications. 7.  Seger 2013.

66

Steven M. Ortiz and Samuel R. Wolff Tel Gezer Master Stratigraphic Chart 2006–2015

Preliminary Strata

Field E (formerly A)

Field W (formerly A-sondage and B)

1

Topsoil, Modern Excavation Dumps

HUC dump

Trenches, rock piles HUC Dump

Bergheim Estate, Abu Shusheh, Macalister

Backfill

Backfill

2

Hellenistic

Wall 61023, two pottery kilns (41010 and 61058), Silo 61038, reused IA walls (?), HUC: reused gate

Domestic buildings (A4/5), several pits above Four Room House

3

4

Persian

5

Late Iron Age II IA IIc

isolated pillar and basin

Pit (A4)

Ceramic

Retaining wall (A4/5)

HUC Excavations

Strata IIA–C, III

IV

Ceramic, Dog burials, pits Room in Area sV/W 4/5, Silo (W2), Large Silo (Z6) wall stubs, pits

V

Destruction: 734/733 (Assyrian Destruction (Tiglath-Pileser III This phase represents minor building/construction modifications 6A

VI

Enlarging south wall of Four Room House Rebuild Industrial Building C IA IIb 9th–8th

6B

Public: Administrative Buildings A–B; Large building: wall & plaster floor (A5/B5); Rebuilt fortification walls, HUC: 4-chambered gate

Rebuild Industrial Building C (Oil Production?); Four Room House, courtyard, street HUC: domestic buildings (Field VII)

VI

Unit D (rebuild of Building 52136, enlarged and strengthened)

VII

VIII (Note: Fields VII [7a & B] and III [5 & 6] each had 2 phases)

Destruction 7

IA IIb (9th)

Domestic: Units A–C; Dog burials Destruction: Shishak

8

IA IIa Late 10th

HUCI: Rebuild of drain in 6-chambered Gate; Casemate 12 door filled in

Rebuild/Strengthen city plan and repair of City Wall — buttressing interior

IA IIa Mid-10th

Public: Central Courtyard Building Fortifications: Casemate city wall HUC: 6-chambered Gate (Field III)

Buildings 52136, 52057: larger walls plus a cobble floor and a tabun; Fortifications: Single-line City-wall and rebuild glacis 11163

Construction sub-structures for defenses 9

Crib walls connected to casemate (Z9) — construction phase of Casemate and Iron IIA city wall Crib walls connected to casemate (B9) — reuse of Stratum 9 city wall as substructure for new city wall — construction phase of Casemate and Iron IIA city wall

IA Ic 11th/10th

Unexcavated

Tabun 62067 Rectangular Building, Courtyard with Tabun 82007

IA Ib 12th /11th

Isolated walls beneath casemate

Destruction 10A 10B 11

Complex of building units, City wall (62005/21097/11133/11157) Glacis and curb Complex of building units, city wall

IA Ia/b (12th)

Complex of building units, city wall

X–IX (late 11th/early 10th Siamun Des.) XI (Phil) XIII–XII

Destruction 12 13

LB IIA MB IIC

Wall 11097 in D9 Ceramic

Patrician House

XVI

Walls and Glacis

XVIII

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

67

Fig. 3.  Aerial Field W (sondage) Middle Bronze glacis (west at top).

MB II Rampart and Glacis.  A stone glacis was uncovered at the lowest excavated levels of the slope of the tel (fig. 3). This stone mantle consists of small bouldersized stones and covers a rampart that extends south from the outer line of the Iron II fortification walls. These stones were exposed for a length of 10 m from east to west and may continue to the east and west of our excavation area. The glacis continues 9 m downslope from its upper limit to the lowest exposed levels. The glacis slopes at ca. 15 degrees but is steeper toward the east. Near the highest part of the glacis and integrated into its steepest portion is a section of a wall more than 6 m long, extending eastward into the balk. This wall is preserved to a height of three courses of boulders and cobble-sized stones; its width remains undetermined,

68

Steven M. Ortiz and Samuel R. Wolff

Fig. 4.  Middle Bronze Age stone glacis on dike and fill layer.

because it extends into the northern balk. The function of this wall is uncertain; it may have stabilized the glacis or served as the base for a tower that was contemporary with the glacis. The stone glacis was founded on a rampart composed of dike and fill layers of alternating plaster and soil (fig. 4). Only the uppermost two layers in this sequence were excavated. This type of rampart construction has already been discerned at Gezer, although no others had a stone glacis. The southern edge of the top plaster layer in the sequence (15–20 cm thick) meets the top stones of the glacis. The plaster extends east–west for at least 15 m, although its traces are more patchy toward the east, and then continues northward from the glacis for at least 3 m as a flat plaster “cap” to this fortification system. A short wall section built of two courses and three rows of small unhewn boulders, was uncovered above this plaster cap. Its function remains enigmatic. A small probe (1 × 1 m) into the sealed locus below the plastered cap yielded only MB II pottery, thus dating the entire structure to that period.

Late Bronze Age The exposure of the Late Bronze Age levels occurred in the course of several seasons, mostly as the excavators were studying and excavating the Iron Age

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

69

Fig. 5.  Late Bronze Age (Stratum 11) [in green].

fortifications. In the first season (2006), shallow fills with Late Bronze Age pottery were found on the slope beneath the line of the Iron Age II casemate wall. In a following season, a large pillar base was found in a probe conducted to determine the foundation level of the city wall. The debris associated with the pillar base had Late Bronze Age pottery. We tentatively proposed that these materials were the remnants of a square-pillared building and associated it with Stratum XIV of the HUC excavations solely because the stratum above it was Iron Age I. We postulated that this might have been associated with Merneptah’s destruction of Gezer (Ortiz and Wolff 2012: 12). 8 The Late Bronze Stratum became clearer with the chance exposure of this Late Bronze Age occupation during the 2013 season. While excavating the Iron Age I wall and removing the stones of the Iron Age glacis, a small exposure of the Late Bronze stratum was excavated. This is isolated on the southern edge of Field W (fig. 5). This chance exposure has revealed more components of the pillared building. The plan of the building is difficult to discern, because the southern part of the building eroded down the slope and an Iron Age I city wall and complex of buildings are built directly on top of this building (see fig. 5 above). The building complex (fig. 6) consists of north–south walls (52131, 62014, 62079) as well as east–west walls (62006, 52128, 11166, 72038). There is a single pillar base (31071), less than a meter in diameter, which probably is evidence of a 8.  In our previous excavation report (Ortiz and Wolff 2012), our stratigraphic chart contained a misprint: the LB stratum should refer to HUC Stratum XV, not Stratum XII.

70

Fig. 6.  Late Bronze Age building (Stratum 12).

Fig. 7.  Tentative plan of Late Bronze Age remains.

Steven M. Ortiz and Samuel R. Wolff

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

71

Fig. 8.  Stone installation 82047 (Late Bronze Age).

courtyard or central room. All the walls are 1 m thick and constructed of two rows of unhewn stones, with chinking stones. Evidence of a mud-brick superstructure was found above Wall 62014. Just south of the pillar base and proposed courtyard/ central room was Platform 62022. This could also be the remnants of a wall with cobbles. It looks like the remnants of a rectangular, single-course structure. This building contained Installation 82047—a limestone basin with a sump. The stone installation broke in half in antiquity, and the northern half has fallen over (see fig.  8). This installation would have been in the northeastern room of the Late Bronze Age building. Evidence of destruction was found south of Wall 52128, around the pillar base, on top of Surface 62008, and around stone Installation 82047. Most of this was a heavy layer of ash and fired mud-brick detritus. One complete storage jar was found above beaten earth Surface 52130. Most of the pottery and finds were found on Surface 62008 between Walls 52131, 62014, and 62006. Some of the finds included a roof roller and large grinders. Several vessels (cooking pot, krater, storage jars) were found in the destruction on this surface, as well as a scarab of Amenhotep III and three cylinder seals. Several fragments of Cypriote and Mycenaean pottery were found, all of which date to the 14th century BCE. This 14th-century BCE destruction matches other LB IIA destructions in the region (e.g., Beth Shemesh, TimnahBatash, and Jaffa). Based on these finds, the original proposal that this Late Bronze Age stratum equates HUC Stratum XIV was incorrect. It appears that the LB building is a square-shaped building, ca. 12 × 10 m. It is unclear if this is a public or private building. Based on the size and finds, it is probable that it is a typical Canaanite patrician courtyard house (like Tel Batash Building 475; Gilboa, Sharon, and Zorn 2014, see especially LB plans on p. 52, fig. 8). The question is whether the northern wall W52128 and eastern wall W62079

72

Steven M. Ortiz and Samuel R. Wolff

are exterior walls or are interior walls surrounding a central courtyard containing a pillar base and installation. If the latter is the case, then this building is perhaps a larger residence (see Shai et al. 2011 for discussion of the differentiation between public and private LB buildings). Previous publications and papers have noted that the Late Bronze Age stratum is found on the edge of the slope, with the southern extent eroded down the slope. It was built directly on the Middle Bronze Age glacis. Based on this data, we proposed that there was no LB city wall (at least in this area) and that the MB fortifications were not reused in the LB. Sometime in the Iron Age I, a city wall was built directly over the Late Bronze Age destruction and occupation. An Iron Age II glacis was built over the Iron Age I wall and provides evidence for the extent of the slope during this period. While our investigations into the Late Bronze Age remains are still in their initial stages, perhaps this destruction is indicative of the unrest between the Canaanite city-states as reflected in the el-Amarna correspondence. There is no evidence in our Field W of any stratum that is contemporary with the HUC Stratum XIV/XV associated with the Merneptah destruction. Perhaps the Late Bronze Age II occupation extended down the eastern slope of the western hill and, after the destruction, the city contracted to the top of the western hill. One of the auxiliary goals of the 2015 season was to connect the line of fortifications found in our excavations with the architectural features in HUC’s Field X. Excavation consisted mostly of removing dumps from previous excavations (including the Tandy, HUC, Macalister digs). While the results are tentative, it is clear that the original date given to W1003 as part of the Iron Age II casemate system is incorrect. This wall (removed by HUC; see fig. 10 and red outline in fig. 9) does not line up with the Iron II city wall line (blue in fig. 9) but does line up with the stones of the Iron I “curb and glacis” (62049 and 62048; fig. 9, light orange). Additionally, W1003 is lower in elevation than the Iron II system, and it rested on top of W1020 (fig 9, yellow), which runs beneath the Iron I city wall materials (orange in fig. 9). W1020 is thus tentatively dated to the Late Bronze Age and likely forms the southern wall of an LB building west of the one/s in Z9–W9. These conclusions are tentative, because we have only started to remove the surface soil and have not yet fully defined the relationships between these walls.

Iron Age I: Stratum 10 Perhaps one of the main results of our excavations is the discovery that the Iron Age I city of Gezer was a walled city and that it extended from the acropolis (western hill) to the center of the southern edge of the tel. More than 20 m of this city wall has been excavated. The city wall is constructed of large unhewn stones, in two courses, with chinking stones between these courses and is preserved in some places for a meter and a half in height. The Iron Age II city wall was built directly on top of this wall, and in some places it appeared to be integrated with this wall. Because of this integration, we originally interpreted this earlier wall as a retention

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

73

Fig. 9.  Iron Age I wall and earlier wall.

system of support walls for the Iron Age II wall (Ortiz and Wolff 2012). Once the the Iron Age II wall was dismantled, it became clear that these were two separate city walls. In fig. 11, the cutaway of the Iron Age II wall (shown in blue) and the Iron Age I wall are visible. Built up against the north face of the city wall is more than 150 m2 of a complex of buildings integrated into the city wall (fig. 12). The project has tentatively isolated nine building units. In almost all units, evidence of destruction was found (Units A, B, C, and 3 in fig. 12).

74

Steven M. Ortiz and Samuel R. Wolff

Fig. 10.  Plan of Iron Age I wall (Tandy) and Wall 1003 (HUC) [plan: G. Arbino].

Along the north face of the city wall were several units that were built as integral units with the wall, perhaps an earlier form of casemate fortification. We have defined five of these units. Each unit is basically a single room that probably served as storage. These units were built directly on the Late Bronze Age materials and remained unchanged throughout Iron Age I (Strata 11–10). Unit 1 is undefined. While we were expanding the excavations of the western end of the Iron Age I wall, a complete storage jar abutting the city wall was found in destruction debris. It is possible that this unit connects with Unit 2 to the east of it, but the debris is at a higher level then the surface level of Unit 2. Unit 2 is 5 × 2 m with a north–south divider wall (W52134). The unit had beaten earth surfaces and a bin in the corner of the western room (L52107). Units 3 and 4 are built directly on the Late Bronze Age destruction, with north–south walls preserved for ca. 2 m in several courses. Unit 3, 3 × 5 m, had extensive Iron Age I destruction debris: ash and burned mud brick on a floor (Surface 52116), where we found two storage jars, a multi-handled krater, and several mushroom-shaped clay stoppers. One of these stoppers had a stamp seal common to the Egyptian 21st Dynasty. Unit 3 appears to be a basement or lower storage area and also served as a support or leveling for Units A and C just north of this room. The founding levels of the walls of Unit 3 as well as the Iron Age destruction debris were directly on the Late Bronze Age occupation. Unit 4 was also built directly on top of the Late Bronze destruction, so much so that the builders were aware of the Late Bronze building and built their walls directly to the north and east to avoid the large vat of the LB building. Much of the Late Bronze debris

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

75

Fig. 11.  Iron I (red) and Iron II (blue) City Walls.

was used as backfill. The surface level of Unit 4 was not found, probably destroyed by the construction of the Iron Age II city wall, which was built on top of these units. Unit 5 was also heavily damaged by the construction of the Iron Age II city wall. Based on wall remnants, this room was 6 × 4 m. An almost complete storage jar was found against the southern wall (wall 11136) of the room, and remnants of a cobble surface (62080) provide a hint of the location of the surface level. On the other hand, Units 3 and 4 had to have deep foundations because of the slopes (west–east and north–south) at this point on the western hill. Each of these five units built up against the city wall follow the slope of the defensive wall, dropping more than a meter in height. For instance, the surface of

76

Fig. 12.  Iron Age I Building Complex.

Steven M. Ortiz and Samuel R. Wolff

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

77

Unit 5 is a meter below the surface of Unit 2, even though these two contemporary surfaces are only ca. 10 m apart. There are four units to the north. Unit B is a large unit about 6 × 3 m. The entrance to this unit was from the west (see fig. 12). There is a threshold that is flanked by two pillars. This unit contained a plaster surface that ran up to the threshold. Against the eastern wall was a bin (Installation 82036), where two restorable storage jars were found. In addition, a bronze needle and a scarab were also found near the southern wall (W52133). The main unit of this group is a Pillared Building: Unit A, which contains three pillars in an east–west line. 9 This large pillared room, ca. 8 × 4 m, was floored with a beaten earth surface (Surface 82015). The entrance to this room was from the north, between wall 82030 and wall 82067. The walls were destroyed at the entrance, but we could estimate that there was probably a 3-m-wide entrance into the pillared room. In the western corner of the room, between the western Pillar 82023 and Wall, we found three complete storage jars covered by a destruction of fired mud brick and stone tumble. In the south of this unit is a room-niche, ca. 2 × 2 m. The function of this niche is unknown; it faces Unit 3, where we had a lower basement room and we found storage jars and several clay stoppers. In addition to the restorable vessels, within the destruction we found several cult items, such as a miniature rattle, fragment of a six-toed zoomorphic vessel, a fragment of a vessel with a possible phallic symbol, as well as a bronze spearhead. Unit A opens to Unit C to the east separated by a wall (82016). Unit C is partially excavated (walls and architectural features of Stratum 9 still remain unexcavated in the area). It is about 7 m in length. Destruction debris in this room contained restorable vessels. Unit D is a small, 2 × 3-m room with a cobble surface that had a chalk overly in the south. It was also entered from the north. Foundations of the walls are above the surface levels of the rooms (from 3 cm to 8 cm). We are tentatively concluding that Unit D was a later addition, because the walls (52012 and 52043) were raised above the surface levels of Units A and B. One of the unique features of the Pillared Building (Unit A) and Unit D is that the walls are constructed of larger stones than the other units. They are large boulder-size stones that require two or three men to move or lift. The other walls of the Iron Age I units are constructed of stones that an average-sized person can carry. All the walls are constructed of a single row of stones. The HUC excavations uncovered two Iron Age I courtyard houses on the acropolis (Field VI). When the Field W Iron Age domestic units are compared to the two on the acropolis, it is clear that these are of a poorer quality. It is possible that the Iron Age Courtyard houses on the acropolis were elite compared to this quarter, found next to the southern city wall. The Iron Age I destruction and city wall were a surprise. Although we knew that the HUC excavations revealed Iron Age I 9.  From east to west: Pillars 82021, 82022, and 82023.

78

Steven M. Ortiz and Samuel R. Wolff

occupation, it was only found in Field VI on the acropolis, with minor ceramic evidence on the southern end of the tel. We now have evidence for at last two domestic quarters of the Iron Age City. Unlike the courtyard houses on the acropolis, we found minimal Philistine Bichrome pottery in the fills and debris of the Iron Age I.

IA I/IIA: Stratum 9 This phase was known but undiscerned in our 2014 season report. It was placed in our general Iron Age I Strata (9 and 10). It is now clear that what we originally thought was a unit/room of the Iron Age I building complexes is actually a later phase. This is a unique phase, because it is tentatively dated to the initial construction of the 10th-century stratum (Stratum 8), yet the builders are aware of the destroyed Stratum 10 phase. One wall (Wall 72028) reuses a pillar (Pillar 82031) of the Stratum 10 pillared building in its line, and a north–south wall (W31025) abuts this pillar. In addition, other walls are built directly on top of or integrate walls of Stratum 10 (e.g., W82046 is built on top of Stratum 10 W52145 and integrates W52144). The excavation of a balk (between Y7 and Z7) revealed several architectural elements that helped define this stratum: (1) a Tabun 82007; (2) cobble surface 82028; and (3) the continuation of W72028 (W82009), which abuts W31025. While the question remains if this stratum is contemporary with the Iron Age IIA city wall (e.g., it was rebuilt later and we do not have the original wall), there was a small rebuild of a domestic structure. Our new understanding of Stratum 9 is that it was an intermediate phase between the Iron Age I buildings of Stratum 10 and the major fortifications of Stratum 8. The HUC excavations also found several phases dating to the Iron Age I and II. No surfaces or debris layers can be associated with the several architectural units that are remnants of these phases. Most were destroyed by the robust foundation and buildings of the Hellenistic Period in Field W. These were domestic units. None of these units, with the exception of Unit E, were the typical courtyard house (Gilboa, Sharon, and Zorn 2014).

Iron Age II: Stratum 8 (10th Century BCE) Casemate Fortification City Wall The Iron Age casemate wall was previous excavated by the HUC and the 1984 University of Arizona Excavations. The wall was also previous known from the Macalister excavations. The Tandy excavations have explored the western extent of the line of the Iron Age city wall. The Iron Age fortification system has been well described in our previous report (Ortiz and Wolff 2012). One of the proposals in our 2012 report was that there was a subterranean wall system built beneath the casemate city wall. During the 2012 season, nearly 15 meters of the Iron Age II city wall was removed to: (1) investigate the fortification system; and (2) expand Field W to the south. The removal of this wall was fortunate: it clarified our understanding of the Iron Age fortification system. We now know (see above) that what we thought

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

79

Fig. 13.  Broad room Plan (Stratum 9).

was a subterranean support system was actually the remnants of the Iron Age I city wall and domestic structures. The casemate wall system consists of two parallel wall lines that extend 30 m to the west from the gate; the portion known to exist to the east of the gate was not

80

Steven M. Ortiz and Samuel R. Wolff

Fig. 14.  Casemate and glacis (Stratum 8).

excavated by our team. The casemate is constructed of two rows of large unhewn stones, with a central row of smaller chinking/fill stones. The northern line continues as a single wall for more than 17 m, whereas the southern wall line ceases. 10 The northern wall line was adapted into a latter 8th-century BCE rebuild that consists of a single row of stones.  11 It contains a reused vat. After the first field season, it became clear that Macalister had previously excavated this area and freely altered the plans. It is interesting to point out that there were no entrances into the casemates, as there were at other sites such as Kh. Qeiyafa (Garfinkel and Ganor 2009).

Wall Retention System We previously proposed (Ortiz and Wolff 2012) that there was a series of seven single-row walls abutting the southern face of the fortification wall. We also assumed that these retention walls were built on top of or were integrated into the Iron Age I city wall and the destruction debris from that wall. This retaining wall system is contemporary with the main casemate system. The narrowness of the socle system indicates that the large, three-row, double-wall casemate system did not continue west beyond our field of excavation. With the removal of part of the Iron II wall, it was discovered that some of what we thought were socle support walls were actually north–south walls of the Iron Age I city that are abutting the Iron Age I city wall. The Iron Age II builders found an already existing Iron Age I 10.  It is possible that the rebuild reflects the original design of the city wall (i.e., a double parallel wall (casemate) from the gate that becomes a single wall line); excavation of earlier strata will determine the history of use. 11.  According to Macalister’s plan, he found only a single wall line.

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

81

Figs. 15 and 16.  Casemate wall in saddle (Stratum 8) [plan by G. Arbino].

city wall with domestic units incorporated into the wall. They added some additional retaining walls, a leveling fill, and then built a new city wall on top of the earlier stratum.

Iron IIA Glacis The Iron Age stone glacis was founded directly atop Late Bronze/Early Iron I destruction debris. It extends from the western edge of excavations eastward almost 15 m and southward ca. 10 m, with a 1.3 m drop in slope. At the southern edge of this glacis, the stones ended in an uneven and erratic edge, which dropped off vertically 1.6 m to the level of the MB II glacis below, indicating that it was robbed out either in antiquity or by modern excavations (Macalister). The stones of this glacis were smaller than those of the MB II glacis. A stepped sloping revetment construction (glacis) was built up against the Iron Age city wall and incorporated into the Iron Age I city wall and socle wall retention system. This structure is built of cobble- to boulder-size unhewn stones in a series of layers or steps from the south up to the outside face of the city wall. It was exposed nearly 15 m from east to west and ca. 15 m in width. The extent of this stone

82

Steven M. Ortiz and Samuel R. Wolff

Fig. 17.  Iron Age IIA administrative central hall/courtyard building.

revetment is unknown because it is only revealed in a probe in the sondage. From the bottom of the revetment to the top of the fortification wall is a height of nearly 7 m. There is a 2½ m east–west slope from the surface levels north of the fortification wall and revetment to the threshold of the Iron Age II gate. The Iron Age fortification system constructed during the 10th century BCE was designed for optimal defense. The six-chambered gate was built in the saddle with the “built-up foundation technique,” 12 and the western fortification wall to the west followed the slope up the western hill. At the point of the western end of the saddle, remnants of the Iron Age I city wall with an additional stone revetment system was constructed. At a later period in the Iron Age, an outer gatehouse was constructed, along with an outer wall—typical of other Iron Age sites. 13

Administrative Quarter (Field E) In Field E, the area to the west of the six-chambered gate was an administrative quarter. This season, with the removal of all balks and the connection between the Tandy excavations and the previous HUC (1970s) and Arizona (1984) excavations, we now have a complete plan of a large administrative building (see fig. 17). This 12.  Ussishkin 1980. 13.  See Dever 1984.

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

83

Fig. 18.  Administrative central hall building: Rooms 1 (left) and Room 2 (right). Note the Ashlar masonry in the corners.

building is the typical bit hilani-type palace found mostly in the northern Levant (Syria). This building type has come under reevaluation: several scholars now question whether the term bit hilani is an appropriate label. Sharon and Zarzecki-Peleg posit that these buildings should be called Lateral-Access Podium (LAP) structures, while Lehmann and Killebrew prefer the term Iron Age Central Hall Tetra-Partite residencies. 14 Our Administrative Central Hall Building (ca. 20 m × 25 m) has features of these types of large central buildings. It is clear that our building, the first that has been found and completely excavated at Gezer, fits this Iron Age tradition of large administrative buildings (compare Megiddo Palace 10,000). 15 This administrative central hall building was built mostly with large, rough field-stones. What is unique is that the corners of the buildings have ashlar stones. In addition, there is a major wall built mostly of ashlar blocks that separates the two 14.  Sharon and Zarzecki-Peleg 2006; Lehmann and Killebrew 2010; See also Ben-Ami and Wazana 2013; Osborne 2012. 15. Note that Dever (1984) found parts of this building, which he identified as Palace 10000. He recognized that this was a large administrative/elite building, based on the construction (ashlar masonry) and partial plan; he did not recognize it as a typical bit hilani-type structure.

84

Steven M. Ortiz and Samuel R. Wolff

central halls (i.e., Room 6 and Rooms 7–9). These blocks were tipped over from the east to the west (see fig. 17). Dever described this wall (the “west wall” of his Palace 10000) as built of “massive, roughly squared blocks . . . [which] had tumbled over in disarray and remained where they fell.” 16 The eastern part of the administrative central hall building  17 and limited areas elsewhere within the building 18 were previously excavated by HUC as part of HUC’s Palace 10,000. 19 Most of these areas have remained exposed to the elements in the decades since their excavation and, although we have revisited some of these areas, they offer little new information apart from aiding in our understanding of the architecture of the larger structure. The central hall building appears to have two entrances. Area 13 20 is an entry corridor providing access to the building from the alley separating the public structure from the city gate. This same alley gave access to Areas 14 and 15, neither of which could be entered from inside the building, as well as the initial casemate. The walls in this area of the building are notably distinct in construction from the rest of the building. They are constructed two-tothree rows wide, with roughly dressed field-stones that are somewhat smaller in size than those utilized in the rest of the building. As is the case elsewhere in the public structure, these walls incorporate the occasional ashlar block. While this southeastern corner of the structure does appear to be part of the public structure, the thickness of the walls and the seemingly exterior entry to the rooms suggest a different use for this portion of the building. Perhaps this area was part of a defensive tower associated with the gate but incorporated into a larger support structure. Entering the central hall building via the entry corridor (Area 13) was by means of a bent-axis entrance. This entrance vestibule was paved with a cobbled surface (S81045). Interestingly, this was the only stone-paved surface; all other surfaces in the central courtyard building were plastered. The stone-lining was likely due to the increased amount of wear due to foot traffic in the eastern entrance (Area 10). Alternatively, the placement of the cobbles in the east entrance may have been due to factors related to the runoff of rain water from the plastered courtyard (area 6). In the absence of any drainage system in the courtyard (for which no evidence was found), it is possible that water may have drained to the area of the entry and that the cobbles were intended to ease movement of water through the entryway under such conditions. To the south of the east entrance are two small rooms that may have served as small storage spaces. The first room (Area 11) is separated from the entrance (Area 10) by a small partition wall (W81044). This partition is built of two rows of rough field-stones laid out in a square approximately 1 m × 1 m, abutting the east face of 16.  Dever 1984: 216. 17.  Areas 4, 13, 14, and 15 in fig. 17. 18.  Within areas 3, 5, 6, 7, 8, 10, and 11 in fig. 17. 19.  For preliminary discussions of HUC’s results regarding the public structure, see Dever 1984: 206–18; 1985: 217–30; 1986: 9–34; Younker 1991: 19–60. 20.  All unit and room numbers in this section refer to the designations in fig. 17.

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

85

wall W31066. The passage into this room (Area 11) is 0.65 m wide, while the room itself is about 1.4 m × 1.7 m. On the south side of the room is a nearly identical passage created by a partition wall (W81043) that leads into another room (Area 12). This room is of similar design to the previous but slightly larger at about 1.5  m × 2 m. The ceramic material recovered from these rooms primarily represents storage jars, strengthening the suggestion that these small rooms were utilized as storage spaces. To the west of these rooms were two elongated halls and/or courtyards. Entering through the bent axis of the entrance (Area 10) leads to a large central courtyard that was subdivided into various spaces (Areas 6–9). The northern end of the eastern courtyard (Areas 7 and 8) is divided by a short protrusion from a north–south wall W31066 that may be the sole remaining evidence of an originally more substantial divider. Just south of this area is Area 9. This area is bounded by the casemate wall W11081 on the south, vat L71084 and abutting wall stub W71076 on the west, wall W31066 on the east, and wall W81046 on the north. Entry into this approximately 4 m × 2 m space was via the northwest corner, next to vat L71084. It is unclear whether this area was open to the rest of the courtyard, perhaps with wall W81046 serving as a short partition wall, or was an enclosed room. The tops of the remaining stones of wall W81046 are at about the floor level of the courtyard, while most of the walls of the public structure are preserved much higher. Activities carried out in the area may have been related to the use of vat L71084, given its proximity, but there were no other finds within the area that might help to discern the use of the space. 21 The second elongated room (Area 6) provides the clearest evidence that this space was an open courtyard. A well-preserved plastered surface (S71080) was excavated in the course of the 2014 and 2015 seasons and was found to extend from the casemate wall (W11081) in the south to the entrance of room 2 in the north, a length of approximately 9 m. Constructed on this surface were three tabuns: L71055 in the south, against vat L71084; L81021 in the north, along the south face of wall W61064; and L81028 in the west, still partially obstructed by the west balk of E7. Also found in the courtyard during the 2014 season was a small concave stone feature (L71079) set into the plaster and surrounded by a ring of small cobbles. This stone feature may have been a grinding installation or perhaps a posthole for some sort of temporary covering. Evidence of destruction was found on the surface, where remnants of charred wooden beams were encountered. Small finds from the courtyard found in proximity to this installation include an Egyptianstyle faience Bastet bead and fragments of a bull figurine with a unique circular appliqué on the forehead. 21.  The vat might have been in secondary use as part of the wall line. It is not clear at this point if this represents its original use in the central courtyard building or if it was put out of use at a later time. One of the problems is that the stratigraphic relationship is lost due to the intrusion of a later stone pit (remnants are visible in fig. 17).

86

Steven M. Ortiz and Samuel R. Wolff

Separating Area 6 from Area 7 is an ashlar wall (W51088) that runs 2.60 m in a north–south orientation. As can be observed in the plan, this ashlar wall is in line with walls W71076 and W61065 but does not connect with either. There are two courses of ashlars preserved, and a number of other ashlars and large rough stones were found tumbled to the west of the wall, having clearly been part of the wall prior to its destruction. It appears that this wall acted as a partition between Areas 6 and 7, though its size may suggest an additional structural purpose that is not yet clear. The area immediately to the west of the central courtyard (Area 5) remains partially obscured by unexcavated balks and squares above the level of the courtyard. What is clear at this point is that Area 5 is separated from the central courtyard by north–south wall W51063. This wall abuts the casemate wall W11081 on the south, but its northern extent remains concealed within the unexcavated balk. Given that tabuns L81021 and L71055 were constructed immediately next to walls, it seems reasonable to expect the same to hold true for tabun L81028, which is also partially concealed by an unexcavated balk, just east of the expected continuation of wall W51063. Future excavations may show wall W51063 to continue north almost as far as east–west wall W71039, stopping just short of the wall to create a passage between Aeas 5 and 6. Such a reconstruction would render Area 5 as a broad room, approximately 9 m × 2.5 m. The western wall of the room, also the westernmost wall of the public structure, appears to have had a second entrance to the public structure at its center. The approximately 3-m gap created by the entrance divides the western wall into wall W81049 (between the gap and the casemate wall W11081) and wall W61071 (between the gap and the northwest corner of the public structure). Although most of Area 5 remains to be excavated, the level of its plastered floor (S81048) was reached during the 2011 season in part of the area. Rooms 1 and 2 were mostly excavated during the 2014 season. Work in these areas during the 2015 season uncovered a plastered surface in each. Room one has a well-preserved plaster surface (S81005) that laps against walls W71039 (south), W61071 (west), and W61059 (north). Room 2 also yielded a plastered surface (S81006), though it was not as well preserved as that of Room 1. Room 3 was originally excavated by the HUC project (fig. 19) and revisited in the 2015 season by our team. The area is similar in size and layout to Rooms 1 and 2, with the exception that its enclosing wall is incomplete. Wall W81035 belongs to a later stratum and is founded on the tumbled debris of the public structure, but we lacked the time this season to remove this wall and it remains in situ, obscuring our view of the remains of any southern enclosing wall that may have existed for Area 3. Also remaining in situ in this area is the large vat L81050, previously exposed by the HUC team. The outer dimensions of this vat are 1.46 m × 0.96 m, with a depth of 0.60 m. Dever correctly identifies the vat as an olive oil installation. 22 22.  Dever 1984: 216–17.

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

87

Fig. 19.  Stone basin in Room 3.

Separating Area 3 from Room 4 is partition wall W81051. We worked this season only in the small area west of Stratum 6 wall W81052, revealing a small patch of poorly preserved plastered surface S81047. Dever’s reconstructed plan of his Palace 10,000 (Dever 1985: 221) placed an entrance in Area 4, but it seems more likely that the gap in the line of wall W61059 that he interpreted as an entrance was caused by the extraction of the stones during a later period of the site’s history. As mentioned above, our work this season has enabled us to note certain adjustments required in Dever’s plan of the public structure (his Palace 10,000) based on the HUC finds. In addition to the absence of the western half of the public structure, 23 the plan is missing all of the interior walls that divide Areas 7–12. Thus, the area Dever interpreted as an open “parade ground” 24 was actually six distinct spaces, changing the character of the structure substantially. 25 The HUC team based their date of the mid-tenth century BCE for the public structure on ceramic finds from within the construction fills associated with the structure, which they found to continue to a depth of up to 2.50 m beneath the floor level. 26 Our excavations have found that the public structure was mostly 23.  Though Dever himself (1985: 221) expected the structure to continue to the west. 24. Ibid. 25. ibid. 26. Younker (1991: 21) states that soundings taken from our Areas 4, 13, 14, and 15 revealed the public structure to have been built upon “built-up foundations.” The area was

88

Steven M. Ortiz and Samuel R. Wolff

Fig. 20.  Plaster floor in Room 1.

empty at the time of its destruction, but the initial pottery calls on those ceramics that were recovered from within the destruction debris and on top of the floors do support a 10th-century BCE date for the use and destruction of the building. The central courtyard building was destroyed in a significant and violent event. The destruction debris from the building buried the floors and walls in up to 1.50 m of debris. This debris was mostly deteriorated mud brick, but certain areas within the structure, typically to the immediate west of a wall, contained high concentrations of large boulders. This was the case in Room 1, to the west of wall W71021; Room 2, to the west of wall W61065; Area 6, just south of wall W61064 and west of wall W51088, as well as along the north face of the casemate wall W11081; Area 3, where much of this debris remains in situ beneath later wall W81035; Area 4, north of wall W81053; Area 7; Area 8, west of wall W31066; and Rooms 11 and 12. From the distribution of these boulders, it seems that when the central courtyard building was destroyed, most of the walls fell toward the west. Perhaps this represents a destructive force moving westward from the area of the city gate. The source of this destructive force appears to have not caught the occupants of the public structure by surprise. The building was for the most part cleaned out prior to its destruction, suggesting that its contents were taken when the occupants first leveled and raised by up to 1.50 m with a construction fill. Then, foundation walls were built up from this level, before being buried by nearly a meter in a second construction fill. The floors of the public structure, per Younker, rested atop this second construction fill.

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

89

Fig. 21.  Ivory game board lid.

fled the coming violence. Little was left behind. A complete cooking pot was recovered from the northeastern corner of Room 2. That room also yielded a complete rattle, a bronze bracelet, a bronze ring, and a unique stone object. Discovered in Room 1 was an incised ivory game board and a handful of unworked astragali (fig. 21). This game board features the standard layout of the Game of Twenty Squares. It features three parallel rows of squares, laid out in groups of four, twelve, and four, with rosettes marking five of the twenty squares. A very close parallel to this board was found in Tomb 58 at En­komi (Kiely 2014). The quality of the game board suggests that it was a prestige object, and this in turn may indicate an elite usage of Room 1. Room 2 contained olive pits, pounding stones, gaming pieces (3), a spindle whorl, sling stones, and projectile points. A number of large bones were recovered, including a well-preserved (sheep?) horn, and bovine mandible from two different animals. A rattle and cooking pot were the only complete ceramic artifacts found.

Stratum 8 (Field West) Not much has been exposed of the Iron Age IIA occupation (10th c. BCE) in Field W. Most of the Iron Age was poorly preserved due to later Hellenistic building activity and the early 20th-century excavations by Macalister. Remnants of buildings were uncovered that are tentatively dated to Stratum 8 (see fig. 22). Two building complexes were discerned: only the remnants of a cobbled surface and the outlines of buildings were left by the north face of the casemate wall (Building 52136), and a second building to its north (Building 52057). Building 52136 consists of a main room with two other rooms that abutted the casemate wall. The main room, 5 × 5 m, contained a cobbled surface with a sump/silo that had an entrance from the east. The western part of this building is unknown; it continues into an unexcavated area. Building 52057 is north of Building 52136 and contained an installation. Unfortunately no surfaces were preserved.

90

Steven M. Ortiz and Samuel R. Wolff

Fig. 22.  10th-century BCE buildings in Field W.

Iron Age II: Stratum 7 (9th Century BCE) Previous excavations by HUC, as well as Dever’s excavations in 1984 and 1990, found limited occupation of their Stratum VII (9th century) and therefore were not anticipating a major stratum dating to the 9th century. Surprisingly, the Tandy excavations have revealed a robust 9th century BCE stratum. 27 This stratum was found mostly in Field E, east of the Iron Age city gate. Built directly on top of the Stratum 8 administrative courtyard building and partially incorporating some of its walls, was a domestic building (see fig. 24). This building was destroyed, as evidenced by a rich layer of destruction debris found in the building. 27.  We are tentatively dating this to the 9th century BCE by default. This stratum (7) is above our Stratum 8 destruction (Iron Age IIA, ca. 925 BCE) and beneath our Stratum 6 (8th century BCE, ca. 733/732 BCE). Work on ceramic seriation as well as 14C samples will hopefully confirm the tentative dating.

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

91

Fig. 23.  Stratum 7 (9th century BCE).

Several other buildings (that is, walls) are associated with this stratum, although they lack associated destruction debris and surfaces (see fig. 23). What is apparent is that the nature of this administrative quarter drastically shifted from large administrative buildings to small domestic structures. There appear to be three domestic complexes (fig. 23). Each complex averaged about 10 × 10 m in area. Most of the walls were constructed of a single row of stones. Only the southern parts of these units were excavated. It is clear that they continue to the north under the 8thcentury occupation. Each of these units is built up against the north face of the casemate city wall. The plan for only one of the three complexes has been drawn.

Stratum 7 Domestic Building In 2014, a major destruction was found inside a pillared domestic structure, and this enabled the project to define the nature of one of these three complexes (fig. 24). This building measures about 12 × 8 m. It consists of a main pillared room, with a storage room to its north. Flanking this room to the east were cobbled steps that led up to an elevated tabun (Area 1, fig. 24). To the south of this pillared building was a courtyard with a separate area also containing a tabun (Room 6, fig. 24). Farther south were two rooms (Rooms 8 and 7) built up against the casemate wall. Entrance to these two rooms was from the courtyard area (Area 5). This building unit is unique and not a typical four-room house. The tabuns, material culture, and construction all point to the fact that this is a domestic unit. It is built on top of the destroyed Stratum 8 and reused some of the Stratum 8 walls (W51063, W71070, W61071). The surfaces are beaten earth, with some cobble and flagstone areas.

92

Steven M. Ortiz and Samuel R. Wolff

Fig. 24.  Stratum 7 (9th century BCE) domestic building. his building contained a rich destruction layer (L71012), which included numerous storage jars as well a large roof roller.

Entrance to the domestic unit was from the east (Area 4) or the northeast end (Area 1). The eastern end of the structure is too poorly preserved to identify or rule out the presence of an entry on that side. Area 1 is the only other possible entrance with an opening to the north. One difficulty with this, however, is the presence of tabun L61070, which was found in Area 1 against the west face of wall W61071. While this tabun would have made passage through Area 1 more difficult, it was not so large as to completely impede foot traffic and rule out Area 1 as an entrance. It is also possible that the tabun is from a later squatter phase. Area 1 has a rough cobbled pavement (S71052) that appears to form two steps leading down into the pillared building from the north. These steps lead to Room 2, which is approximately 2 m × 4 m and features a small stone-lined bin L81016

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

93

Fig. 25.  Stratum 7 destruction, Room 8.

set into the floor against the east wall W71070. Room 2 is separated from Room 3 by two pillars (L71050, L71051) of similar size. Each of these is 0.50 m × 0.50 m and 0.80–0.90 m in height. The pillars stand in the center of the space created by Rooms 2 and 3, so that the two rooms mirror each other in size and shape. Ceramic finds indicate that Room 3 was used for storage, particularly in its southern half, while Room 2 housed fewer vessels, likely due to its status as a passageway connecting the northern half of the building with rooms further to the south. Just north of this pillared space is an enclosed room (Room 10) for which we were unable to discern an entrance. The north wall (W61053) of this room runs at a somewhat different angle from all the other east–west walls in the pillared building and is built of marginally larger boulders than those used elsewhere in the structure. It may be that this room was added to the exterior of the pillared building subsequent to the initial construction of the structure. This may help to explain the apparent lack of a clearly defined entrance into the room. Also of note is that, while most of the pillared building was rich in ceramic and other small finds, Room 10 was by comparison nearly devoid of objects. South of, and accessed directly from, Room 2 is Room 5. This room acted as a hub of sorts for the pillared building; from it one could access Room 6 to the west, Area 4 to the east, or Rooms 7 and 8 to the south. This room featured a plastered installation consisting of three basins. One basin was a shallow, elongated depression

94

Steven M. Ortiz and Samuel R. Wolff

Fig. 26.  Stratum 7 destruction, Room 8.

that was well plastered and lined with cobbles. The precise use of this feature is unclear, but it was well constructed. Just to the southeast of this plaster-and-cobble basin, we uncovered the remains of a plastered installation and another one to the southwest, plastered bin L81042. Both of these plaster-lined bins had been sunk into the floor. A number of storage jars and other vessels were found in this area (Area 5). Thus, Room 5 seems to have been utilized both as a work space and a storage area. Room 6 was entered from the west side of Room 5 and may constitute a small, open courtyard. This space is approximately 1.50 m × 2.50 m and features a tabun (L81034) in its southwest corner. From this tabun we recovered numerous (100+) olive pits for potential carbon-14 dating. Room 7 was entered via Room 5 and in turn gave access to Room 8. Both of these rooms abut the casemate wall W11081. Numerous objects were recovered in Room 8 during the 2011 season, including several vessels and two stone game boards. Room 7, by contrast, was mostly empty. Two storage jars were discovered against a stone bench (L81014) built against the casemate wall, one of which contained three pieces of a long metal object.

Summary, Stratum 7 During Stratum 7, the administrative quarter of Stratum 8 changed to a domestic quarter. It is clear that Stratum 7 reused the earlier fortifications and casemate wall line. The 10th-century-BCE monumental architecture (e.g., pillars, walls) is missing from this stratum. Although only one unit has been fully excavated, there is evidence for other units. All walls are constructed of a single course of unhewn stones, with several remnants of tabuns found in the units. All of these walls were stratigraphically below our Stratum 6 administrative buildings. One of the difficulties

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

95

Fig. 27.  Stratum 6a.

of establishing plans of these units is that this area was previously excavated by Macalister, who removed a lot of the stratigraphic connections between the various wall fragments. In addition, this area was heavily damaged by the activity of the Hellenistic period. The pottery found in the destruction of the pillared building has not been restored. Most of the restoration of Stratum 7 pottery is from Room 8 (excavated in the 2011 season). Typical ceramic forms are store-jars, cooking pots, and cooking jugs (see fig. 27). We tentatively date this destruction as contemporary with the destructions of Gath, Tel Zayit, and Tel Goded (Maeir 2004b; 2012: 38, 43–49; Tappy 2008; Tappy et al. 2006: 9–16, Gibson 1994: 230–31). Most scholars associate this destruction with the campaign of Hazael, King of Aram-Damascus, in the second half of the 9th century BCE (2 Kgs 12:18; Maeir 2004a; 2012; Finkelstein and Piasetzky 2007; 2009).

Iron Age II: Stratum 6 (8th Century BCE) Our Stratum 6 has already been discussed in previous publications (Ortiz and Wolff 2012) and will be briefly summarized here. Stratum 6 consists of a major phase, with some minor rebuilds. In addition to the rebuilding of the gate into a four-chambered gate, 28 the Stratum 7 city was drastically changed. A series of public buildings was constructed just west of the gate. These buildings were built up against the north face of the casemate wall. The domestic structures abutting the 28.  Dever et al., “Further Excavations at Gezer, 1967–71,” Biblical Archaeologist 34 (1971): 118.

96

Steven M. Ortiz and Samuel R. Wolff

Fig. 28.  Four-room house.

line of the casemate wall went out of use, and three major administrative buildings were built, changing the area west of the gate back into an administrative quarter. The domestic quarter was relocated to the northwest. Most of the archaeological data was removed by Macalister; all that remained were the remnants of the foundation of the wall lines. All the pottery was mixed, with Iron Age and Hellenistic sherds intermingled. Only one surface in the three administrative buildings was preserved (cobble surface 21071 in Building C), where an Iron Age juglet was found. To the north of this building, the project was fortunate that Hellenistic buildings prevented Macalister from penetrating to the Iron Age levels, and in them we found a large four-room house with more than a meter of destruction debris.

Iron Age Four-Room House To the northwest of these administrative buildings was a domestic quarter where we excavated a four-room house. The four-room house consisted of the southern

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

97

Fig. 29.  Aerial photo with plan of the four-room house (Field W).

wing, which had (1) a plaster-floored room to the west and a cobble-surfaced room to the east; (2) a central room, which contained pottery vessels (mainly storage jars, grinding stones, and loomweights; a few tabuns were also identified here; (3) a northern room, which remains largely unexcavated; and (4) a broadroom subdivided into two rooms to the west. The latter was filled mostly with broken storage jars. The four-room house consists of three long rooms (a central room flanked by parallel northern and southern rooms), separated by large limestone pillars (average dimensions 0.50 × 0.50 × 1.00 m), with a broadroom to the west, which was subdivided into two smaller rooms by a transverse wall. The eastern and northern borders of the house remain unexcavated. The area of the house as a whole is estimated at 135 m2, considerably larger than typical four-room houses found previously at Gezer and at other urban sites. The building and its contents were sealed by burnt mud-brick destruction debris, testimony to a considerable conflagration. The portable finds from the building included basalt grinding stones, loom weights, and a sizable ceramic assemblage, consisting primarily of restorable storage jars, with lesser numbers of bowls and kraters (no cooking pots or jugs). The four-room house contained more than 30 whole vessels. The vast majority were store-jars. Notably, only 1 cooking pot was found The juxtaposition in the same

98

Steven M. Ortiz and Samuel R. Wolff

Fig. 30.  Vessels found in the Assyrian destruction of the four-room house (Stratum 6).

room (southern broadroom) of lamelek storage jars (without the typical stamps), typical of Iron IIB Judean sites and Phoenician torpedo-shaped storage jars, characteristic of coastal assemblages, clearly illustrates that Gezer straddled the geopolitical boundary between these two zones. The ceramic assemblage and small glyptic finds date to the eighth century BCE and are tentatively associated with the destruction of the site by Tiglath-Pileser III in 734 BCE. Evidence of a similarly dated destruction was found in the excavations in nearby Field VII. To the south of the four-room house, remains of a cobbled surface (ca. 6.5 m long and 3 m wide, exposed thus far) was uncovered, perhaps representing a portion of a street leading uphill from the Iron Age gateway toward the west. One weight from an olive press installation, like that found in nearby Field VII, was found on this surface. The ridged holemouth jar, the only one found in the assemblage, is a late-8th and 7thcentury type. Its presence here, in what is apparently a 734 BCE context, provides us with a terminus post quem for this type. Note that the cooking pot is not the typical Judean form but rather a coastal type. When viewed in conjunction with the contemporary remains in Field VII, it seems that this area was domestic in function. Nonetheless, the architecture revealed in our excavation seems more massive and more robust than that revealed in Field VII. Note that the pillars, for example, are smaller in size and are situated closer together in Field VII. Our building measures some 140 m2, much larger than most examples found in urban settings. The building was destroyed in a massive conflagration. The four-room house was located immediately north of what appears to have been an area devoted to the production of olive oil. Note that there several installations and a weight were found here, which lead us to this conclusion. In Field W, as we removed the stones from Building C, the pillar base was discovered to be an upturned olive oil vat. This vat has a rectangular shape and was probably in use in Phase 1 of Building C. It went out of use with the olive oil installation. In addition, it was determined that a large circular installation excavated by Macalister dates to Stratum 6a. The pottery

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

99

and the glyptics all seem to point to Tiglath-Pileser III’s ca. 733 BCE campaign as the occasion of the building’s destruction. We do not know precisely when the building was constructed. Evidence in the southern end points toward two phases of occupation, with the cobble-surfaced and plaster-surfaced rooms perhaps being later in date than the rest.

Overview of Results Although the project has completed eight seasons of excavation, most of the results have appeared in the past two seasons of excavation. This was anticipated, because Gezer has been extensively excavated. Unfortunately, while the Macalister excavations revealed that Gezer was an important city during the second and first millennium BCE, it is difficult to distinguish the various historical periods based on those excavations. The stratigraphic picture was clarified with the HUC excavations of the 1960s and 1970s. A majority of these excavations focused on the Middle Bronze Age. Although the Phase Two project shifted focus to the Iron Age, many questions were still left unanswered. 29 The renewed Tandy Excavations, while still emphasizing the stratigraphic foundation of the HUC excavations, has intentionally accomplished this task by incorporating broad architectural overviews. The emphasis was placed on an area where a picture of shifts in the process of urbanization would be most likely. Though the results are tentative, a picture of this process is starting to be become clear. First, the 2nd-millennium-BCE ancient city of Gezer extended its occupation to the east slope of the western hill. While evidence of the MB glacis has been found encompassing both the western and eastern hills, there is now evidence of an LB IIB patrician house in this area. Scholars have attempted to understand the Amarna Period city of Gezer by attempting to find elite buildings in Macalister’s plans. 30 We now have a partial plan of an elite building dating to the 14th century BCE. In a few more seasons, as the excavations of the Iron Age I strata proceed, the project should develop a complete plan of this LB building. The results of the Tandy excavation have shown that Gezer was an unwalled city during the Late Bronze period. Even though the well-known debate about whether Gezer was a walled city during the Late Bronze has now been settled, 31 ironically, we have determined that Gezer during the Iron Age I was also a walled city. It is clear that Gezer experienced growth and contraction in its urbanization. The Iron Age I Canaanite city was destroyed and the city in this area experienced 29.  The Phase Two excavations were directed by J. Seger. The Iron Age occupation strata was limited to the Iron Age I buildings on the acropolis, the reexcavation of the Iron Age gate, and the domestic units in Field VII. 30.  See Ortiz and Wolff (forthcoming) on the history of debates about the character of Late Bronze Age Gezer. 31.  See Bunimovitz (1983), Dever (1986; 1993), and Finkelstein (1981; 1994) for discussion. Also, see recently Seger and Hardin (2013: 31–33).

100

Steven M. Ortiz and Samuel R. Wolff

a drastic shift to public buildings, particularly a large administrative central courtyard building adjacent to a gate complex and fortification system. This shift is associated with the rise of the United Monarchy. That Gezer was an administrative city is attested in the biblical memory of Solomon’s building projects (1 Kgs 9:15). This new urban center and central authority was short-lived, because Gezer (Stratum 8) was destroyed by Shishak. The character of the public quarter shifted into domestic units. These units continued to use the Stratum 8 casemate wall and gate system, although the central courtyard building went out of use. The domestic units were not the typical four-room house associated with the Israelites but probably represent a local Canaanite household tradition. 32 These units were destroyed in the 9th century BCE, and a new city was rebuilt. At that time, the quarter reverted back to public buildings, and the domestic quarter moved to the northwest, where we excavated a large four-room house adjacent to the four-room houses excavated by HUC (Gitin 1990). During Stratum 6, evidence of Judean influence is seen in the form of several fragments of baking trays and lmlk handles. One of the research questions that the project is addressing is the shifting ethnic and political boundaries of Gezer in the context of its location in the Aijalon Valley, between the Israelite and Judahite highlanders and the Philistine coast. More recently, scholars are realizing that ethnic and political boundaries are more complex then the simplistic juxtaposition of Israelites vs. Philistines. 33 Because scholars now are considering the complex ethnic and political groups in the Yarkon basin and the northern Shephelah (Gadot 2011; Mazar 2009), Tel Gezer is an ideal site to address these developments. Gezer sits on a border (between the Judean Hills, the Philistine Coastal Plain, and the north [that is, Israel]). In addition, it also has a history of shifting borders (during the Iron Age I, when it was Canaanite, compared with during the early Iron Age II, when it was Israelite, and during the later Iron Age, when it was Judahite). We will start to see this picture clearly only when we put all the artifacts together according to their stratigraphic context. 32.  Gilboa, Sharon, and Zorn 2014. 33.  Bunimovitz and Lederman (2008) and (Faust (2012).

Bibliography Ben-Ami, D., and Wazana, N. 2013 Enemy at the Gates: The Phenomenon of Fortifications in Israel Reexamined. Vetus Testamentum 63: 368–82. Bunimovitz, S., and Lederman, Z. 2008 A Border Case: Beth-Shemesh and the Rise of Ancient Israel. Pp. 21–31 in Lester Grabbe (ed.). Israel in Transition: From Late Bronze II to Iron IIa (c. 1250–850 B.C.E.). Volume I: The Archaeology. New York: T. & T. Clark. 2009 The Archaeology of Border Communities: Renewed Excavations at Tel Beth Shemesh, Part I: The Iron Age. Near Eastern Archaeology 72: 114–42.

Tel Gezer Excavations 2006–2015: The Transformation of a Border City

101

Dever, W. G. 1984 Gezer Revisited: New Excavations of the Solomonic and Assyrian Period Defenses. The Biblical Archaeologist 47: 206–18. 1985 Solomonic and Assyrian Period ‘Palaces’ At Gezer. Israel Exploration Journal 35: 217–30. 1986 Late Bronze Age and Solomonic Defenses at Gezer: New Evidence. Bulletin of the American Schools of Oriental Research 262: 9–34. Fantalkin, A., and Finkelstein, I. 2006 The Sheshonq I Campaign and the 8th-Century BCE Earthquake: More on the Archaeology and History of the South in the Iron I–IIA. Tel Aviv 33:18–42. Finkelstein I., and Piasetzky, E. 2007 Radiocarbon, Iron IIa Destructions and Israel: Aram Damascus Conflict in the 9th Century. Ugarit Forschungen 39: 261–76. Finkelstein, I., and Piasetzky, E. 2009 Radiocarbon-Dated Destruction Layers: A Skeleton for Iron Age Chronology in the Levant. OJA 28: 255–74. Gadot, Yuval 2011 “Houses and Households in Settlements along the Yarkon River, Israel, During the Iron Age I: Society, Economy, and Identity.” Pp. 155–82 in A. Yasur-Landau, J.  R. Ebeling, and L. B. Mazow (eds.), Household Archaeology in Ancient Israel and Beyond. Leiden: Brill. Garfinkel, Y., and Ganor, S. 2009 Khirbet Qeiyafa Volume I: Excavation Reports 2007–2008. Jerusalem: Israel Exploration Society and the Institute of Archaeology, Hebrew University of Jerusalem. Gibson, S. 1994 The Tell el-Judeideh (Tel Goded) Excavations: A Re-appraisal Based on Archival Records in the Palestine Exploration Fund. Tel Aviv 21: 194–234. Kiely, T. 2014 Game Box with Chariot Hunt. P. 43 in J. Aruz, S. B. Graff, and Y Rakic, eds., Assyria to Iberia at the Dawn of the Classical Age. New York: Metropolitan Museum of Art. Lehmann, G., and Killebrew, A. 2010 Palace 6000 at Megiddo in Context: Iron Age Central Hall Tetra-Partite Residencies and the Bit-Hilani Building Tradition in the Levant. Bulletin of the American Schools of Oriental Research 359: 14–33. Maeir, A. M. 2004a The Historical Background and Dating of Amos VI 2: An Archaeological Perspective from Tell es-Safi/Gath. Vetus Testamentum 54: 319–34. 2004b Bronze and Iron Age Tombs at Tel Gezer, Israel: Finds from the Excavations by RaymondCharles Weill in 1914 and 1921, with contributions by N. Panitz-Cohen, D. Barag, O.  Keel, N. Applbaum and Y. Applbaum. British Archaeological Reports International Series 1206. Oxford: Archaeopress. 2009 The Iron Age Dwelings at Tell Qasile. Pp. 319–36 in D. Schloen, ed. Exploring the Longue Duree: Essays in Honor of Lawrence E. Stager. Winona Lake, IN: Eisenbrauns. 2012 The Tell es-Safi/Gath Archaeological Project 1996–2010: Introduction, Overview and Synopsis of Results. Pp. 1–88 in A. M. Maeir, ed. Tell es-Safi/Gath I: The 1996–2005 Seasons, Part 1: Text. Ägypten und Altes Testament 69. Wiesbaden: Harrassowitz. Ortiz, S. M., and Wolff, S. 2012 Guarding the Border to Jerusalem: The Iron Age City of Gezer. Near East Archaeology 75: 4–19.

102

Steven M. Ortiz and Samuel R. Wolff

forthcoming  A Reevaluation of Gezer in the Late Bronze Age in Light of Renewed Excavations and Recent Scholarship. In Aren M. Maeir, Itzik Shai, and Chris McKinny, eds., And the Canaanites Were Then in the Land: New Perspectives on the Late Bronze Age of Southern Canaan and Its Surroundings. Winona Lake, IN: Eisenrauns. Osborne, J. 2012 Communicating Power in the Bit-Hilani Palace. Bulletin of the American Schools of Oriental Research 368: 29–66. Seger, J., and J. W. Hardin 2013 Gezer VII: The Middle Bronze and Later Fortifications in Fields II, IV, and VIII. Winona Lake, IN: Eisenbrauns. Shai, I., Maeir, A., Gadot, Y. and Uziel, J. 2011 Differentiating between Public and Residential Buildings: A Case Study from Late Bronze II Tell es-Safi/Gath. Pp. 107–131 in A. Yasur-Landau, J. R. Ebeling, and L. B. Mazow. Household Archaeology in Ancient Israel and Beyond. Leiden: Brill. Sharon, I., and Zarzecki-Peleg, A. 2006 Podium Structures with Lateral Access: Authority Ploys in Royal Architecture in the Iron Age Levant. Pp. 145–67 in Confronting the Past: Archaeological and Historical Exxays on Ancient Israel in Honor of William G. Dever, ed. S. Gitin, J. E. Wright, and J. P. Dessel. Winona Lake, IN: Eisenbrauns. Shavit, Alon 2000 Settlement Patterns in the Ayalon Valley in the Bronze and Iron Ages. Tel Aviv 27: 189–230. Tappy, R. E. 2008 Tel Zayit and the Tel Zayit Abecedary in their Regional Context. Pp. 1–44 in R. E. Tappy and P. K. McCarter, eds. Literate Culture and Tenth-Century Canaan. Winona Lake, IN: Eisenbrauns. Tappy, R. E., McCarter, P. K., Lunberg, M. J., and Zuckerman B. 2006 An Abecedary of the Mid-Tenth Century BCE from the Judean Shephelah. Bulletin of the American Schools of Oriental Researc 344: 5–46. Ussishkin, D. 1980 Was the “Solomonic” City Gate at Megiddo Built by King Solomon? Bulletin of the American Schools of Oriental Research 239: 1–18. Wolff, S., and Finkielsztejn, G. 2009 Two New Hellenistic Lead Weights of the Tanit Series. Pp. 497–506 in J. D. Schloen, ed. Exploring the Longue Durée: Essays in Honor of Lawerence E. Stager. Winona Lake: Eisenbrauns. Wolff, S., Arbino, G., and Ortiz, S. 2015 Macalister at Gezer: A Perspective from the Field. Pp. 43–49 in S. Wolff, ed. Villain Visionary? R. A. S. Macalister and the Archaeology of Palestine. Proceedings of a Workshop Held at the Albright Institute of Archaeological Research, Jerusalem, on 13 December 2013. Leeds: Maney. Younker, R. W. 1991 A Preliminary Report of the 1990 Season at Tel Gezer: Excavations of the “Outer Wall” and the “Solomonic” Gateway ( July 2 to August10, 1990).” Andrews University Seminary Studies 29: 19–60. Zerlwer-Silberberg, Paule, 2012 Les Fouilles de M. Raymond Weill à Tell-Gezer (1914 et 1924): Le mémoire perdu et retrouvé de Mme Silberberg-Zelwer (1892–1942). Publication de deux campagnes de fouilles archéologiques conduites par M. Raymond Weill à Tell Gezer (1913–1924). 1942. Jerusalem: Centre de recherché français à Jèrusalem.

Tell Halif in the Late Bronze and Iron Age Oded Borowski

Location Tell Halif (in Arabic, Tell el-Khuweilifeh) is a 3-acre mound located in the foothills of Mount Hebron, some 12 miles northeast of Beersheba. The site is located at a meeting point between four environmental zones: the Judean hill-country to the east; the Shephelah to the north; the Negev to the south; and to the west the Coastal Plain (Seger et al. 1990). The site is positioned on the first line of hills facing the plain, providing a vantage point guarding its agricultural and grazing lands and its water sources. Tell Halif is also situated on a major branch of the Via Maris leading from the coastal plain to the Judean hill-country toward Hebron and Jerusalem.

Identification Tell Halif has been identified at various times and by various scholars with several biblical sites such as Ziklag, Sharuhen, Kiriath-sepher, Goshen, Hormah, and Rimmon or En-Rimmon (for summaries, see Borowski 1988, Dessel 2009: 12–13). The latter identification is presently the one accepted by the majority of scholars.

History of Exploration and Excavation The site attracted attention already in the 19th century (Guerin 1869: 341–62) and in the early 20th century (Albright 1924); however, no archaeological work was performed until the 1950s, when salvage excavations were undertaken by D. Alon. Further salvage work was carried out by Alon and others during the 1960s (HA 1964) and 1970s (HA 1972, HA 1974). Long-term excavations by the Lahav Research Project (LRP) commenced in 1976 under the direction of J. D. Seger (Seger and Borowski 1977) and have been conducted through four phases: Phase I (1976–77, 1979–80), directed by J. D. Seger, was carried out in Fields I–III, Sites 72 (Borowski 2013) and 101 (Dessel 2009), and other probes on and around the mound, all of which investigated the basic stratigraphy of the site; a regional survey was also conducted in conjunction with this phase. Phase II (1983, 1985–87, 1989), also directed by Seger, continued the work in all three fields and concentrated on the recovery of remains from the Early and Late Bronze Ages (Seger et al. 1990); Phase III (1992–93, 1999), co-directed by J.  F. Jacobs and O. Borowski, investigated the Iron Age II remains 103

104

Oded Borowski

Fig. 1.  Tell Halif: excavated areas.

in Field IV at the western end of the tell ( Jacobs 1998, Hardin 2010); Phase IV (2007–2009, 2014–2016), directed by Borowski, continues the work of Phase III at an adjacent area (Field V) (Borowski 2008, Borowski 2009, Borowski 2010). 1 Additional archaeological work was carried out in 1995–96 under the direction of T. E. Levy and D. Alon by the Nahal Tillah Project, which investigated the Eastern Terrace of Tell Halif and other Chalcolithic and Early Bronze Age sites in the vicinity (Levy et al. 1997). 1.  The LRP has continuously been affiliated with the American Schools of Oriental Research (ASOR) and the W.  F. Albright Institute of Archaeological Research (AIAR); during Phases I and II, it was also affiliated with the Hebrew Union College–Nelson Glueck School of Biblical Archaeology in Jerusalem. Additional support has been provided by consortium member institutions and private donations. Special thanks are due to Kibbutz Lahav and the Joe Alon Center for Regional and Folklore Studies for the help they extended throughout the years.

105

Tell Halif in the Late Bronze and Iron Age

Fig. 2.  Stratigraphical Table: Tell Halif Stratum

Archaeological Period

Date

I

Modern Arab

1800–1948

II

Islamic–Crusader

 700–1500

III

Late Roman/Byzantine

 200–600

Early Roman

  100 BCE–200 CE

IV

Hellenistic

 300–100

V

Persian

 500–300

Late Iron II/Babylonian

 650–500

Iron II

 700–650

Gap

Gap VIA

— Destruction VIB

Iron II

 800–700

VIC

Iron II

 850–800

VID

Iron II

 900–850

VII

Iron I

1200–900

VIII

Late Bronze IIB

1300–1200

IX

LB IIA

1400–1300

— Destruction X XI

LB IB

1475–1400

LB IA

1550–1475

Gap

Middle Bronze II

1850–1550

Gap

Early Bronze IV

2300–1850

XII

EB IIIB2

2400–2300

XIII

EB IIIB1

2450–2400

XIV

EB IIIA2

2500–2450

— Destruction XV

EB IIIA1

2600–2500

EB II

2900–2600

XVI

EB IC

3000–2900

XVII

EB IB

3100–3100

XVIII

EB IA

3200–3100

XIX

Chalcolithic

3500–3200

Gap?

Site History The general stratigraphy, which emerged from the work carried out at and around the tell throughout the four phases, contains more than 20 major strata (fig. 2, stratigraphy).

Early Periods The earliest settlements, Strata XIX–XVI (3500–2900 BCE), occupied the Eastern Terrace close to the agricultural lands, first in improved natural caves and later

106

Oded Borowski

in domestic structures that formed a small village (Levy et al. 1997, Dessel 2009). During this period, the site exhibits strong connections with Early Dynastic Egypt. After a gap of about 300 years, the site was reoccupied during the Early Bronze IIIA1 (2600–2500 BCE) further up the hill (Stratum XV); however, this time it was massively fortified, as can be seen by the remains of a 7.5-m-wide tower, a 3.5-mwide defensive wall, and a glacis. Remains from this period of occupation were discovered in Fields I and III. The settlement came to its end in a fiery destruction that produced a layer of ash and other debris 3.5 m deep. The site was reoccupied again in the Early Bronze III (Strata XIV–XII; 2500–2300 BCE) but was not refortified (Seger et al. 1990). The occupants of Early Bronze Tell Halif were engaged in the production of flint tools: several cores and the remains of a knapping workshop were found in Stratum XIII (Seger et al. 1990: 16–17, 22–23). It seems that they specialized mostly in the production of Canaanean blades.

Late Bronze Age Following a gap of about 750 years (EB IV–MB II), settlement was resumed at the tell during the Late Bronze Age IA (Stratum XI; 1550–1475 BCE). After a transitional period that yielded several sub-phases that produced typical ceramics and a Hyksos-type scarab, the site where Field I is located was occupied by a large building with a central courtyard (6 × 6.5 m) surrounded by rooms on all sides (Stratum X; 1475–1400 BCE) ( Jacobs 1987). This type of structure, known as a “residency” building, is associated with Egyptian rule in Palestine, and the stratum can be related to the reign of Thutmoses III (1479–1426 B.C.E.) and probably was mentioned as r-n-m in the Karnak city-list of this pharaoh (Borowski 1988: 25). This stratum came to its end with a massive destruction and was followed by a partial resettlement as Stratum IX (1400–1300 BCE). Some time afterward, in the latter part of the Late Bronze Age (LB IIB, Stratum VIII; 1300–1200 BCE), a large mud-brick platform was constructed on top of the remains of the previous strata (X–IX). A series of stone-lined storage pits was built into the platform, and it seems that the site was used as a grain-storage center, probably serving as an entrepôt (Seger et al. 1990). A storage jar with a handle inscribed in proto-Canaanite script was recovered from one of the pits (Seger and Borowski 1977, Borowski 1982).

Iron Age I Remains from the Iron Age I (Stratum VII; 1200–900 BCE) were uncovered in Fields I–III and possibly also in Field V. Because of the limited exposure, very little can be said about the nature of the settlement during this period. Architectural finds in Fields II and III, and most recently in Field V, include stone-lined storage pits, probably for grain. The pits in Field III were built into a massive stone platform that was constructed on the remains from the EB and LB strata. The existence of the storage pits suggests that the agricultural nature of the site did not change from the LB IIB to the Iron Age I. Remains from Fields I and II include degenerate-style Philistine pottery, which suggests connections with the coastal plain. Remains from

Tell Halif in the Late Bronze and Iron Age

107

varioius strata of fish bones and seashells strengthen this conclusion. From Field II came a headless clay female figurine, the front of which was decorated with punctates and the back with long, etched vertical lines resembling hair (Dessel 1988).

Iron Age II The greatest exposure of any strata is that of the Iron Age II Stratum VIB (800– 700 BCE), followed by Stratum VIA (700–650 BCE), both of which are close to the tell’s surface and could be exposed easily. These strata yielded remains of domestic structures built adjacent to the fortification system, yielding much information about daily life in their respective periods, a topic to be discussed in detail below.

Later Periods The latest strata on the tell date from the Persian (Stratum V; 500–300 BCE) and Hellenistic (Stratum IV; 300–100 BCE) periods. Pits dated to the Persian period were found in Fields I, III, and IV, while in Field II the remains of a large building, possibly administrative, were uncovered. Ceramic evidence suggests that the building was constructed in the late Persian period. In Field IV, more than 560 fragments of clay figurines dated to the late Persian–Hellenistic period were found scattered, mixed with topsoil ( Jacobs 1998; 2015). More figurine fragments were discovered in the adjacent Field V (Borowski 2010). It seems that these fragments originated at a cult site on the tell and were brought to the area of Fields IV and V during ground leveling activities that probably were carried out for agricultural purposes. On the summit of the mound in Field II, architectural remains belonging to Stratum IV were found, including remnants of a residential building, as can be determined by the several ovens found on the living surfaces. During the Roman/Byzantine periods (Stratum III; 200–600 CE), the tell itself was sparsely occupied, and its surface might have been used mostly for agricultural and grazing activities. Remains of isolated structures from this period were recently uncovered in Field V. The main settlement took place on the Eastern Terrace and at other sites in the vicinity, such as Horvat Rimmon (Khirbet Umm er-Rammamin), about 1 km south of Tell Halif (Kloner 1993). The Jewish nature of the population in the region at the time is attested in Eusebius’s Onomasticon, where the early church father describes the existence of two Jewish villages in the area, Tilla (or Talla) and Rimmon. Cemeteries belonging to these two villages have been excavated by several archaeologists at various times (Kloner 1984). The Terrace was resettled in the Early Arab period (ca. 700 CE), and the area surrounding the tell saw some activity during the Crusades. Modern Arab settlement, mostly in caves, occupied the foot of the tell at Khirbet Khuweilifeh until 1948.

Iron Age II (the Judahite Settlement) As previously mentioned, the greatest exposure was of Strata VIB and VIA. These strata are related to each other; both belong to the Judahite town that occupied the mound during the Iron Age II. Both strata produced a tremendous amount of

108

Oded Borowski

Fig. 3.  Pillared house.

information that enables us to reconstruct what life was like in a small town in the Judean kingdom in the Shephelah. This is possibly due to the fact that, at the end of the eighth century BCE, the site suffered sudden and total destruction by fire. Stratum VIB, parts of which were preceded by Stratum VIC, which is dated to the 9th century BCE, contains remains of destroyed domestic structures adjacent to the city wall, covered with a thick layer of ash produced by a fire hot enough to calcify the mud bricks. This fire also helped preserve organic material that was charred or carbonized. That the fire was a result of a military action is evident from the numerous arrowheads, sling stones, and remains of other projectiles and weapons uncovered in the destruction context. Ceramic evidence similar to Lachish III suggests very strongly that the site was destroyed during Sennacherib’s military campaign in Judah in 701 BCE (Borowski and Seger forthcoming). LMLK stamped jar handles found on the surface and in situ in Field V suggest very strongly that the town at Tell Halif participated in Hezekiah’s revolt against Sennacherib (Borowski 2005). Because the end came quickly, most of the contents of the structures remained in situ, and we get a freeze-frame picture of life at the point just before the destruction occurred. Had the inhabitants had the time to leave the place leisurely, they would have removed much of the contents. Furthermore, Stratum VIA, which was inhabited most likely by returning refugees from Stratum VIB, was abandoned in

Tell Halif in the Late Bronze and Iron Age

109

Fig. 4.  Loom weights in situ on the floor.

haste, with the contents of the structures also left behind intact. The similarities between the material culture assemblages and the rebuilt structures of the two strata lead us to conclude that their inhabitants were related. The way these two strata reached their end enables us to reconstruct life as it was just before the demise of the town. It appears that the town was well designed. It was surrounded by a casemate wall into which four-room, pillared houses were built (Borowski 2008, Hardin 2010). Protecting and fronting the city wall, a steep glacis was constructed at an angle of 35–40 degrees. It was built by importing large amounts of dirt, ash, and stones placed in layers, anchored by earlier walls, stone piles and a retaining wall at the bottom. The glacis was covered with a layer of dry-laid flagstones. In Field III, an offset tower was projecting out to protect the wall-line. It is apparent that a major occupation of the local inhabitants was textile production, as seen from the large number of donut-shaped, sun-dried clay loom weights and other tools such as spindles made of various materials (e.g., pottery sherds and stones) and bone spatulas (pick-up sticks) for pattern weaving found in almost every Iron Age II structure. Many of the loom weights were found stored in circular or semicircular installations built specifically for that purpose. Field V

110

Oded Borowski

Fig. 5.  Incense altar.

yielded a textile workshop that was in operation when the city fell, as can be determined by the rows of loom weights found in situ on the floor where the looms stood before the collapse of the building. One major source for spinning the yarn used at Tell Halif was wool and hair that were produced on site, as indicated by remains of sheep and goat bones. Another occupation was wine-making, as indicated by the evidence found in Area M of the F7 dwelling in Field IV. This includes “12 large storage jars (3 of which tested positive for wine), the funnel and strainer (which also tested positive for wine), and numerous stoppers” (Hardin 2010: 159). The size of the operation suggests that it was meant for distribution, not just for personal consumption. Several of the domestic structures yielded cult objects, some of which were related to the domestic and industrial activities. Part of Area B in Room 2 of the F7 dwelling can be identified as a “shrine room” on the basis of the two standing stones, a fenestrated stand, and the head of a pillar figurine (Hardin 2010: 134). Heads of pillar figurines were found at other parts of the site, including Field V, where other cult objects were found, including two small limestone incense altars (one with traces of soot), kernos oil lamps, a fragment of a zoomorphic vessel, and more. The altar containing soot is decorated with etched figures of animals and a

Tell Halif in the Late Bronze and Iron Age

111

human in a style reminiscent of Arabian incense altars. However, analyses of the altars suggests very strongly that they were produced locally (Bang et al. 2012).

The Iron Age II Cemetery Southwest of the tell, across the valley, in the Lahav forest, a large cemetery was identified and partially investigated. The first tomb was excavated by A. Biran and R. Gophna in 1965 (Biran and Gophna 1969). Additional tombs were excavated since then (for a summary, see Borowski 2013: 2–3.). The cemetery was used between the 10th and 7th centuries BCE, with the height of its use in the 8th century. The tombs are of the typical Judahite type, containing a burial chamber with benches along the walls and repositories in the corners. Some of the tombs were reused in the Roman–Byzantine period and in more recent times by Bedouin. Although most of the burial caves were disturbed some time in the past, most yielded ceramic evidence that enabled their dating, and personal items such as jewelry and metal objects (arrowheads, etc.) (e.g., Borowski 2013: pls. 9, 16). One unique find in one of the tombs was a rimmon (pomegranate) bowl, a shallow ceramic bowl with a hollow pomegranate protruding in its center in a manner reminiscent of a kernos vessel (Borowski 1995). Other objects worth noting are a horse-and-rider figurine and a scaraboid seal (Borowski 2013: 68–69).

Contributions of the Project There are several contributions that can be assigned to the work at Tell Halif, and they can be divided into two classes: (a) technological; and (b) historical. Being in the field since the mid-1970s, our project was the first, in 1983, to introduce Osborne portable computers into the field; they were used to record data such as material culture samples, pottery readings, and produce final field reports on a daily basis. Another technological innovation was the introduction in 1987 of Ground Penetrating Radar (GPR) for surveying possible areas for future excavations. With the help of the GPR, we discovered an Iron Age II cistern the opening of which was buried and hidden. The plastered cistern belonged to a run-off water collection system on the tell proper. The GPR also indicated the locations of buried walls and pillars and helped to determine excavation sites such as Fields IV and V (Borowski and Doolittle 1996). Another contribution was the introduction of household archaeology, beginning in 1992 with the initiation of Phase III of the project and the opening of Field IV. This was shown in the approach taken to the total recovery of material culture from the living floors of the excavated structures. When a surface was identified, it was divided into small units (0.25 m × 0.25 m; or 0.50 m × 0.50 m, depending on the richness and scatter of the assemblages), each of which was assigned a separate pottery basket. The location of each was plotted on a top plan and the information collected in this manner assisted in the restoration and placement of vessels and other objects. Additionally, samples of micro-artifacts were collected in association

112

Oded Borowski

with the pottery baskets, which assisted in the reconstruction of activity areas. For an example of the results of this approach, see Hardin 2010. The same methodology was applied in the recovery of loom weights in the textile workshop in Field V, enabling the study of weaving styles. The work at Tell Halif provides archaeological evidence and context for historical and political events that took place in this region. It appears that, during the LB IB (Stratum X) and LB IIA (Stratum IX), the site was under strong Egyptian influence (for example, see Jacobs and Seger 2007: pl. 4:15). During the succeeding stratum (LB IIB; Stratum VIII), the site served the collection, storage, and most likely dispensation of grain. Throughout the Late Bronze Age, the site of Tell Halif was unfortified (Seger et al. 1990, Seger 1993). It is important to observe that the transition from LB IIB to Iron Age I seems to have occurred peacefully. No destruction level separates the two. In fact, in Field I it appears that the nature of the site did not change and some of the old grain pits continued to be used while new ones were constructed as well. The interpretation of the site during this phase as a collection center for agricultural produce is supported by the presence of grain pits in Fields II and III. The latest work in Field V revealed another grain pit that might belong to the same stratum. Some domestic structures in Field I were used in the LB period and reused in the Iron Age I as well ( Jacobs and Seger 2007). The continuity of culture from the LB to the Iron Age I is evident in the ceramic assemblage. The decline of Egyptian and rise of Philistine influence is also exemplified in the ceramic repertoire; however, Tell Halif did not become a Philistine outpost or stronghold ( Jacobs and Seger 2007: 154–55). Stratum VIB and Stratum VIA, dated to the Iron Age II (800–650 BCE), provide extensive evidence concerning daily life during this period. The sudden and complete destruction of Stratum VIB that is attributed to Sennacherib’s campaign in the Levant in 701 BCE left us copious information related to the economy of the site, trade relations, cultic practices, diet, and more. The materials from Tell Halif suggest connections during this period with areas as far north as Syria and as far south as the Red Sea and the Indian Ocean. Connections were maintained with the Coastal Plain and reached as far as Egypt.

Bibliography Albright, W. F. 1924 “Researches of the School in Western Judaea.” Bulletin of the American Schools of Oriental Research 15: 5–9. Bang, S. H., et al. 2012 A Petrographic Provenance Analysis of Cuboid Limestone Incense Altars from Tell Halif and Their Implications. A lecture at the ASOR Annual Meeting, Chicago. Biran, A., and R. Gophna 1969 An Iron Age Burial Cave at Tel Halif. Eretz Israel 9: 29–39, pls 24–26. Jerusalem: Israel Exploration Society.

Tell Halif in the Late Bronze and Iron Age

113

Borowski, O. 1982 “Four Seasons of Excavations at Tell Halif/Lahav.” Qadmoniot 15: 57–60 (Hebrew). 1988 “The Biblical Identity of Tel Halif.” Biblical Archaeologist 51: 21–27. 1995 “The Pomegranate Bowl from Tell Halif.” Israel Exploration Journal 45: 150–54. 2005 “Tell Halif: In the Path of Sennacherib.” Biblical Archaeology Review 31.3: 24–35. 2008a “Notes and News: Tel Halif, 2007.” Israel Exploration Journal 58: 100–103. 2008b “Tell Halif–2007.” Hadashot Arkheologiyot; Excavations and Surveys in Israel 120 (March 2008). http://www.hadashot-esi.org.il/report_detail_eng.aspx?id=730&mag_id=114 2009 “Notes and News: Tel Halif, 2008.” Israel Exploration Journal 59: 119–22. 2010 “Tell Halif – 2009.” Hadashot Arkheologiyot: Excavations and Survey in Israel 122. http://www.hadashot-esi.org.il/report_detail_eng.aspx?id=1362 2013 Lahav III: The Iron Age II Cemetery at Tell Halif (Site 72). Winona Lake, IN: Eisenbrauns. Borowski, O., and J. Doolittle 1996 A Penetrating Look: An Experiment in Remote Sensing at Tell Halif. Retrieving the Past: Essays on Archaeological Research and Methodology in Honor of Gus W. Van Beek, ed. J. D. Seger. Winona Lake, IN: Eisenbrauns. Borowski, O., and J. D. Seger forthcoming  Lahav VIII: Tell Halif Excavations in Field IIIn 1977–1987. Winona Lake, IN: Eisenbrauns. Dessel, J. P. 1988 “An Iron Age Figurine from Tel Halif.” Bulletin of the American Schools of Oriental Research 269: 59–64. 2009 Lahav I: Pottery and Politics; The Halif Terrace Site 101 and Egypt in the Fourth Millennium B.C.E. Winona Lake, IN: Eisenbrauns. Guerin, V. 1869 Description Geographique, Historique et Archaeologique de la Palestine. Amsterdam: Amsterdam Oriental Press Hadashot Arkheologiyot 1964 “An Iron Age Burial Cave at Kibbutz Lahav.” Hadashot Arkhe’ologiot 11: 24–25. 1972 “Lahav (Tel Halif).” Hadashot Arkhe’ologiot 43: 18. 1974 “Lahav–Tell Halif.” Hadashot Arkheologiot 51: 28–29. Hardin, J. W. 2010 Lahav II: Households and the Use of Domestic Space at Iron II Tell Halif; An Archaeology of Destruction. Winona Lake, IN: Eisenbrauns. Jacobs, P. F. 1987 Tell Halif: Prosperity in a Late Bronze Age City in the Edge of the Negev. Pp. 67–86 in Archaeology and Biblical Interpretation, ed. L. G. Perdue, L. E. Toombs, and G. L. Johnson. Atlanta:, John Knox. 1998 “DigMaster: Cobb Institute of Archaeology.” http://www.cobb.msstate.edu/Research. html. 2015 Lahav IV: The Figurines of Tell Halif. Winona Lake, IN: Eisenbrauns. Jacobs, P. F., and J. D. Seger 2007 Glimpses of the Iron I period at Tell Halif. Pp. 146–55 in “Up to the Gates of Ekron”: Essays on the Archaeology and History of the Eastern Mediterranean in Honor of Seymour Gitin, ed. S. White-Crawford, A. Ben-Tor, J. P. Dessel et al. Jerusalem”, AIAR and IES. Kloner, A. 1984 The Cemetery at Horvat Thala. Eretz-Israel 17: 325–32, 314*; pls. 340–43. Jerusalem: Israel Exploration Society. (Hebrew)

114

Oded Borowski

1993 Rimmon, Horvat. Pp. 1284–85 in vol. 4 of The New Encyclopedia of Archaeological Excavations in the Holy Land. Edited by E. Stern. Jerusalem: Israel Exploration Society and Carta. Levy, T. E., et al. 1997 “Egyptian-Canaanite Interactionat Nahal Tillah, Israel (ca. 4500–3000 B.C.E.): An Interim Report on the 1994–1995 Excavations.” Bulletin of the American Schools of Oriental Research 307: 1–51. Seger, J. D. 1993 Halif, Tell. Pp. 553–59 in vol. 2 of The New Encyclopedia of Archaeological Excavations in the Holy Land. Edited by E. Stern. Jerusalem: Israel Exploration Society and Carta. Seger, J. D. et al. 1990 “The Bronze Age Settlements at Tell Halif: Phase II Excavations, 1983–1987.” BASOR Supplement 26: 1–32. Seger, J. D., and O. Borowski 1977 “The First Two Seasons at Tell Halif.” Biblical Archaeologist 40: 156–66.

The Iron Age City of Khirbet Qeiyafa Yosef Garfinkel

1. Introduction A major research question that has attracted much attention in recent years is to attempt to determine when the Kingdom of Judah spread from the core area of the hill country into the lowlands of the Judean Shephelah. Four major answers to this question have been proposed: 1. The early- to mid-tenth century BCE, the time of the United Monarchy. Accordingly, Level V at Lachish has been attributed to the United Kingdom and its destruction related to the campaign of Pharaoh Shoshenq I (see, for example, Aharoni 1979). 2. The late tenth century BCE, the time of Rehoboam’s fortifications (Garfinkel 2012). 3. The late ninth century BCE, after the destruction of the major Philistine city of Gath (Tell es-Safi). Today, this possibility is espoused by many scholars (see, for example, Koch 2012; Lehmann and Niemann 2014). 4. The late eighth century BCE, after the destruction of the northern Kingdom of Israel. Many refugees fled to the south at that time, enabling the Kingdom of Judah to develop fully (see, for example, Finkelstein 1996).

Needless to say, all of these suggestions are based on historical considerations, and none rest on the solid radiometric dating of relevant archaeological layers. It is also interesting that most of these proposals view the establishment of the Kingdom of Judah as a single, short event, in which the entire kingdom was established in the hill country, the Judean Shephelah, the Beer-sheba Valley, and even the so-called fortresses in the Negev. My suggestion is that the only solution that views the establishment of the Kingdom of Judah as a long, historical and demographic process, with a number of defined stages is no. 2 (Garfinkel 2012). The Khirbet Qeiyafa excavations have altered the debate in major ways. For the first time in the archaeology of the Iron Age Shephelah, a large body of radiometric dates has been presented. Initially, only a few dates were published, but we now have 27 measurements, all made on short-lived samples (mainly olive pits) deriving from a single occupation layer that existed for 20 or 30 years at most. Khirbet Qeiyafa was built around 1000 BCE and was destroyed around 980 or 970 BCE (Garfinkel et al. 2012; Garfinkel et al. 2015). 115

116

Yosef Garfinkel

Fig. 1.  Aerial photograph of Khirbet Qeiyafa at the end of the 2012 excavation season (view to the north).

All the finds from this layer (fortifications, urban planning, local pottery, Cypriot pottery, Egyptian scarabs, inscriptions, art, and cult objects) are dated to 1000–970 BCE. This body of fresh data permits us to take a fresh look at the questions relating to the establishment and expansion of the Kingdom of Judah. Furthermore, building on the data from Khirbet Qeiyafa, we have initiated a new project at Tel Lachish intended to obtain additional data pertinent to this research question. In this essay, I will give a short description of the Iron Age city of Khirbet Qeiyafa and its major finds. For detailed information, the reader is referred to the two final excavation reports published thus far (Garfinkel and Ganor 2009; Garfinkel, Ganor, and Hasel 2014), as well as numerous articles (see, for example, Garfinkel and Ganor 2008, 2012; Garfinkel, Ganor, and Hasel 2010, 2011, 2012; Garfinkel and Kang 2011; Garfinkel and Mumcuoglu 2013). In the discussion, I will focus on the contribution of Khirbet Qeiyafa to the understanding of the early phase of the Kingdom of Judah.

2.  The Site of Khirbet Qeiyafa and the Iron Age City Khirbet Qeiyafa is located in the western part of the upper Shephelah (Israel map grid 14603–12267), on the summit of a hill that borders the Elah Valley on the north. This is a key strategic location in the biblical Kingdom of Judah, on the

The Iron Age City of Khirbet Qeiyafa

117

Fig. 2.  The Iron Age city and the excavation areas at Khirbet Qeiyafa.

main road from Philistia and the coastal plain to Jerusalem and Hebron in the hill country. Even prior to excavation, visitors to Khirbet Qeiyafa could discern a massive city wall, 2–3 m in height, encompassing the summit of the hill. The city wall demarcates an area of 2.3 hectares and its total length is ca. 700 m (figs. 1–2). Due to the local topography, only the external face of the wall is exposed and the inner part is buried under archaeological remains. The base of the city wall is composed of cyclopean stones, some weighing 4–5 tons, while its upper part is built with medium-sized stones. Two city gates had already been located prior to their excavation, one in the south and one in the west.

118

Yosef Garfinkel

Fig. 3.  Aerial photograph of the city wall and gate in Area C (view to the south).

Research into the history of Khirbet Qeiyafa started in the mid-nineteenth century, when the site was reported by the French explorer V. Guerin (1868: 331–32). In the Survey of Western Palestine, Khirbet Qeiyafa was described in only a few words: “heaps of stones” (Conder and Kitchener 1883: 118). In the summary list of Arabic and English names, the site appears in Arabic as Kh. Kiafa, “the ruin of tracking footsteps” (Conder and Kitchener 1881: 308). During the twentieth century, the site was neglected; it is not referred to in the works of the leading scholars in the field of biblical historical geography but is mentioned in a few surveys conducted in the 1980s and 1990s in the Shephelah (Dagan 1993, 1996; Greenhut et al. 2001: 115–17). None of these surveys, however, recognized that the site represented a heavily fortified city of the early tenth century BCE. Seven excavation seasons (2007–2013) were conducted at Khirbet Qeiyafa by Y. Garfinkel and S. Ganor on behalf of the Institute of Archaeology of the Hebrew University of Jerusalem (in 2009–2011, with M. G. Hasel of the Southern Adventist University). Altogether, ca. 5000 square meters were uncovered in six excavation areas (Areas A–F). About 100 m west of the fortified city, a small tower measuring 6 × 6 m was uncovered (Area W). The pottery of this area, including four rosetteimpressed handles, is typical of the seventh and early sixth centuries BCE. Here, however, I will focus only on the early Iron Age city. Due to the shallow accumulation of debris and the massive stone construction, it was possible to uncover a large part of the Iron Age city during a relatively short

The Iron Age City of Khirbet Qeiyafa

119

Fig. 4.  Schematic plan of the two gates and the casemate city wall. Note that the casemate openings are always located away from the gate.

time. The expedition uncovered ca. 20% of the city and unearthed two gates, large open piazzas adjoining the interior of each gate, a casemate city wall, a peripheral belt of houses abutting the city wall, a large storage building, and a major public structure occupying the highest point of the site. The wide exposure allowed us to tackle questions that usually remain unanswered because of lack of data. Two of these questions are: (1) how the city was built (Garfinkel, Ganor, and Hasel 2012) and (2) how the building stones were quarried from nearby locations (Keimer 2014). The fortification system at Khirbet Qeiyafa includes a casemate wall (fig. 3). The outer wall is more massive, about 1.5 m wide, and was built of large stones, sometimes 2–3 m long and weighing up to 8 tons. The inner wall was less massive, about 1 m wide, and was usually constructed from medium-sized stones weighing 100–200 kg. Short perpendicular walls divided the city wall into casemates, and these walls had an average length of ca. 6.5 m. The openings of the casemates are consistently located in the corner that is furthest from the city gate. In Area C, the fifth casemate northeast of the gate is twice the width of an ordinary casemate and has thicker walls. This appears to have been a watchtower, located at a strategic point where the road approaching the city from the Elah Valley was visible. Abutting this tower was a stable, a square structure with three massive stone pillars and two troughs. The location of these two structures in proximity to each other is clearly for functional reasons: if the watch saw the need to move manpower or supplies quickly, animals kept in the stable could be utilized.

120

Yosef Garfinkel

Fig. 5.  Schematic plan of the dwellings in Area C.

Two city gates were uncovered at Khirbet Qeiyafa: the western gate in Area B and the southern gate in Area C, both located at the end of roads leading to the city (fig. 4). The façade of the southern gate is particularly monumental and includes two enormous stones, one on each side. This is the most monumental gate façade excavated to date at any Iron Age city in Judah or Israel. Clearly, the use of a single huge stone rather than three or four smaller ones did not merely stem from engineering considerations related to the strength of the construction but served as a propaganda device, demonstrating the power of the city’s ruler to all who entered it. The two gates are similar to each other in their size, plan, and the pattern of the casemates openings, which are always located away from the gate. Adjacent to the interior of each gate was an open piazza. In this area, the casemate wall was freestanding, and no houses abutted the inner wall. The piazza next to the southern gate is 20 m long and the piazza next to the western gate is 30 m long. The gate piazzas of Khirbet Qeiyafa are noteworthy for an additional feature unknown at other sites: adjacent to each of them is a cultic room.

The Iron Age City of Khirbet Qeiyafa

121

Fig. 6.  A large pillared building abutting the city wall in Area F in the northern area of Khirbet Qeiyafa.

A belt of buildings abutting the city wall and incorporating the nearby casemates as rooms was found in each of the relevant excavation areas. In accordance with our approach of excavating complete architectural units, we uncovered 11 buildings in Areas B, C, and D. Six complete buildings were uncovered in Area C (fig. 5). The uncovering of entire buildings enables us to understand their plan, the size of their rooms, and their spatial organization. Each of these buildings seems to have had an open courtyard, several rooms, a number of casemates, and often a corridor connecting its different parts. In the courtyards, we often found tabuns for cooking, showing that this activity was conducted outdoors. A large, long, tripartite pillared building was uncovered in Area F; it measures 11 × 15 m and has an area of ca. 160 square meters (fig. 6). This is a large storage building of the type that characterized public storage buildings in Iron Age cities (see, for example, Kochavi 1998). Buildings of this kind are indicative of a strong central authority that collected taxes and redistributed them to the relevant part of the population. A large, massive building occupied the highest point of the site, near its center, in Area A. Even after the major damage caused by the construction of a later Byzantine structure, the Iron Age building was preserved to a length of 30 m on its southern edge, between its southeastern and southwestern corners. The walls are two to

122

Yosef Garfinkel

three times wider than those of the regular Iron Age houses uncovered in Areas B, C, and D, a width that indicates a structure about three storeys high. Because this building was also located at the highest point of the site, it made an impressive statement in the city and in the entire regional landscape. It is a clear case of the use of architecture to symbolize political power. This was the central building in the city, apparently the seat of the governor and the local administration. The data presented above clearly indicate that Iron Age Khirbet Qeiyafa was a well-planned city. A pleasing symmetry is evident in the urban layout. The two gates are almost identical: each has a drain on the left of the entrance, next to each is a large open piazza, and the openings of the casemates in the city wall are always located in the corner farthest from the gate. Adjacent to each piazza is a cultic room.

3.  Categories of Major Finds and Trade Connections The sudden destruction and abandonment of the city left a very rich assemblage of objects of daily life on the floors and destruction debris in every room. The three largest categories of finds are pottery, stone artifacts, and metal objects. Hundreds of restorable pottery vessels were found, the best examples of early tenth-century-BCE pottery ever found in Judah. Until our excavations, the pottery typology of the early Iron Age IIA was known from Arad XII, Beersheba VII, Tel Batash IV, and Lachish V. All of these levels suffer from the same fundamental problems (Garfinkel and Kang 2011): they were excavated in rather limited areas without clear architectural contexts, the assemblages consisted mainly of small sherds rather than complete vessels, no radiometric dating is available to enable absolute dates, the sites are characterized by a long sequence of Iron Age strata rather than one clear-cut phase, and the assemblages are usually composed only of diagnostic sherds, since the rest of the pottery was discarded during excavation. Although the analysis of the pottery is not yet complete, I can already point to some of the contributions of this assemblage: 1. A previously unrecognized pottery horizon in the cultural sequence of the southern Levant has been identified (Kang and Garfinkel 2009a; Kang 2013). The Khirbet Qeiyafa assemblage has already helped to define the tenth century BCE in the excavations of Jerusalem (Mazar 2011). The early Iron Age IIA assemblages from Jerusalem and Khirbet Qeiyafa display similarities in cooking pots and in jar handles with finger-impressed handles. 2. The painted pottery of Ashdod Ware type from Khirbet Qeiyafa has enabled us to subdivide this pottery tradition into two groups, earlier (Ashdod I) and later (Ashdod II) (Kang and Garfinkel 2009b). The pottery assemblage from Tell es-Safi, designated “Late Philistine Painted Ware” (Ben-Shlomo, Shai, and Maeir 2004), corresponds to Ashdod II in both its typology and its later date in the second half of the ninth century BCE. 3. The very low percentage of pottery decorated with red slip, together with the even lower percentage of red slip with irregular hand burnish, can serve as a

The Iron Age City of Khirbet Qeiyafa

123

Fig. 7.  Nearly 700 finger-impressed handles of large storage jars were discovered at Khirbet Qeiyafa.

chronological marker, showing that this type of pottery decoration became more popular over time. 4. The mass production of jars with finger-impressed handles (fig. 7) is the beginning of a long tradition in the region of Judah, represented by the jars with lmlk, rosette, Gibeon, “lion,” mh, yhd, and five-pointed star stamps ranging from the eighth century BCE to the Hellenistic period. Because all of the abovementioned groups of storage jars derive from Judah, it seems that the nearly 700 impressed jar handles from Khirbet Qeiyafa are not coincidental and reflect the beginning of this very long tradition. 5. The assemblage includes four Black Juglets, a pottery vessel typical of the Iron Age II (Cohen-Weinberger and Panitz-Cohen 2014). 6. The Cypriot juglets from Khirbet Qeiyafa (fig. 8) connect the site with the Mediterranean trade and have far-reaching implications for the dating of these juglets at other sites (Gilboa 2012; Gilboa and Waiman-Barak 2014).

More than 200 stone artifacts were found. They were made from hard limestone, soft limestone, chalk, basalt, beach rock, flint, and other minerals. A few fragments of small alabaster vessels were discovered as well, probably indicative of trade relations with Egypt. Basalt is not a local raw material in the Judean Shephelah and had to have been imported from volcanic deposits more than 100 km from the site. Basalt and alabaster artifacts have not been reported in the Iron Age I

124

Yosef Garfinkel

at sites such as Giloh, Shiloh, Khirbet Raddana, and ʿIzbet Sartah. The appearance of basalt grinding tools at Khirbet Qeiyafa, including quite heavy items, attests to an intensification of economic activities in the early tenth century BCE. More than 30 iron and bronze tools, mainly weapons, were uncovered; they include swords, knives, arrowheads, spearheads, and one bronze axe (fig. 9). Two pottery crucibles with bronze slag were found as well, indicating that metal was smelted on site. The dominant use of iron rather than bronze should not be overlooked. As indicated by Gott­ lieb (2010), in the late eleventh and early tenth centuries BCE, two traditions can be found side by side: some sites yield mainly bronze items, while iron artifacts are rare, while at other sites, the opposite is true. The first group includes northern sites such as Fig. 8.  A small barrel-shaped juglet of imported Megiddo, Taanach, Beth-Shean, and Cypriot Black-on-White ware. Hazor, where the bronze metalwork continues the earlier Canaanite tradition. At southern sites such as Arad and Beer-sheba, iron is dominant. An exception is Tel Masos; despite its location in the Beer-sheba Valley, the metal assemblage of this site is dominated by bronze artifacts. The metal assemblage at Khirbet Qeiyafa clearly resembles those from Arad and Beersheba. In contrast to these small Iron Age I sites in the hill country, at Khirbet Qeiyafa, we have extensive information about trade on three different levels: 1. Local, regional trade: this category includes trade with the nearby Philistine coastal plain. Petrographic analysis has shown that the Ashdod Ware pottery came from Philistia. In the same way, grinding stones made from beach rock originated on the Mediterranean coast. 2. Interregional trade within the southern Levant: this category includes basalt grinding tools and copper. 3. International trade: this category includes two Cypriot juglets, Egyptian scarabs and amulets, alabaster vessels, tin for the bronze industry, and a miniature gold leaf.

The Iron Age City of Khirbet Qeiyafa

125

Fig. 9.  A rich collection of metal objects has been found at Khirbet Qeiyafa, including various weapons: iron daggers, iron swords, a bronze axe, and bronze arrowheads.

The large variety of objects imported to Khirbet Qeiyafa points to trade connections with a rather wide geographical area, from the west (Ashdod Ware and beach rock), north (basalt), and south (copper). Since Khirbet Qeiyafa did not have access to a port, the Cypriot and Egyptian artifacts could have been transported through one of the Philistine harbors (Ashdod or Ashkelon). As with all trade, the immediate question is what was given in return. What goods could the inhabitants of Khirbet Qeiyafa have provided to the inhabitants of the Philistine coastal plain? It seems to us that timber may have been a major commodity, because it was in short supply in Philistia: the sandy coastal plain is not an appropriate ecological zone for trees. The large population of the sites of Ashdod, Ashkelon, and Tell es-Safi would have needed timber for construction, cooking, and heating. This would create a motivation for regional trade connections between Philistia and the Judean Shephelah.

4.  The Contribution of Khirbet Qeiyafa’s Absolute Dating to Iron Age Cultural Phasing The transition between the Iron Age I and II is currently debated. The debate concentrates on two different aspects that have traditionally been linked: the first appearance of typically Iron Age IIA pottery assemblages and the beginning of urbanism in the Kingdom of Judah. The traditional view of this transition, now

126

Yosef Garfinkel

designated the “high chronology,” dates it to ca. 1000 BCE. Advocates of a low chronology place the end of the Iron Age I at ca. 920 BCE, and an ultra-low chronology dates it as late as ca. 900 BCE (for review of the matter, see Garfinkel et al. 2012; Garfinkel et al. in press). One of the benefits of the chronological debate is the recognition that the Iron Age IIA was a rather long phase that occupied the tenth and ninth centuries BCE rather than only the tenth century, as previously believed. The subdivision of this long phase has occupied the attention of several scholars. Herzog and Singer-Avitz (2004) propose that the Iron Age IIA in the south should be subdivided into two phases. They place Arad XII, Beersheba VII, Lachish V, Tel Batash IV, and Tel Masos II in the early Iron Age IIA. These levels are not true fortified cities but enclosures with adjoining houses arranged along the periphery of the site. Herzog and SingerAvitz argue that the first fortified cities were constructed only in the late Iron Age IIA, approximately in the mid-ninth century BCE, citing Arad XI, Beersheba VI, and Lachish IV in this context. Contrary to this conclusion, the pottery assemblage of Khirbet Qeiyafa resembles those of the earlier group of sites but is associated with a true fortified city. Khirbet Qeiyafa, with its massive fortification system, shows that the construction of cities in Judah started in the very early Iron Age IIA. The newly recognized pottery assemblage of Khirbet Qeiyafa enables us to pre­ sent a new subdivision of the Iron IIA Age in Judah into three phases (Garfinkel 2011). 1. The pottery assemblage identified so clearly at Khirbet Qeiyafa was also found at other settlements in the Judean Shephelah or the hill country, such as BethShemesh 4, Arad XII, and Beersheba VII. It marks the beginning of a new period in the history of Judah. Some of the sites, like Khirbet Qeiyafa, were fortified, but others in this region, like Beth-Shemesh 4, Arad XII, and Beersheba VII, were still unwalled villages in this phase, while many others, such as Lachish and Tell Beit Mirsim, were completely uninhabited. The main characteristics of this early phase are infrequent red slip and irregular hand-burnish, early Ashdod Ware, imported Cypriot White Painted vessels, and archaic (Canaanite) script.   In the extensive surveys conducted in the Judean Shephelah, virtually no sites of the early Iron Age IIA were observed (Dagan 1993, 1996; Lehmann 2003). The same picture was reported from various other surveys conducted in the hill country, creating the mistaken impression that Judah was an empty land during the tenth and ninth centuries BCE and became a full-blown state only in the late eighth century BCE. Finkelstein did not hesitate to present a bold picture, in which the settlements consisted only of “500 people with sticks in their hands shouting and cursing and spitting” (Draper 2010, quoting Finkelstein).   If the surveys failed to recognize a city in which the city wall and two gates were still standing, we can see how this period has been overlooked in the large number of sites where its levels are buried under later occupations. This is a clear case of circular reasoning: because it was poorly-known, the tenth century BCE was not recognized in the surveys, and hence the conclusion that no sites of the period existed in these regions was reached.

127

The Iron Age City of Khirbet Qeiyafa

2. In the second phase, in the late tenth and early ninth centuries BCE, additional settlements were built, including Beth-Shemesh 3, Lachish V, and Tel Zayit. This phase is characterized by irregular hand-burnish on bowls (sometimes in geometric patterns), imported Cypriot Black-on-Red vessels, and early Phoenician-Hebrew script (at Tel Zayit). 3. In the third phase, in the second half of the ninth century BCE, fortified cities were gradually established in more southerly sites as well, such as Lachish IV, Tell Beit Mirsim, Beersheba, Tel ʿIra, and Arad. The most representative ceramic assemblages for this phase are those of Tell es-Safi IV and Lachish IV. This phase is characterized by large amounts of red slip and irregular hand burnish, together with late Ashdod Ware.

Table 1.  Division of the Iron Age IIA in Judah and the Shephelah into three chronological phases and the prominent characteristics of each phase (Garfinkel 2011)

Cultural phase within Iron Age IIA

Cultural characteristics

Sites

Late eleventh/early tenth century BCE

Infrequent red slip and irregular hand burnish; archaic (Canaanite) script; import of Cypriot White Painted vessels; early Ashdod Ware

Khirbet Qeiyafa, Khirbet ed-Dawwara, BethShemesh 4, Arad XII, Beersheba VII

Second half of tenth century–early ninth century BCE

Irregular hand burnish on bowls, sometimes in geometric patterns; early Phoenician-Hebrew script; import of Cypriot Black-on-Red vessels

Beth-Shemesh 3, Lachish V, Tel Zayit

Mid to late ninth century BCE

Very common red slip and irregular hand burnish; late Ashdod Ware

Tell es-Safi IV, Lachish IV

This basic subdivision has recently been accepted by Katz and Faust (2014) and has helped them to analyze the pottery assemblage from Tomb C3 at Tel ʿEton. As is usual with rich burial caves, this assemblage represented a long sequence lacking refined stratigraphic divisions and radiometric dates. The recognition of the clearcut pottery phase of Khirbet Qeiyafa has made it possible to discern a tripartite division of the Iron Age II in other sites.

5. Discussion The central question regarding Khirbet Qeiyafa is its relationship with the biblical text, which describes state-formation processes in Judah, King David’s activities, and intensive military clashes against the Philistine city of Gath in the Elah Valley. These biblical traditions are contemporaneous with the settlement of the fortified city at Khirbet Qeiyafa. Thus, our excavations have direct implications for these complex matters. If Khirbet Qeiyafa was a Philistine or Canaanite city or belonged to the Kingdom of Israel, it cannot be connected with the traditions about David and state formation in Judah (on such interpretations, see Naʾaman 2008, 2012; Finkelstein 2013; Lehmann and Niemann 2014). If, however, Khirbet Qeiyafa was a

128

Yosef Garfinkel

Judean city, it is of crucial importance for the history of David and state-formation processes in Judah (Garfinkel 2011; Garfinkel, Ganor, and Hasel 2011). When evaluating the various possible interpretations of Khirbet Qeiyafa, the following distinctive components should be considered: 1. The site was built according to the typical Judean urban plan, a plan that is not found in any city in the Kingdom of Israel. 2. There are nearly 700 impressed jar handles, a typically Judean administrative device. Impressed jar handles are not found in the Kingdom of Israel in meaningful quantities. 3. The site did not yield the figurines that are characteristic of sites in the Kingdom of Israel in this period. 4. Five early alphabetic (Proto-Canaanite) inscriptions are known today from the tenth century BCE: three from Khirbet Qeiyafa, one from Beth-Shemesh, and one from Jerusalem. These sites are located in the core area of Judah. Not a single inscription of this kind has been found in sites of the Kingdom of Israel. 5. The dominance of iron tools in the assemblage of metal objects is characteristic of Judean sites; in the Kingdom of Israel, bronze was dominant at this time. 6. The site’s location in the Elah Valley on the main road from the Philistine centers of Ashdod and Ashkelon to Jerusalem had no geopolitical importance for the Kingdom of Israel. In order to defend its supposed territory from Philistine attacks, the northern kingdom would have needed to build fortified cities in the Sorek and Ayalon Valleys.

The material culture of Khirbet Qeiyafa does not accord with the characteristics of the Philistine city-states of the coastal plain and lower Shephelah, the Canaanites, or the Kingdom of Israel. On the other hand, all of these aspects fit the Kingdom of Judah very well. Indeed, those who have suggested that Khirbet Qeiyafa could be a site belonging to the Philistines, the Canaanites, or the Kingdom of Israel have not discussed its material culture. One should also be aware that the finds at Khirbet Qeiyafa do not imply that a large kingdom extending all the way to Megiddo or Hazor existed. In this phase, even Beersheba and Arad were not yet fortified. Thus, Khirbet Qeiyafa represents the first stage in the development of the Kingdom of Judah, which was at that time a relatively small political unit.

References Aharoni, Y. 1979 The Land of the Bible: A Historical Geography. Translated and edited by A. F. Rainey. Philadelphia: Westminster. Ben-Shlomo, D.; Shai, I.; and Maeir, A. M. 2004 Late Philistine Decorated Ware (“Ashdod Ware”): Typology, Chronology, and Production Centers. Bulletin of the American Schools of Oriental Research 335: 1–34.

The Iron Age City of Khirbet Qeiyafa

129

Cohen-Weinberger, A., and Panitz-Cohen, N. 2014 The Black Juglets. Pp. 403–14 in Khirbet Qeiyafa, Vol. 2: Excavation Report 2009–2013: Stratigraphy and Architecture (Areas B, C, D, E), ed. Y. Garfinkel, S. Ganor, and M. G. Hasel. Jerusalem: Israel Exploration Society. Conder, C. R., and Kitchener, H. H. 1881 The Survey of Western Palestine: Arabic and English Name List. London: Palestine Exploration Fund. 1883 The Survey of Western Palestine. London: Palestine Exploration Fund. Dagan, Y. 1993 Bet Shemesh and Nes Harim Maps, Survey. Excavations and Surveys in Israel 13: 94–95. 1996 Cities of the Judean Shephelah and Their Division into Districts Based on Joshua 16. Eretz-Israel 25: 136–46 (Hebrew). Draper, R. 2010 Kings of Controversy. National Geographic, December: 67–91. Finkelstein, I. 1996 The Archaeology of the United Monarchy: An Alternative View. Levant 28: 177–87. 2013 The Forgotten Kingdom: The Archaeology and History of Northern Israel. SBL Ancient Near East Monographs 5. Atlanta: Society of Biblical Literature. Garfinkel, Y. 2011 The Davidic Kingdom in Light of the Finds at Khirbet Qeiyafa. City of David Studies of Ancient Jerusalem 6: 13*–35*. 2012 The Settlement History of the Kingdom of Judah from its Establishment to its Destruction. Cathedra 143: 7–44 (Hebrew). Garfinkel, Y., and Ganor, S. 2008 Khirbet Qeiyafa: Shaʿarayim. Journal of Hebrew Scriptures 8, Article 22, http://www​ .jhsonline.org/Articles/article_99.pdf 2009 Khirbet Qeiyafa, Vol. 1: Excavation Report 2007–2008. Jerusalem: Israel Exploration Society. 2012 Cult in Khirbet Qeiyafa from the Iron Age IIa: Cult Rooms and Shrine Models. New Studies in the Archaeology of Jerusalem and its Region 6: 50–65 (Hebrew). Garfinkel, Y., and Kang, H. G. 2011 The Relative and Absolute Chronology of Khirbet Qeiyafa: Very Late Iron Age I or Very Early Iron Age IIA? Israel Exploration Journal 61: 171–83. Garfinkel, Y., and Mumcuoglu, M. 2013 Triglyphs and Recessed Doorframes on a Building Model from Khirbet Qeiyafa: New Light on Two Technical Terms in the Biblical Descriptions of Solomon’s Palace and Temple. Israel Exploration Journal 63: 135–63. Garfinkel, Y.; Ganor, S.; and Hasel, M. G. 2010 The Contribution of Khirbet Qeiyafa to our Understanding of the Iron Age Period. Strata: Bulletin of the Anglo-Israeli Archaeological Society 28: 39–54. 2011 Khirbet Qeiyafa Excavations and the Rise of the Kingdom of Judah. Eretz-Israel 30: 174–94 (Hebrew). 2012 The Iron Age City of Khirbet Qeiyafa after Four Seasons of Excavations. Pp. 149–74 in The Ancient Near East in the 12th–10th Centuries BCE: Culture and History, Proceedings of the International Conference Held at the University of Haifa, 2–5 May, 2010, ed. G. Galil et al. Alter Orient und Altes Testament 392. Münster: Ugarit-Verlag. 2014 Khirbet Qeiyafa, Vol. 2: Excavation Report 2009–2013: Stratigraphy and Architecture (Areas B, C, D, E). Jerusalem: Israel Exploration Society.

130

Yosef Garfinkel

Garfinkel, Y.; Streit, K.; Ganor, S.; and Hasel, M. G. 2012 State Formation in Judah: Biblical Tradition, Modern Historical Theories and Radiometric Dates at Khirbet Qeiyafa. Radiocarbon 54: 359–69. Garfinkel, Y.; Streit, K.; Ganor, S.; and Reimer, P. J. 2015 King David’s City at Khirbet Qeiyafa: Results of the Second Radiocarbon Dating Project. Radiocarbon 57/5: 881–90. Gilboa, A. 2012 Cypriot Barrel Juglets at Khirbet Qeiyafa and Other Sites in the Levant: Cultural Aspects and Chronological Implications. Tel Aviv 39: 133–49. Gilboa, A., and Waiman-Barak, P. 2014 Cypriot Ceramic Imports at Khirbet Qeiyafa: Provenience, Chronology and Significance. Pp. 391–402 in Khirbet Qeiyafa, Vol. 2: Excavation Report 2009–2013: Stratigraphy and Architecture (Areas B, C, D, E), ed. Y. Garfinkel, S. Ganor, and M. G. Hasel. Jerusalem: Israel Exploration Society. Gottlieb, Y. 2010 The Advent of the Age of Iron in the Land of Israel: A Review and Reassessment. Tel Aviv 37: 89–110. Greenhut, Z.; Strul, L.; Bardah, L.; and Weiss, D. 2001 Jerusalem District Master Plan 30/1: Archaeological Appendix. Jerusalem: Ministry of the Interior (Hebrew). Guérin, V. 1868 Description géographique, historique et archéologique de la Palestine. Paris: Imprimé par autorisation de l’empereur à l’Impr. Herzog, Z., and Singer-Avitz, L. 2004 Redefining the Center: The Emergence of State in Judah. Tel Aviv 31: 209–44. Katz, H., and Faust, A. 2014 The Chronology of the Iron Age IIA in Judah in the Light of Tel ʿEton Tomb C3 and Other Assemblages. Bulletin of the American Schools of Oriental Research 371: 103–27. Kang, H. G. 2013 Pottery Assemblage of Khirbet Qeiyafa and its Implications for Understanding the Early 10th Century BCE in Judah. Unpublished Ph.D. Dissertation, Hebrew University of Jerusalem. Kang, H. G., and Garfinkel, Y. 2009a Ashdod Ware I: Middle Philistine Decorated Ware. Pp.  151–60 in Khirbet Qeiyafa, Vol. 1: Excavation Report 2007–2008, ed. Y. Garfinkel and S. Ganor. Jerusalem: Israel Exploration Society. 2009b The Early Iron Age IIA Pottery. Pp. 119–50 in Khirbet Qeiyafa, Vol. 1: Excavation Report 2007–2008, ed. Y. Garfinkel and S. Ganor. Jerusalem: Israel Exploration Society. Keimer, K. H. 2014 Iron Age Stone Quarries. Pp.  333–45 in Khirbet Qeiyafa, Vol. 2: Excavation Report 2009–2013: Stratigraphy and Architecture (Areas B, C, D, E), ed. Y. Garfinkel, S. Ganor, and M. G. Hasel. Jerusalem: Israel Exploration Society. Koch, I. 2012 The Geopolitical Organization of the Judean Shephelah during Iron Age I–IIA. Cathedra 143: 45–64 (Hebrew). Lehmann, G. 2003 The United Monarchy in the Countryside: Jerusalem, Judah and the Shephelah during the 10th Century B.C.E. Pp.  117–64 in Jerusalem in Bible and Archaeology, ed. A. G. Vaughn and A. E. Killebrew. Atlanta: Society of Biblical Literature.

The Iron Age City of Khirbet Qeiyafa

131

Lehmann, G. and Niemann, H. M. 2014 When Did the Shephelah become Judahite? Tel Aviv 41: 77–94. Mazar, E. 2011 Discovering the Solomonic Wall in Jerusalem: A Remarkable Archaeological Adventure. Jerusalem: Shoham Academic Research and Publication. Naʾaman, N. 2008 In Search of the Ancient Name of Khirbet Qeiyafa. Journal of Hebrew Scriptures 8, Article 21, http://www.jhsonline.org/Articles/article_98.pdf. 2012 Khirbet Qeiyafa in Context. Ugarit-Forschungen 42: 497–526.

This page left intentionally blank.

Philistine Gath after 20 Years: Regional perspectives on the Iron Age at Tell eṣ-Ṣafi/Gath Aren M. Maeir

Introduction The ongoing archaeological project at Tell eṣ-Ṣafi/Gath, now at the beginning of its third decade in the field, is in the envious position of being one of the longestrunning, ongoing projects currently working in the region of the Shephelah (fig. 1). As such, we have benefited in our research from the various projects that have worked in this and adjacent regions before us and alongside us, and most importantly, from the “boom” of projects that have commenced in the Shephelah in the last few years. This situation enables current research, at Tell eṣ-Ṣafi/Gath specifically and the region in general, to move beyond “the basics”—such as the rudimentary study of the regional material culture—and to deal with various, and in many cases, broader issues. In the following pages, I do not intend to recapitulate the many finds and results from the Iron Age that have been found at Tell eṣ-Ṣafi/Gath since the beginning of the project (see, e.g., Maeir 2012a; 2012c; 2013a; Maeir and Uziel in press) but rather to utilize this platform to discuss a choice set of specific finds and issues that relate to the site itself, as well as to consider wider issues that are of relevance to the understanding of central issues in Iron Age archaeology in the Shephelah and beyond.

The Development of the Philistine Culture and Its Ramifications Study of the Philistine culture in the last decade or so has brought about quite a major change in the understanding of the Philistines and their culture. This includes major revisions on aspects such as: who the Philistines were; the origins of their culture and of the populations in Iron Age Philistia; the sociocultural processes that occurred during the Iron Age in Philistia; the character of the relations between the Philistines and neighboring Levantine cultures; and more. By and large, these new conceptual frameworks are based on the finds from the excavations in Philistia and surrounding regions and, in particular, from those of 133

134

Aren M. Maeir

recent, state-of-the-art excavations. The finds from Tell eṣ-Ṣafi/Gath and their analyses have contributed substantially to a better understanding of many of these issues (and by and large, have been extensively published elsewhere). Briefly, I would like to mention some of the more interesting insights that have developed out of our excavations at Tell eṣ-Ṣafi/Gath and discuss some of the issues relevant for the Iron Age Shephelah in general. While many researchers unsuccessfully have searched for the homeland of the Philistines in this or that specific region, cumulative research on early Philistine culture indicates that one cannot speak of a single place of origin; rather, the early Philistine population is comprised of peoples of mixed origins, from various places in the eastern and central Mediterranean, who settled among local Canaanites. This is exemplified by many different cultural influences seen in early Iron Age Philistia—not Mycenaean, Cypriote, Anatolian, or a single other cultural origin, but a combination of many (e.g., Maeir, Hitchcock, and Horwitz 2013; Frumin et al. 2015). We have argued that early Philistine culture should be seen as an “entangled” culture (e.g., Hitchcock and Maeir 2013; 2014; 2017; Davis, Maeir, and Hitchcock 2015; Maeir and Hitchcock 2016; in press; Maeir, Davis, and Hitchcock 2016. In addition, the sociocultural background of the Philistines is far from clear, and the circumstances under which the Philistines appeared, in the general context of the terminal Late Bronze Age Mediterranean world, are simply not known. We have suggested (e.g., Hitchcock and Maeir 2014; 2017) that major groups of early Philistine peoples may have been pirate-like in nature, in the context of the collapse of the Late Bronze Age system in the late 13th and early 12th centuries BCE. Thus, I can only reiterate that close attention should be paid to the complex cultural, ethnic, and geographic origins of the Philistines, and the reflection of these origins in the processes, influences, and mechanisms seen in early Iron Age Philistia and surroundings. Otherwise, simplistic monolithic explanations of the cultural dynamics during this time in this and surrounding regions will still be proposed.

On a Supposed “Canaanite Enclave” in the Iron I Shephelah In sundry recent discussions on the early Iron Age of southern Canaan, and on Philistia and the Shephelah in particular, it has been suggested that a distinct Canaanite entity existed in the Shephelah, situated between the Philistines to the west and the Israelites on the east. As noted below, I believe this view is not without problems (for a more extensive discussion on this issue, see Maeir and Hitchcock 2016). The very question of how to identify a site as being associated with the Philistine culture and, even more basically, how the various levels of “Philistine identity” can be archaeologically defined, have been extensively discussed. Some of these attempts to differentiate between the “Philistines” and other ethnicities, based on a small set of material correlates, have led to simplistic or simply mistaken

Philistine Gath after 20 Years

135

differentiations. As noted in the past, many of these cultural attributes can appear on both sides of the supposed Philistine/Israelite ethnic boundaries—and even beyond (Hitchcock and Maeir 2013; Maeir, Hitchcock, and Horowitz 2013). There is no doubt that the material assemblages at major sites in Iron Age Philistia are different from those of sites in regions associated with other groups (Israelite, Judahite, Phoenician, etc.). Nonetheless, specific types of objects can be seen in many areas and used by many groups (such as pottery types appearing in different cultural areas; see, e.g., Ben-Shlomo et al. 2008). The appearance of supposedly Philistine objects should not be seen as necessarily indicating the expansion of the Philistine culture into other zones and, similarly, the appearance of Israelite/Judahite facets among the Philistines. Instead, artifact assemblages should be examined in their broader contexts in order to draw out various cultural encounters, relationships, and entanglements, as well as to elucidate new ones (e.g., Ross 2012). Philistine cultural identity is often seen as set in binary opposition to Israelite group identity, with this “otherness” as a major impetus for the formation of Israelite identity (e.g., Faust 2013a; 2013b; 2015a; 2015b; 2015c; Faust and Lev-Tov 2011; 2014). But a straightforward connection between the ideologies reflected in the biblical text and the archaeological record is problematic. To start with, most of the assumptions regarding the antagonistic relationship between the Philistines and the “Israelites” are based solely on the biblical text—and it is not clear how much of this represents an early Iron Age reality and what part of this is a reflection of later Iron Age, or even post-Iron Age, realities and ideologies. To this one can add that the very assumption that one can define and relate to a singular “Israelite” identity, which defined the Philistines as the “other” in the early Iron Age, is far from certain (see now Ehrlich 2016 on the “fuzziness” of biblical gentilics). With this in view, the recent suggestion that there was a Canaanite enclave in the Shephelah should be assessed. Various scholars (e.g., Jasmin 2006: 227–28; Bunimovitz and Lederman 2009; 2011; Naʾaman 2010; Faust and Katz 2011; 2015; Faust 2013b; 2015c; Lederman and Bunimovitz 2014) believe that not only can one identify Philistine and Israelite/Judahite ethnicities, but they also believe that there is an additional “Canaanite” group in the Shephelah, located between the Philistines and Israelites. For example, the excavators of Beth-Shemesh have identified a process of “resistance” among the local “non-Philistine” population at the site. They suggest that, with the arrival of the Philistine migrants, the inhabitants of Beth-Shemesh stopped eating pork and ceased using decorated Philistine pottery in what they believe is an act of resistance (Bunimovitz and Lederman 2009; 2011; Lederman and Bunimovitz 2014). Although there may be some basis for these suggestions, several cautionary notes are in order. (a) Philistine decorated pottery has been found at Beth-Shemesh (e.g., Münnich 2013; as well as in the recent excavations [Lederman pers. comm.]); (b) The Philistines themselves were comprised of both foreign and local Canaanite components.

136

Aren M. Maeir

Attempts to formulate definitions of the various ethnic groups as living in clearly demarcated regions with unambiguous borders are even more difficult to accept (e.g., Naʾaman 2010; Faust and Katz 2011; Faust 2013b; 2015a; 2015c; 2015d). The definition of “who is” and “who is not” a Philistine or an Israelite/Judahite is hardly agreed upon (see Maeir and Hitchcock 2016), and attempts to define the supposedly static ethnic identity of a group situated in the contact zone between these groups is problematic. Because “Canaanite” (local Levantine) features are seen in Iron Age Philistia, and much of the so-called “early Israelite” culture can be traced to local Levantine (“Canaanite”) origins, the existence of a third, unique “Canaanean” entity is problematic to posit. Faust (2015c) has suggested that the relations between the Philistines and the Shephelah Canaanites reflects a colonial-like relationship. Accordingly, the Philistines, arriving in Canaan in the early Iron Age and forcibly occupying Philistia and destroying the Canaanite cities, were overlords of the Canaanites. He believes that, in the early Iron I, the Canaanites in the Shephelah did not use emblematic Philistine cultural items due to boundary definition. Later on, though, in the middle-to-late Iron I, the Canaanites, and in particular their elites, started to use certain classes of Philistine material culture, as a result of classic interaction between colonizer and colonized peoples. This suggestion is likewise beset with problems. To start with, it seems to be based on an outmoded understanding of the processes involved in the initial stages of the development of the Philistine culture. Although traditional suggestions considered the appearance of the Philistine culture as the result of a clear-cut invasion of a foreign group (or groups), recent studies (e.g., Yasur-Landau 2010; Maeir and Hitchcock 2016; in press; Hitchcock and Maeir 2013; 2014; 2017; Cline 2014) reflect a very different picture. The Philistines are seen as a mixed, entangled sociocultural entity deriving from various foreign and local Levantine groups. Likewise, the mechanisms through which the Philistine culture appeared are quite complex, including, inter alia, collapse of the Mediterranean Late Bronze Age “world order” and the appearance of pirate-like groups in the eastern Mediterranean (Hitchcock and Maeir 2014; 2017). To this one can add that there is very little evidence of substantial destruction at the Canaanite sites in the southern Coastal Plain, despite Faust’s (2015c: 215) claims to the contrary. It should be stressed as well that there is very little evidence of warfare-related material in the Iron I Philistine culture—which belies a view of the Philistines as a conquering and dominating colonial culture. Although colonialism encompasses a very broad range of cultural interactions (see Gosden 2004; van Dommelen 2012), colonial relationships require, condicio sine qua non, domination by one party of the other (Horvarth 1972; Osterhammel 2005; Jordan 2009; Kohn 2011; Ypi 2013: 162; Steinmitz 2014: 79–80; Loomba 2015: 20)—and this may very well not be the case with the Philistines’ relations with neighboring cultures.

Philistine Gath after 20 Years

137

Clearly then, because it is not at all certain that the Philistines maintained a colonial dominance over the Shephelah; the conclusion that relations between the Philistines and the peoples of the Shephelah reflects a colonial situation rests on shaky ground. A central issue on which this entire question hinges is how the ethnic identity of a population may be identified. Because this topic has been dealt with in regard to the “Canaanite enclave” in detail (Maeir and Hitchcock in press a), I will touch on it only briefly here. The suggestion that an ethnic group may be defined on the basis of a list of archaeologically identified “markers,” which then can serve to define geographical boundaries (e.g., Bunimovitz and Faust 2001; Faust 2015a; 2015b; 2015c; 2015d; Faust and Lev-Tov 2011, 2014; Faust and Katz 2011), often based on Barth’s (1969) seminal contribution to the study of ethnicity, needs to be carefully reviewed (Maeir and Hitchcock in press). Are not the very definitions of ethnic groups presumed to appear in Iron I an ideological reflection derived from later texts? Are we sure that there were distinct Philistine and Israelite ethnicities in the early Iron Age—or were there several groupings, some of them somewhat hard to define? And even if these groups did exist, are the suggested archaeological markers (such as pig consumption, pottery, etc.) in fact “boundary markers” for the different groups? As noted in previous studies (Maeir, Hitchcock, and Horowitz 2013; Maeir and Hitchcock in press), the appearance of some of these supposed markers in early Iron Age Philistia itself raises questions regarding the validity of simplistically using such markers. The fluidity and rapidly changing nature of ethnic identity is well known (e.g., Hall 2000; Malkin 2001; Dougherty and Kurke 2003; Casella and Fowler 2005; Siapkas 2014). Conclusions regarding the identities of groups during the early Iron Age that are based, by and large, on written sources from later periods may reflect the social and/or ideological environment of these later periods, instead of the Iron Age. Thus, it is far from clear that distinct, archaeologically visible ethnic identities during the early Iron Age can be identified, when, in fact, a heterogeneous and constantly changing matrix of identities might have existed at the time. 1 As previously suggested (Maeir and Hitchcock 2016), the identification of the “Canaanite enclave” might be influenced by modern reception of the biblical text, in particular the “Tamar and Judah narrative” in Genesis 38, in which the Canaanites are located in the Shephelah region. On the one hand, there is no textual corroboration of this in contemporaneous Iron Age texts. And on the other hand, very few biblical scholars would date Genesis 38 to the early Iron Age (e.g., Leuchter 1. While Garfinkel, Kreimerman, and Zilberg (2016: 173–86) at first appear to have a skeptical view of identifying ethnicities in the Iron Age Shephelah (and criticize the identification of a Canaanite entity in this region), in the end they opt for a somewhat essentialist interpretation, at least as concerns the identity of the inhabitants of Qeiyafa (as Judahites).

138

Aren M. Maeir

2013; in general, see Uehlinger 1999; 2000). Thus, it is far from clear that this text reflects a historical reality. Although I fully accept the existence of what might be termed a Canaanite “identity” (or better, identities) during the Late Bronze Age, this does not mean that this identity continued into the Iron Age, a period when demographic, technological, and sociopolitical structures were in flux. This claim would require the assumption that the groups living in the border zone between the Coastal Plain and the Central Hills, where later sources located the Philistines on the one hand and Israelites/Judahites on the other, retained the cultural and ethnic identities they had during the Late Bronze Age. This, however, cannot be taken for granted. Identities can change quickly, and unless there is explicit evidence, there is no reason to assume that Late Bronze Age identity packages endured throughout such an extended period of social and cultural upheaval. Although there is some continuity in some aspects of the material culture between the Late Bronze Age and early Iron Age, this does not mean that there was a continuity in Canaanite (or any other) group identity. Nestor (2010; and see Eriksen [2010: 213–14] regarding other contexts) rightfully cautions against simplistic attempts to demonstrate explicit continuity between Israelite identity in Iron Age I and Iron Age II; and even more so, in the case of the Canaanite identity in the Shephelah during the early Iron Age, where we are completely dependent on scholarly assumptions, a cautious and skeptical approach is warranted. While the theoretical possibility of the existence of a “Canaanite identity” in the Shephelah cannot be denied, this is nothing more than a modern postulate! Because both the Philistines and the Israelites/Judahites had substantial “Canaanite” components in their formative stages (Maeir and Hitchcock 2016), how can “Philistine Canaanites,” “Real Canaanites” (supposedly living in the Shephelah), and “Israelite Canaanites” be distinguished? As already suggested (Maeir and Hitchcock 2016; in press), it would be preferable to look at the transitions between the Philistia and the Shephelah and the latter and the Central Hills as regions in which boundaries did exist but were “fuzzy” and constantly changing (Gardner 2007). There is no question that during the early Iron Age there were peoples who identified themselves distinctly—perhaps as “Philistines” (and they resided mainly in Philistia) and as Israelites/Judahites (and they resided mainly in the Central Hills); and for the argument’s sake, perhaps there were people who self-identified as “Canaanites” (residing in the Shephelah). Nonetheless, it would be very hard to determine on the basis of the archaeological evidence, at any given time, the cultural/ethnic affiliation of any specific group, let alone the exclusive, or even static, group identity of the inhabitants of a specific site in the border zones (see now Lehmann and Niemann 2014 and Mazar 2014: 362–64 for attempts to deal with the fluidity of cultural identities in the early Iron Age Shephelah).

Philistine Gath after 20 Years

139

What might have existed are overlapping “micro-identities” (see Whitmarsh 2010; Poblome, Malfitana, and Lund 2014), in which intensive cultural “codeswitching” (following van Nijf 2010) occurred in the Iron I Shephelah, accompanied by regular switching between emblematic identifying facets, depending on specific contexts and needs. Similarly, perhaps identities in the Shephelah region were “nested identities,” in which identities (including, but not only, ethnicity) operated simultaneously at different levels. 2 Thus, there may have been overlapping identities in use simultaneously that were, over time, reflected in various ways in the archaeological record. The suggestion that changes in material culture and related differences directly reflect different ethnic identities is entirely too simplistic. As previously suggested (Maeir and Hitchcock 2016, in press; see now also Faust 2015c), viewing the Shephelah as a “Middle Ground” (White 1991; Woolf 2011; Reger 2014) or a “Third Space” (Bhabba 1994) may be insightful. In this perspective, the material culture “packages” identified in this region perhaps reflect the “Social Imaginaries” (Castoriadis 1975; Taylor 2002; Strauss 2006; Stavrianopoulou 2013) in a region that had intense intercultural contacts (see Mengoni 2010). These “social imaginaries” may also be reflected in later biblical sources mentioning the alleged cultural and ethnic makeup of the region; but this does not mean that these biblical images accurately reflect the complex sociocultural makeup and identity politics of this region during Iron I. In summary, the suggestion that there was a “Canaanite enclave” in the early Iron Age Shephelah is far from proved. To continue to put forward this suggestion requires much more explicit archaeological evidence.

The Status of Philistine Gath during the 10th–9th Centuries BCE Recently, a debate has been raging as to the date when the Kingdom of Judah expanded westward into the Shephelah. On the one hand, some scholars posit that already in the 10th century BCE the Judahite Kingdom expanded into the Shephelah, to its northern, central, and southern regions (e.g., Faust 2013b; Tappy 2011; Garfinkel, Kreimerman, and Zilberg 2016; Lederman and Bunimovitz 2014; Hardin, Rollston, and Blakely 2014; etc.). Accordingly, this development reflects the growing political power of the Kingdom, and at the same time, the weakening of the Philistine cities—and their subjugation by the newly ascendant Judahite Kingdom. On the other hand, various scholars have suggested (Koch 2012; Sergi 2013; Lehmann and Niemann 2014) that the Judahite Kingdom’s expansion westward, at least into the central Shephelah, did not occur until later, perhaps not until after the conquest of Gath by Hazael. Without going into too much detail, I would like to reiterate the dominant role that Gath played in the region until its destruction by Hazael ca. 830 BCE. Not only 2.  On “nested identities” in general, see Herb and Kaplan 1999. For archaeological applications, see, e.g., Janusek 2005; Hakenbeck 2007; 2011; Roberts 2011; Salazar et al. 2014; Scopacasa 2014.

140

Aren M. Maeir

Fig. 1.  Map of the Shephelah and Coastal Plain, with Tell eṣ-Ṣafi/Gath and primary Iron Age sites.

Philistine Gath after 20 Years

141

Fig. 2.  Aerial photograph (at the end of the 2016 season) of the fortifications and city gate of the lower city of Gath (north is at bottom of the picture).

is the city of Gath most probably the largest city in the southern Levant up until this destruction, there is now clear evidence (and on this see below in relation to the siege trench) that the entire city, including the extensive and impressively fortified lower city to the north (fig. 2). And to this one can add that, so far, with two decades of excavations at the site in the Iron Age layers, there is no evidence of a major destruction at the site from the early Iron Age until the destruction by Hazael and, similarly, there is no evidence of distinct, sudden changes in the material culture and its orientation. In other words, in light of the size, strength (extensive fortifications), and continuity exhibited at the site—and the very fact that it was perceived as a major target for conquest by Hazael—there can be little doubt that the Kingdom of Gath was the dominant polity in the central and northern part of the western Shephelah (and eastern Philistia) until the late 9th century BCE. Gath’s strong presence would effectively block any attempt on the part of the nascent Judahite kingdom to expand westward. Gath’s dominant status cannot be explained away, as Faust (2013b) attempted: it is impossible to claim that Gath should be seen as a unique case in Philistia, while other parts of Philistia were dominated by the Judahite Kingdom. One cannot simply sweep the major polity in the region under the carpet.

142

Aren M. Maeir

Fig. 3.  Early Iron Age Cooking Jugs from Tell eṣ-Ṣafi/Gath.

Similarly, claims that Khirbet Qeiyafa should be seen as evidence of the expansion of the early Judahite Kingdom (Garfinkel, Kreimerman, and Zilberg 2016) should be put into proper perspective. While I do tend to accept that Khirbet Qeiyafa is a Judahite site, it is important to stress that the site was abandoned soon after its construction (Garfinkel, Kreimerman, and Zilberg 2016: 94–96). 3 Needless to say, a substantial reason must lie behind the abandonment of such a carefully constructed site; the omnipresent threat of the Kingdom of Gath just to the west can be seen as the most likely cause. Thus, while Khirbet Qeiyafa may very likely have been an attempt of the early Judahite polity to expand into the central and western Shephelah, this attempt seems to have been quickly curtailed by the dominant polity in the region—the Kingdom of Gath.

Cultural Relationships between Gath and Judah The influence of various cultures in the southern Levant on Philistine culture is well known. Although the timing and meaning of these influences are at times 3.  Although the excavators suggest that the Iron IIA phase of the site was abandoned, in light of the large quantity of objects found in this level, including smashed cultic objects, one can wonder whether this stratum ended in a destruction. In any case (abandonment or destruction), the chances are very good that the underlying causes behind the end of this phase were pressure from, if not direct conquest by, Gath.

Philistine Gath after 20 Years

143

debated (see various opinions noted above), it is quite clear that from its earliest, formative stages until the end of the Iron Age, there is evidence of regular Levantine influence on Philistine culture. On the other hand, influences in the other direction, from Philistia to surrounding Levantine cultures, has been of much less focus. Over the years, various suggestions regarding Philistine linguistic influence on Israelite/Judahite culture, as seen in biblical and other textual materials, have been made (Rabin 1974). But very little has been said regarding influences as manifested in the material remains. A vessel type that appears in early Iron Age Philistia, with clear parallels from the Aegean Late Helladic cultures, Fig. 4.  Iron IIA (9th century BCE) jar, made with clay from the Jerusalem region, with a and which is often seen as a fossil directeur Judahite inscription, found in the temple in of early Philistine culture, is the “cooking the lower city of Gath. jug.” This form subsequently appears in various late Iron I and Iron II cultures in the southern Levant (fig. 3). Several years ago (Ben-Shlomo et al. 2008), it was pointed out that the “cooking jug” may reflect certain food preparation techniques typical of the Philistine culture and these techniques may have been adopted, or appropriated, by other Levantine Iron Age cultures; when this adoption took place, other cultures incorporated the cooking jug into their pottery repertoire. Recently, Kisilevitz (2015: 166–68) published several figurines from an Iron IIA temple at Moza, near Jerusalem, which appear to share similarities to Philistine figurines in their decorations. This probably is additional evidence of Philistine influence on surounding Levantine cultures and, in this case, in the cultic realm. Interestingly, to the list of bi-directional cultic influences between Philistia and Judah we can now add a jar, made in the region of Jerusalem, with what appears to be a Judahite inscription on it (Maeir and Eshel 2014; fig. 4 here). The jar was found in the Iron Age IIA temple in the lower city of Gath, right next to the two-horned monolithic stone altar (fig. 5), which likewise shows a combination of local and nonlocal influences in Philistine cult (Maeir 2012d). 4 4.  The stone altar from Area D at Tell eṣ-Ṣafi/Gath (Maeir 2012) will be published in detail in the future. In the meantime, it should be stressed that it is clear that this altar never had four stone horns—but only the two in the front. On the back of the altar, the original quarrying marks can still be seen, and there is no evidence on the back of the altar of two additional

144

Aren M. Maeir It appears that, at least until the 9th century BCE, and most likely in later phases of the Iron Age as well, ongoing bi-directional cultural influences existed between Philistia and Judah (and other regions in the Southern Levant as well). There is no reason to assume that the Philistine culture was not dominated, and did not become completely influenced by, Judean culture from the 10th century BCE onward (as Faust 2013b suggests). Not only did the Philistine material culture develop from Iron I into Iron IIA—with, among other features, continuity of Iron I symbolism (e.g., Maeir and Shai 2015)—but mutual influences between the two cultures continued without a doubt until the fall of Gath in the late 9th century BCE and very likely in later phases of the Iron Age as well.

Fig. 5.  Monolithic stone altar from the Iron IIA (9th century BCE) lower city of Gath. Note the two horns at the front of the altar and clear signs that the back portion of the altar was never fully finished; there is absolutely no evidence that originally two additional horns had existed in the back and that subsequently they had been removed.

The Siege System at Tell eṣ-Ṣafi/Gath

The siege system at Tell eṣ-Ṣafi/ Gath (fig. 6), which is dated to the Iron Age IIA and probably was constructed to attack the city by Hazael, the King of Aram Damascus, has been discussed in various publications (Ackermann, Maeir, and Bruins 2004; Ackermann, Bruins, and Maeir 2005; Maeir 2009; 2012a; 2012c; Maeir and Gur-Arieh 2011); the detailed report is currently in press (Gur-Arieh and Maeir in press). This identification has been accepted by all, save for Usshishkin (2009; 2014; 2015), who has questioned whether this feature is in fact a siege system at all. 5 Because a horns, which fell off or were removed at some stage. Thus, Faust and Lev-Tov’s (2014: 17 n. 32) and Nigro’s (2014: 3 n. 9) suggestions that this was in fact, originally, a four-horned altar cannot be accepted. 5. Garfinkel, Kreimerman, and Zilberg (2016: 112–13) have now joined Ussishkin in questioning the identification of this feature as a siege trench, but their arguments are rather unpersuasive. They ask: “Why would Hazael exhaust the strength of his forces by carrying out such an operation instead of attacking the city immediately upon arrival, or besieging it

Philistine Gath after 20 Years

145

Fig. 6.  Schematic section of elements of the Iron IIA (9th century BCE) siege system surrounding Tell eṣ-Ṣafi/Gath (based on finds in Area C6).

detailed response to his 2009 article has already been published (Maeir and GurArieh 2011), I do not intend of go over all the arguments for its identification as a siege system. Instead, in light of Ussishkin’s most recent views on the issue (2014; by other means? How did these warriors support themselves during the long march back and forth and during the protracted siege” (ibid.). These questions are somewhat surprising, because any familiarity with ancient siege warfare (as for example explained in Maeir and GurArieh 2011), would have made it clear that, despite these questions, time and again ancient armies besieged sites for extended periods and expended the effort to construct extensive and labor-intensive siege systems, when other methods of attacking the site were not viable. The reservations they raise regarding the identification of this installation as a siege system and it attribution to an Aramean conquest of Gath are similarly unconvincing. Several details can be cited: (1) They question the very identification of Tell eṣ-Ṣafi as Gath, without providing any compelling argument for this (Garfinkel, Kreimerman, and Zilberg 2016), despite the fact that this identification is certain (see Schniedewind 1998; Maeir 2012b). (2) They suggest that Gath was destroyed by another Philistine city-state (Garfinkel, Kreimerman, and Zilberg 2016: 113), once again without providing any support. (3) Furthermore, the interpretation of the brief mention of the conquest of Gath by Hazael that they prefer, particularly when compared to their positivistic and simplistic analysis of large sections of the biblical text in relation to the early Judahite monarchy, is difficult to accept. This lack of reference to and failure to utilize large swaths of up-to-date textual analyses and interpretive approaches to the biblical text seems to fit well with their methodological approach. They seem to ignore previous research on a topic when suggesting their own interpretations of the same topic (Garfinkel, Kreimerman, and Zilberg 2016:113)! (4) And, finally, it should be noted that they ignore what they themselves (Garfinkel, Kreimerman, and Zilberg 2016: 186) argue for— namely, an “Occam’s Razor” approach to interpreting archaeological evidence: identifying Hazael as the agent behind the destruction of Gath is the simplest solution in comparison to all other possibilities that have been suggested (see Maeir 2012a: 43–49).

146

Aren M. Maeir

Fig. 7.  View of a portion of the siege system surrounding Tell eṣ-Ṣafi/Gath, to the south of the site. The clearly visible berm (1) situated to the south of the trench (2) (on the side away from the city of Gath). Note the location of one of the besiegers’ towers (3), inserted into the berm.

2015), I will point out some of the weaknesses in the latest arguments against this interpretation. (a) Ussishkin (2014; 2015) attempts to disconnect the trench from the suggestion that it had a siege function and returns to Ephʿal’s (1996: 77; 2008: 81 n. 143) suggestion that the ḥrz mentioned in the Zakur inscription does not refer to a trench but to tunneling activity carried out by Bir-Hadad, the son of Hazael during his siege of Hadrach. Unfortunately, he does not consider the detailed discussion of Ephʿal’s suggestion, and the refutation of it, which we have published (Maeir 2009; Maeir and Gur-Arieh 2011), which makes his argument largely irrelevant. (b) Ussishkin (2014; 2015) does not believe that there was a northern side to the trench and does not accept our very logical suggestion that the river bed of the Elah Valley, which runs just to the north of the site served as such. Ussishkin (2014; 2015) states that, because the river bed runs just to the north of Area D, it could be easily crossed by inhabitants of Gath and would not have served as a substantial obstacle. Once again, this does not take into account several factors, most of which I have already published. (1) It has been clearly demonstrated, using remote sensing, that the trench goes far to the north toward the Elah Valley riverbed, and there cannot be any explanation for this unless the trench was intended to connect with the riverbed, thus forming the northern side of the siege barrier. (2) The riverbed, as well as the surrounding landscape, was substantially deeper in antiquity than today (Ackermann et al. 2014), and the riverbed would have been a substantial

147

Philistine Gath after 20 Years

topographical obstacle,not to mention that in antiquity, the river itself may have been perpetual rather than seasonal (Ackermann, pers. comm.) (3) Ussishkin (2014; 2015) assumes that substantial portions of the city were unfortified. This simply is not the case. Evidence of fortifications of the upper tell are now evident in Area F. Although the site was founded in the Early Bronze, the city wall continued to be used in the Middle Bronze, Late Bronze, Iron I, and Iron II. More importantly, the recent excavation at Tell eṣ-Ṣafi/Gath (as of summer 2016) has now demonstrated, beyond doubt, that a large-scale fortification, probably including a gate, surrounded and fortified the lower city as well (fig. 2). Even though evidence of this was seen on the surface prior to excavations, this area and this feature have now been excavated and clearly dated to the Iron Age I and IIA. (4) Ussishkin’s belief (2014: 10) that there is no evidence of a berm on the southern side of the site, as he claims in his interpretation of the photograph that he published (2014: fig. 4), is simply incorrect. Although the berm is not fully preserved on the southern side of the trench, several clear sections of the berm have been preserved (see fig. 7). In fact, one can even see evidence of the berm in the picture that Ussishkin presented. 6 (5) Finally, I a firm believer in the principal of “Occam’s Razor” when it comes to the interpretation of archaeological remains. Ussishkin (2014: 11) admits that he has no alternative explanation for the trench but still prefers not to identify it as a siege system. Since I believe that convincing arguments and evidence to identify this feature as a siege system have been presented, and even though a sign identifying it as such has yet to be found, the available evidence lends strong support to our suggestion.

Summary The issues covered in this overview represent a selection of the issues relating to the study of the Iron Age Shephelah, Philistine culture, and other associated matters that have emerged from the first twenty years of excavation at Tell eṣ-Ṣafi/ Gath. As the excavations continue, and more finds are revealed and analyzed and placed within their wider context, additional insights and reassessments of our understanding on these and related issues will no doubt emerge. 6.  As noted, a detailed study of the siege system will appear in the near future (Gur-Arieh and Maeir in press). An explanation of the geomorphological features connected with the trench and berm, including evidence for these features in various locations around the site, has appeared in previous publications (Ackermann, Maeir, and Bruins 2004; Ackermann, Bruins, and Maeir 2005).

References: Ackermann, O.; Maeir, A. M.; and Bruins, H. J. 2004 Unique Human-Made Catenary Changes and Their Effect on Soil and Vegetation in the Semi-Arid Mediterranean Zone: A Case Study on Sarcopterium Spinosum Distribution near Tell eṣ-Ṣâfi/Gath, Israel. Catena 57: 309–30.

148

Aren M. Maeir

Ackermann, O.; Bruins, H. J.; and Maeir, A. M. 2005 A Unique Human-Made Trench at Tell eṣ-Ṣafi/Gath, Israel: Anthropogenic Impact and Landscape Response. Geoarchaeology 20(3): 303–28. Ackermann, O.; Greenbaum, N.; Ayalon, A.; Bar-Matthews, A.; Boaretto, E.; Bruins, H.; Cabanes, D.; Horwitz, L. K.; Neumann, F.; Porat, N.; Weiss, E.; and Maeir, A. M. 2014 Using Palaeo-Environmental Proxies to Reconstruct Natural and Anthropogenic Controls on Sedimentation Rates, Tell eṣ-Ṣafi/Gath, Eastern Mediterranean. Anthropocene 8: 70–82. Bhabba, H. K. 1994 The Location of Culture. London: Routledge. Barth, F. 1969 Ethnic Groups and Boundaries. Boston: Little, Brown and Company. Ben-Shlomo, D.; Shai, I.; Zukerman, A.; and Maeir, A. M. 2008 Cooking Identities: Aegean-Style and Philistine Cooking Jugs and Cultural Interaction in the Southern Levant during the Iron Age. American Journal of Archaeology 112(2): 225–46. Bunimovitz, S.; and Faust, A. 2001 Chronological Separation, Geographical Segregation, or Ethnic Demarcation? Ethnography and the Iron Age Low Chronology. Bulletin of the American Schools of Oriental Research 322: 1–10. Bunimovitz, S.; and Lederman, Z. 2009 The Archaeology of Border Communities: Renewed Excavations at Tel BethShemesh, Part 1: The Iron Age. Near Eastern Archaeology 72(3): 114–42. 2011 Close Yet Apart: Diverse Cultural Dynamics at Iron Age Beth-Shemesh and Lachish. Pp. 33–53 in The Fire Signals of Lachish: Studies in the Archaeology and History of Israel in the Late Bronze Age, Iron Age and Persian Period in Honor of David Ussishkin, ed. I. Finkelstein and N. Naʾaman. Winona Lake, IN: Eisenbrauns. Casella, E. C.; and Fowler, C., eds. 2005 The Archaeology of Plural and Changing Identities: Beyond Identification. New York: Kluwer Academic. Castoriadis, C. 1975 L’institution imaginaire de la société. Paris: Seuil. Cline, E. H. 2014 1177 B.C.: The Year Civilization Collapsed. Princeton, NJ: Princeton University Press. Davis, B.; Maeir, A. M.; and Hitchcock, L. A. 2015 Disentangling Entangled Objects: Iron Age Inscriptions from Philistia as a Reflection of Cultural Processes. Israel Exploration Journal 65(2): 140–65. Dougherty, C.; and Kurke, L. 2003 The Cultures within Greek Culture. Cambridge: Cambridge University Press. Ehrlich, C. S. 2016 Biblical Gentilics and Israelite Ethnicity. Pp. 413–21 in The Books of Samuel. Stories  – History –Reception History, ed. W. Dietrich. Bibliotheca Ephermeridum Theologicarum Lovaniensium 284. Leuven: Peeters. Ephʿal, I. 1996 Siege and Its Ancient Near Eastern Manifestations. Jerusalem: Magnes. [Hebrew] 2008 The City Besieged: Siege and Its Manifestations in the Ancient Near East. Culture and History of the Ancient Near East 36. Leiden: Brill. Eriksen, T. H. 2010 Ethnicity and Nationalism: Anthropological Perspectives. 3rd ed. London

Philistine Gath after 20 Years

149

Faust, A. 2013a Decoration versus Simplicity: Pottery and Ethnic Negotiations in Early Israel. Ars Judaica 9: 7–18. 2013b From Regional Power to Peaceful Neighbour: Philistia in the Iron I–II Transition. Israel Exploration Journal 63: 174–204. 2015a The Bible, Archaeology, and the Practice of Circumcision in Israelite and Philistine Societies. Journal of Biblical Literature 134(2): 273–90. 2015b The Emergence of Iron Age Israel: On Origins and Habitus. Pp. 467–82 in Israel’s Exodus in Transdisciplinary Perspective: Text, Archaeology, Culture, and Geoscience, ed. T. E. Levy, T. Schneider and W. H. C. Propp. Quantitative Methods in the Humanities and Social Sciences. Heidelberg: Springer. 2015c The “Philistine Tomb” at Tel ‘Eton: Culture Contact, Colonialism, and Local Responses in Iron Age Shephelah, Israel. Journal of Anthropological Research 71: 195–230. 2015d Pottery and Society in Iron Age Philistia: Feasting, Identity, Economy, and Gender. Bulletin of the American Schools of Oriental Research 373: 167–98. Faust, A.; and Katz, H. 2011 Philistines, Israelites and Canaanites in the Southern Trough Valley during the Iron Age I. Egypt and the Levant 21: 231–47. 2015 A Canaanite Town, a Judahite Center, and a Persian Period Fort: Excavating Over Two Thousand Years of History at Tel ‘Eton. Near Eastern Archaeology 78(2): 88–102. Faust, A.; and Lev-Tov, J. 2011 The Constitution of Philistine Identity: Ethnic Dynamics in Twelfth to Tenth Century Philistia. Oxford Journal of Archaeology 30: 13–31. 2014 Philistia and the Philistines in the Iron Age I: Interaction, Ethnic Dynamics and Boundary Maintenance. HIPHIL Novum 1(1): 1–24. Frumin, S.; Maeir, A. M.; Horwitz, L. K.; and Weiss, E. 2105 Studying Ancient Anthropogenic Impact on Current Floral Biodiversity in the Southern Levant as Reflected by the Philistine Migration. Scientific Reports 5(13308): 1–10. Gardner, A. 2007 Fluid Frontiers: Cultural Interaction on the Edge of Empire. Stanford Journal of Archaeology 5: 43–60. Garfinkel, Y.; Kreimerman, I.; and Zilberg, P. 2016 Debating Khirbet Qeiyafa: A Fortified City in Judah from the Time of King David. Jerusalem: Israel Exploration Society. Gosden, C. 2004 Archaeology and Colonialism: Cultural Contact from 5000 BC to the Present. Topics in Contemporary Archaeology. Cambridge: Cambridge University Press. Gur-Arieh, S.; and Maeir, A. M. In press  Chapter 3. Area C: The Siege Trench and Other Features. In Tell eṣ-Ṣafi/Gath II: Excavation Reports and Studies, ed. A. M. Maeir and J. Uziel. Ägypten und Altes Testament. Münster: Ugarit-Verlag. Hakenbeck, S. 2007 Situational Ethnicity and Nested Identities: New Approaches to an Old Problem. Anglo-Saxon Studies in Archaeology and History 14: 19–27. 2011 Roman or Barbarian? Shifting Identities in Early Medieval Cemeteries in Bavaria. Post-Classical Archaeologies 1: 37–66. Hall, J. 2000 Ethnic Identity in Greek Antiquity. Cambridge: Cambridge University Press.

150

Aren M. Maeir

Hardin, J. W.; Rollston, C. A.; and Blakely, J. A. 2014 Iron Age Bullae from Officialdom’s Periphery: Khirbet Summeily in Broader Context. Near Eastern Archaeology 77(4): 299–301. Hitchcock, L. A.; and Maeir, A. M. 2013 Beyond Creolization and Hybridity: Entangled and Transcultural Identities in Philistia. Archaeological Review from Cambridge 28(1): 51–74. 2014 Yo-Ho, Yo-Ho, a Seren’s Life for Me! World Archaeology 46(3): 624–40. 2017 Fifteen Men on a Dead Seren’s Chest, Yo Ho Ho and a Krater of Wine. In Context and Connection: Essays on the Archaeology of the Ancient Near East in Honour of Antonio Sagona, ed. A. Batmaz, G. Bedianashvili, A. Michalewicz, and A. Robinson. Orientalia Lovaniensia Analecta. Leuven: Peeters. Horvarth, R. J. 1972 A Definition of Colonialism. Current Anthropology 13: 45–57. Janusek, J. W. 2005 Of Pots and People: Ceramic Style and Social Identity in the Tiwanaku State. Pp. 34– 53 in Us and Them: Archaeological and Ethnicity in the Andes, ed. R. M. Reycraft. Cotsen Institute of Archaeology, Monograph 53. Los Angeles: Cotsen Institute of Archaeology. Jasmin, M. 2006 L’étude de la transition du Bronze récent II au Fer I en Palestine méridionale. British Archaeological Reports, International Series 1495. Oxford: Archaeopress. Jordan, K. A. 2009 Colonies, Colonialism and Cultural Entanglement: The Archaeology of Postcolumbian Intercultural Relations. Pp. 31–49 in International Handbook of Historical Archaeology, ed. T. Majewsji and D. Gaimster. New York: Springer. Kisilevitz, S. 2015 The Iron IIA Judahite Temple at Tel Moza. Tel Aviv 42(2): 147–64. Koch, I. 2012 The Geopolitical Organization of the Judean Shephelah during Iron Age I–IIA. Cathedra 143: 45–64. [Hebrew] Kohn, M. 2011 Colonialism. In The Stanford Encyclopedia of Philosophy, ed. E. N. Zalta. Lederman, Z.; and Bunimovitz, S. 2014 Canaanites, “Shephelites” and Those Who Will Become Judahites. Pp. 61–71 in New Studies in the Archaeology of Jerusalem and Its Region, Vol. 8, ed. G. Stiebel, O. PelegBarkat, D. Ben-Ami, and Y. Gadot. Jerusalem: Institute of Archaeology. [Hebrew] Lehmann, G.; and Niemann, H. M. 2014 When Did the Shephelah Become Judahite? Tel Aviv 41(1): 77–94. Leuchter, M. 2013 Genesis 38 in Social and Historical Perspective. Journal of Biblical Literature 132(2): 209–27. Loomba, A. 2015 Colonialism/Postcolonialism. New Critical Idiom. London: Routledge. Maeir, A. M. 2009 Hazael, Birhadad, and the Ḥrẓ. Pp. 273–77 in Exploring the Longue Durée: Essays in Honor of Lawrence E. Stager, ed. J. D. Schloen. Winona Lake, IN: Eisenbrauns. 2012a Chapter 1: The Tell eṣ-Ṣafi/Gath Archaeological Project 1996–2010: Introduction, Overview and Synopsis of Results. Pp. 1–88 in Tell eṣ-Ṣafi/Gath I: Report on the 1996–2005 Seasons, ed. A. M. Maeir. Ägypten und Altes Testament 69. Wiesbaden: Harrassowitz.

Philistine Gath after 20 Years

151

2012b Chapter 2A: History of Research–1838 to 1996. Pp. 89–107 in Tell eṣ-Ṣafi/Gath I: Report on the 1996–2005 Seasons, ed. A. M. Maeir. Ägypten und Altes Testament 69. Wiesbaden: Harrassowitz. 2012c Insights on the Philistine Culture and Related Issues: An Overview of 15 Years of Work at Tell eṣ-Ṣafi/Gath. Pp. 345–403 in The Ancient Near East in the 12th–10th Centuries BCE, Culture and History: Proceedings of the International Conference Held in the University of Haifa, 2–5 May 2010, ed. G. Galil, A. Gilboa, A. M. Maeir, and D. Kahn. Alter Orient und Altes Testament 392. Münster: Ugarit-Verlag. 2012d Prize Find: Horned Altar from Tell eṣ-Ṣafi Hints at Philistine Origins. Biblical Archaeology Review 38(1): 35. 2013a Gath. Pp. 443–51 in The Oxford Encyclopedia of the Bible and Archaeology, ed. D. M. Master. New York: Oxford. 2013b Review of A. Faust. 2012. The Archaeology of Israelite Society. Eisenbrauns: Winona Lake, IN. Review of Biblical Literature. 2016 The Aramaean Involvement in the Southern Levant: Case Studies for Identifying the Archaeological Evidence. Pp. 79–87 in In Search of Aram and Israel: Politics, Culture and Identity, ed. O. Sergi, M. Oeming, and I. J. de Hulster. Orientalische Religionen in der Antike. Tübingen: Mohr-Siebeck. Maeir, A. M.; Davis, B.; and Hitchcock, L. A. 2016 Philistine Names and Terms Once Again: A Recent Perspective. Journal of Eastern Mediterranean Archaeology and Heritage 4(4): 321–40. Maeir, A. M.; and Gur-Arieh, S. 2011 Comparative aspects of the Aramean Siege System at Tell eṣ-Ṣāfi/Gath. Pp. 227–44 in The Fire Signals of Lachish: Studies in the Archaeology and History of Israel in the Late Bronze Age, Iron Age and Persian Period in Honor of David Ussishkin, ed. I. Finkelstein and N. Naʾaman. Winona Lake, IN: Eisenbrauns. Maeir, A. M.; and Hitchcock, L. A. 2016 “And the Canaanite Was Then in the Land”? A Critical View on the “Canaanite Enclave” in Iron I. Pp. 209–26 in Alphabets, Texts and Artefacts in the Ancient Near East: Studies Presented to Benjamin Sass, ed. I. Finkelstein, C. Robin, and T. Römer. Paris: Van Dieren. In press  The Appearance, Formation and Transformation of Philistine Culture: New Perspectives and New Finds. In The Sea Peoples Up-To-Date: New Research on the Migration of Peoples in the 12th Century BCE, ed. P. Fischer. Contributions to the Chronology of the Eastern Mediterranean. Vienna: Austrian Academy of Sciences. Maeir, A. M.; Hitchcock, L. A.; and Horwitz, L. K. 2013 On the Constitution and Transformation of Philistine Identity. Oxford Journal of Archaeology 32(1): 1–38. Maeir, A. M.; and Shai, I. 2015 The Origins of the “Late Philistine Decorated Ware”: A Note. Tel Aviv 42(1): 59–66. Malkin, I.; ed. 2001 Ancient Perceptions of Greek Ethnicity. Studies in Greek and Roman Religion 3. Washington, D.C.: Center for Hellenic Studies. Mazar, A. 2014 Archaeology and the Bible: Reflections on Historical Memory in the Deuteronomic History. Pp. 347–69 in Congress Volume Munich 2013, ed. C. M. Maier. Supplements to Vetus Testamentum 163. Leiden: Brill.

152

Aren M. Maeir

Meiri, M.; Huchon, D.; Bar-Oz, G.; Boaretto, E.; Kolska Horwitz, L.; Maeir, A. M.; SapirHen, L.; Larson, G.; Weiner, S.; and Finkelstein, I. 2013 Ancient DNA and Population Turnover in Southern Levantine Pigs- Signature of the Sea Peoples Migration? Scientific Reports 3: 3035 | DOI: 10.1038/srep03035. Mengoni, L. E. 2010 Identity Formation in a Border Area: The Cemeteries of Baoxing, Western Sichuan (Third Century BCE–Second Century CE). Journal of Social Archaeology 10(2): 198–229. Münnich, M. M. 2013 Beth-Shemesh in the Early Iron Age. Ugarit-Forschungen 44: 217–41. Naʾaman, N. 2010 Kh. Qeiyafa in Context. Ugarit-Forschungen 42: 497–526. Nestor, D. A. 2010 Cognitive Perspectives on Israelite Identity. Library of Hebrew Bible/Old Testament Studies 519. New York: T&T Clark. Nigro, L. 2014 David e Golia: Filistei e Israeliti ad un tiro di sasso. Recenti scoperte nel dibattito sull’archeologia in Israele. Quaderni di Vicino Oriente 8: 1–17. Osterhammel, J. 2005 Kolonialismus: Geschichte, Formen, Folgen. 5th ed. Munich: Beck. Poblome, J.; Malfitana, D.; and Lund, J. 2014 It’s Complicated.. Past Cultural Identity and Plain Broken Pottery. Pp. xi–xxvii in Congressus Vicesimus Octavus. Rei Cretariae Romanae Fautorum. Catinae Habitus Acta 43, ed. S. Biegert. Rei Cretariae Romanae Fautorum 43. Bonn: Habelt. Reger, G. 2014 Ethnic Identities, Borderlands, and Hybridity. Pp. 112–29 in A Companion to Ethnicity in the Ancient Mediterranean, ed. J. McInerney. Blackwell Companions to the Ancient World. Malden, MA: Wiley. Roberts, C. M. 2011 Practical Identities: On the Relationship Between Iconography and Group Identity. Pp. 86–95 in Identity Crisis: Archaeological Perspectives on Social Identity. Proceedings of the 42nd (2010) Annual Chacmool Archaeology Conference, University of Calgary, Calgary, Alberta, ed. L. Amundsen-Meyer, N. Engel, and S. Pickering. Calgary: Chacmool Archaeological Association, University of Calgary. Ross, D. E. 2012 Transnational Artifacts: Grappling with Fluid Material Origins and Identities in Archaeological Interpretations of Culture Change. Journal of Anthropological Archaeology 31: 38–48. Salazar, D.; Niemeyer, H. M.; Horta, H.; Figueroa, V.; and Manríquez, G. 2014 Interaction, Social Identity, Agency and Change During Middle Horizon San Pedro de Atacama (Northern Chile): A Multidimensional and Interdisciplinary Perspective. Journal of Anthropological Archaeology 35: 135–52. Schniedewind, W. M. 1998 The Geopolitical History of Philistine Gath. Bulletin of the American Schools of Oriental Research 309: 69–77. Scopacasa, R. 2014 Building Communities in Ancient Samnium: Cult, Ethnicity and Nested Identities. Oxford Journal of Archaeology 33(1): 69–87. Sergi, O. 2013 Judah’s Expansion in Historical Context. Tel Aviv 40: 226–46.

Philistine Gath after 20 Years

153

Siapkas, J. 2014 Ancient Ethnicity and Modern Identity. Pp. 66–81 in A Companion to Ethnicity in the Ancient Mediterranean, ed. J. McInerney. Blackwell Companions to the Ancient World. Malden, MA: Wiley. Stavrianopoulou, E.; ed. 2013 Shifting Social Imaginaries in the Hellenistic Period: Narrations, Practices, and Images. Mnemosyne Supplements. Monographs on Greek and Latin Language and Literature 363. Leiden: Brill. Steinmitz, G. 2014 The Sociology of Empires, Colonies, and Postcolonialism. Annual Review of Sociology 40: 77–103. Strauss, C. 2006 The Imaginary. Anthropological Theory 6: 322–44. Tappy, R. 2011 The Depositional History of Iron Age Tel Zayit: A Response to Finkelstein, Sass, and Singer-Avitz. Eretz Israel (A. Ben-Tor Volume) 30: 127*–43*. Taylor, C. 2002 Modern Social Imaginaries. Public Culture 14(1): 91–124. Uehlinger, C. 1999 The “Canaanites” and Other “Pre-Israelite” Peoples in Story and History, Part I. Freiburger Zeitschrift für Philosophie und Theologie 46: 546–78. 2000 The “Canaanites” and Other “Pre-Israelite” Peoples in Story and History, Part II. Freiburger Zeitschrift für Philosophie und Theologie 47: 173–98. Ussishkin, D. 2009 On the So-Called Aramaean ‘Siege Trench” in Tell eṣ-Ṣafi, Ancient Gath. Israel Exploration Journal 59(2): 137–57. 2014 Gath, Lachish and Jerusalem in the 9th Century BCE: The Archaeological Perspective (In Hebrew with English abstract). Pp. 7–34 in New Studies on Jerusalem, Vol. 20, ed. E. Baruch and A. Faust. Ramat-Gan: Ingeborg Rennert Center for Jerusalem Studies. 2015 Gath, Lachish and Jerusalem in the 9th Century BCE: An Archaeological Reassessment. Zeitschrift des deutschen Palästina Vereins 131(2): 129–49. Van Dommelen, P. 2012 Colonialism and Migration in the Ancient Mediterranean. Annual Review of Anthropology 41: 393–409. Van Nijf, O. 2010 Being Termessian: Local Knowledge and Identity Politics in a Pisidian City. Pp. 163– 88 in Local Knowledge and Microidentities in the Imperial Greek World, ed. T. Whitmarsh. Cambridge: Cambridge University Press. White, R. 1991 The Middle Ground: Indians, Empires, and Republics in the Great Lakes Region, 1650– 1815. Cambridge: Cambridge University Press. Whitmarsh, T. 2010 Local Knowledge and Microidentities in the Imperial Greek World. Cambridge: Cambridge University Press. Woolf, G. 2011 Tales of the Barbarians: Ethnography and Empire in the Roman West. Malden, MA: Wiley-Blackwell.

154

Aren M. Maeir

Yasur-Landau, A. 2010 The Philistines and Aegean Migration at the End of the Late Bronze Age. Cambridge: Cambridge University Press. Ypi, L. 2013 What’s Wrong with Colonialism. Philosophy and Public Affairs 41(2): 158–91.

The Archaeology and History of Tel Zayit: A Record of Liminal Life Ron E. Tappy

One nineteenth-century geographer eloquently described the western slopes of the Judean hill country as “that natural rampart” which formed the best defense for the highland cultures (Ritter 1866: 246). At the base of this topographical feature, the ancient town of Tel Zayit lay along a road that ran through the strategic Naḥal Guvrin on its way out to the large coastal port at Ashkelon. The 30-dunam site (fig. 1) was situated nearly 30 km east of Ashkelon and roughly halfway between Lachish to the south and Tell eṣ-Ṣâfi to the north. Although this area generally belonged to the lowlands district of ancient Judah, it often was a contested zone wherein cultural and, certainly, political associations shifted from time to time, primarily between the highlands to the east and the coastal plain to the west. Since 1998, the exploration of Tel Zayit has proceeded under my direction and with the sponsorship of Pittsburgh Theological Seminary. When viewed through the lens of historical and comparative methods espoused by the Annales school of historiography (compare Braudel 1972; 1980: 25–54; Bloch 1953; 1974; 1995), the archaeological remains thus far exposed at Tel Zayit speak to the enduring status of the town’s strategic position as a borderland community—one that existed and sometimes thrived in the middle ground between larger entities. The location of Tel Zayit in the fluid margin between Judah and Philistia constituted the principal contributing factor to the site’s long-term political and cultural significance. Having a clear interpretive model against which to examine this “liminal zone,” with its varied topographical features, is a basic requirement for proper appreciation of Tel Zayit’s extended (at least 3,700-year) history. After outlining the interpretive framework that I believe best suits the Shephelah, or “Lowlands,” of biblical Judah, I shall draw on historical, textual, and archaeological data from three different periods to demonstrate the applicability of this model to life at ancient Tel Zayit.

I. Reimagining Peripheries The Latin word līmen captures nicely the notion of existence between two or more entities or realities. Latin texts often employ this word to denote a threshold, a passageway from point A to point B. Ancient cultures tended to view these 155

156

Ron E. Tappy

Fig. 1.  Tel Zayit, view from the north; Site Plan (R. E. Tappy).

transitional points (including gates, doorways, windows, crossroads, shrines, etc.) as thresholds where the magical, the invisible, the world of deities met and merged with or influenced the mundane, the visible, the world of humans. Literature produced by these cultures spanning various genres (including laws, incantations, and medical texts) and geocultural areas (from Mesopotamia to Egypt)―repeatedly attests to this mythology. Writers have long recognized liminal space as a place where one meets the other and where elements of both fission and fusion play themselves out. Clearly, the idea of liminality is not merely a modern construct. Not surprisingly, then, this concept has emerged in multiple areas of scholarship investigating both ancient and modern contexts. Following Arnold Van Gennep’s Rites of Passage (1909), Karl Theodor Jaspers (1953) utilized the concept of a liminal period to describe the interstices between the great empires of the “axial age.” ( Jasper’s use of liminality builds on that of Max Weber and Eric Voegelin; see recently Bellah 2011: 265–323; Bellah and Joas 2012.) Scholars have drawn fullscale studies of the liminal warrior from biblical (the Samson narrative in Judges 13–16, which some have classified as a border epic; Cross and Stager 1991), Mesopotamian (Gilgamesh), Greek (Odyssey), and even Old English (Beowulf) narratives and epic poems (compare, in chronological order, Reiner 1967: 116–20; Tigay 1982; Wolf 1987; Mobley 2006). One easily recognizes the enduring motif in both the writings and personal lives of many modern-day authors (from Percy Bysshe and Mary Shelley to Edgar Allen Poe, Robert Lewis Stevenson, Thomas Hardy, Joseph Conrad, and many more). Others have identified liminal aspects in the erotic activities of the goddess Inanna (e.g., Assante 2002: 27–52), in biblical law (Stahl 1995), ancient maritime activities (Monroe 2008), and traditions dealing with gender and sexuality in the biblical world (Ackerman 2005). One thing seems clear: sometimes entire communities,―not just individuals,―find themselves having to exist and seek

The Archaeology and History of Tel Zayit

157

Fig. 2.  Sites in and around the liminal zone of the Shephelah (R. E. Tappy.)

or create self-identification within a transitional zone (note, for example, studies of closely-spaced, inner-montane villages situated in the alpine valleys of northern Italy [Cole and Wolf 1974]). Such was the case for the ancient residents living in the borderland town at Tel Zayit. Although many archaeologists working in the Shephelah of Israel now describe their sites as borderland locales, the field as a whole has not articulated a proper model against which to study these areas. Renewed excavations at Beth Shemesh (e.g., Bunimovitz and Lederman 1997a; 1997b; 2001), the impressive publication of fieldwork at Lachish (Ussishkin 2004), and projects at other sites across the area (including Tel Zayit) have now yielded a wealth of information that enables us to understand this general region as never before. Through the course of the tenth century BCE, the large domestic village at Beth Shemesh (Level 4) evolved into an urban administrative center (Level 3), with symbolic architecture and material culture that reflect an organized political structure, which presumably was located in the mountains to the east. A similar pattern played out, albeit on a more modest scale, at both Tel Zayit and Lachish, where newly established towns followed 200-year occupation gaps during the Iron Age I. Although sites in the southern Shephelah show a more moderate development over the course of the tenth century, from a geopolitical standpoint the settlement at Tel Zayit is quite significant, and the presence there of a mature, 22-letter alphabetic inscription from this period attests to that significance.

158

Ron E. Tappy

Tectonic activity and the runoff of water from the highlands resulted in a network of east–west wadis that descend down the seaward slopes of the hill country and through the Shephelah as they approach the inner coastal plain (fig. 2). These drainage systems include, from north to south, Naḥal Sorek-Refaʿim (Nahar Rubin), which bisects the hills immediately west of Jerusalem, Naḥal HaElah (Wâdī eṣ-Ṣunt), Naḥal Guvrin (Valley of Zephatha), Naḥal Lachish (Wâdī Qubeibeh), Naḥal Shiqma-Adorayim (Wâdī el-Ḥesi), and Naḥal Besor-Gerar (Wâdī Ghazzeh). These environmental features accommodate the runoff on the seaward slopes of the Judean highlands as far south as the Beersheba Basin. At least three principal north–south and three east–west roadways traversed these lowland areas and converged near Tel Zayit during various phases of the Iron Age (see Dorsey 1991: 58, Map 1). The longitudinal roads connected Egypt and the northern Sinai Peninsula with the southernmost Philistine capital at Gaza and the lowland area of Judah. At least three laterally oriented routes (Dorsey 1991: 182, 194, Maps 13–14) linked coastal centers with the interior hill country by exploiting the natural landscape provided by three of the six major drainage systems just mentioned—namely, the Elah, Guvrin, and Lachish valleys. These passageways prove especially important to our work at Tel Zayit, which lies at the western entrance to the central arena, the Naḥal Guvrin. The determinative role that geography plays in defining the frontier of a particular cultural set must always be considered. The Shephelah provides an ideal example of this phenomenon, since these lateral valleys helped to organize the sites of the lowlands and inner coastal area into discernible groups. Such natural groupings of towns along and within the liminal zone molded to a large extent boundary lines and economic exchange systems there. Seminal essays by Anson Rainey (1980; 1983) long ago became the bellwether of Shephelah studies and clarified the layout of the local districts in this area. More recent studies (Tappy 2008a; 2008b; 2008c; 2012; Hardin, Rollston, and Blakely 2012) have refined our understanding of these units. The physical terrain of these lateral passageways correlates well with the outline given in Josh 15:33–44 of the districts and cities belonging to the western lowlands of Judah. The author(s)/redactor(s) of this list organized the settlements of the Shephelah into three geographical groups that follow roughly the Elah (vv. 35–36 = District 2), Lachish (vv.  37–41 = District 3), and Guvrin (vv.  42–44 = District 4) systems. In effect, water basins created political boundaries within and between regions. The Joshua roster names nine towns in the Guvrin District 4, with Libnah apparently representing the main city in that district. These facts bear on the ancient identity of Tel Zayit, which likely relates in some way to the list of sites. If Tel Zayit is not ancient Libnah, the town undoubtedly lay so close to Libnah that it would have followed this important local center (even over Lachish) in most regional matters. Recently, discussion of the tenth century has shifted focus from Jerusalem to the Shephelah—that is, to the western border of Judah. Scholars have sometimes

The Archaeology and History of Tel Zayit

159

used the concepts of “core” and “periphery” to illustrate the symbiotic relationship between a cultural or political center (such as Jerusalem) and the surrounding territory (here, the Shephelah). One might illustrate this relationship by two concentric circles, in which the center seeks to promote its self-interests by wielding partial or absolute control over the territory around it. But the concept of periphery ultimately proves inadequate for describing the cultural history and complex relationships in the lowlands of southern Canaan. The term implies an overly specific, discernible line in the sand that marks the outer edge of the center’s real or symbolic presence, commercial relations, and more. Although the core–periphery paradigm aptly describes the desire of regal-ritual cities (such as Jerusalem) to recreate themselves in the outlying countryside (Fox 1977), it is, more often than not, overly centrist and delimiting. So while this model may symbolize the flow of goods and services to and from an urban hub, it is ultimately centripetal in nature, oriented toward its own center. Appeals to this framework typically seek to show how the two entities (core and periphery) relate to each other. This approach cannot capture the complexities inherent in multiple cores whose peripheries collide, merge, or overtake one another or form their own local sense of self-identity or definition. Even galaxies collide (as we now expect Andromeda and our own Milky Way to do four billion years from now). How much more so the ill-fitted pieces of the tiny Iron Age puzzle we call Canaan? Consequently, a more dynamic model is needed to examine relationships between these combined entities (a core and its periphery) and other similar or dissimilar component units around them. Early studies in symbolic anthropology help provide a different lens through which to study the archaeology and history of the Shephelah. Arnold van Gennep (1960), whose landmark study I mentioned above, investigated the rituals associated with transitional stages in human life. He ultimately coined the now wellworn phrase “rite of passage” and outlined three discernible phases of ritual passage: préliminaire, liminaire, and postliminaire. Building on van Gennep’s threefold framework, Victor Turner (1964, 1967, 1969, 1977) later applied this model to his studies of the Ndembu rituals in northwestern Zambia. Turner used the rubrics “separation, margin (līmen), and aggregation” to describe customs that attend such transitions as males passing from boyhood to manhood, and he focused most of his attention on the liminaire, or the liminal aspect of transition. The typical rite of passage, said both Van Gennep and Turner, begins with the separation of a person from his/her original status and proceeds as the ritual passenger moves through a marginal period before reaching aggregation (or reincorporation)—that is, the point at which the transition is consummated. Entrance into the līmen entails both leaving the structured conditions that apply to nonparticipants and entering a phase of unstructured conditions. The status of the passenger becomes ambiguous. After initiating the rites of passage, the young male is temporarily neither boy nor man, and all ritual applicants are viewed as equals (at least in terms of their temporary lack of status).

160

Ron E. Tappy

In adapting this general concept to the lowlands of ancient Judah, however, I am not suggesting that the conjunction of highland and coastal cultures gave rise to a status- or classless, egalitarian society for those who lived in the borderland area. Instead, I think of this region as one that struggled to hold and manage what Eade has called the “co-existence of numerous oppositions” (Eade 2000: 52; see also Sallnow 1981: 177). The marginal zone between Judah and Philistia did not undergo an economic or cultural leveling process at the hands of either core ( Jerusalem or the Philistine Pentapolis). In my adaptation of Turner’s model of liminality, the Shephelah does not represent a structureless area that, for that reason alone, set itself against the highly structured cores that surrounded it. Unlike Van Gennep’s or Turner’s passenger, who relinquished all status while in the transitional middle-ground of the ritual process, the liminality of the lowlands did not offer the residents there a release from the sociocultural constraints of their respective homeland cores; if anything, the practical expectations and symbolic culture that those native cores attempted to impose on them were designed to highlight their differences. Although interstructural liminality maintains a sociological focus in the anthropological studies of Van Gennep and Turner, then, this principle takes on cultural and political aspects for biblical towns that, because of their mere location, had to function in relation to two or more cores that straddled them. Borderland towns such as Tel Zayit existed betwixt and between the cultural, political, social, economic, judicial, ideological, and theological trappings of the larger, more structured units around them, all of which sought advantage in the balance of control at their borders. Such towns, then, present somewhat of a paradox in that they may at once constitute the core’s (or cores’) principal building blocks of outwardly bound solidarity and also some of the most vulnerable elements in that solidarity. Yet despite their somewhat tenuous status, these cultural smorgasbords often prove quite durable; they may outlast significant changes or even political decline within the opposing cores that surround and attempt to exploit them. Thus the concept of a līmen, or threshold, accurately characterizes the narrow cultural zones that lie just beyond recognized, more stable political boundaries, where competing cores seek to reach out and stake their claims through the use of myriad symbols, such as architecture, language, ethnicity, cultural or religious traditions, dietary taboos, even literary trends and traditions, and so on. Cores find it difficult to control the fringe areas around their frontiers totally, although they wish to do so. More than a binary, core–periphery model, the partially overlapping circles of a simple Venn diagram (fig. 3) depict more incisively situations in which entire politico-cultural entities (core + periphery) are juxtaposed to and collide with similar entities with which they interact in multiple and varied ways, including competition for territory, goods, resources, and control. This approach works especially well in the analysis of both small, borderland units that are geographically or

The Archaeology and History of Tel Zayit

161

Fig. 3.  Schematic of a bidirectional liminal zone (R. E. Tappy).

topographically defined (such as the Shephelah political districts of ancient Judah) and of larger regional kingdoms that must fit within relatively circumscribed areas (such as Judah and Philistia in southern Canaan). The liminal zone, then, embodies a truly middle and often contested area that must, by necessity, relate in various ways and at various times to the disparate cultural sets (two or more) that surround it. The concept embodies a place where cultures—through group encounters, positive interactions, and conflict—compete vigorously for presence, meaning, and interpretation. The Shephelah represents the theater in which the historical record of relations between the highland cultures to the east and those of the coastal plains to the west played itself out in spatial and temporal terms. I sometimes use the mathematical term “set” to describe each of the competing cores whose outer edges have collided. The totality of two overlapping sets represents the union of all aspects of both cores, whereas the overlap itself reflects the intersection/collision of disparate symbols and influence and therefore heightens the liminal feel of everyday life. It becomes apparent that the hybrid character of the liminal zone involves boundaries, borders, and frontiers. Boundaries entail those areas in which the real or symbolic presence of a particular core remains quite discernible through idiosyncratic architectural designs, governmental structures, social customs, and the like—that is, cultural emblems that substantially identify a core by promoting or enforcing a range of meanings shared by the vast majority of people living there. (On the concept of boundary, see Cohen 1985; Cohen, ed. 1986.) The symbols of a cultural or political set are generally clear, understood, accepted, and promoted within the boundaries of that set. This concept, then, connotes the outer limits of an established, relatively stable sphere of influence. Towns along and within the boundary of a particular core will ordinarily display similar traits that reflect

162

Ron E. Tappy

a coherent definition or presence of that core. The monumental architecture witnessed at Beth Shemesh (Level 3), for example, served symbolically to demarcate the boundary of Judah in the Sorek Valley, and we should expect other towns or cities (e.g., Azekah, Khirbet Qeiyafa, Socoh, Tel Burna, Lachish) filling the same role in adjacent or nearby locations to display similar features. A border, on the other hand, represents a relatively narrow area lying just outside the boundary of a core. If a boundary conjures the idea of a discernible line, a border typically lies along but also just beyond that line, like the matte that frames a picture or the ornamental fringe around a rug. As such, the border represents a secondary, liminal area that both attracts and proves vital to the self-serving interests of the cultural sets around it. But governing influences may shift here, evenly or sporadically. A given core’s influence is generally stronger near its actual boundary and diminishes as one moves away from it and through the borderlands—that is, through the more marginal (liminal) area. Taken together, the boundary and border of a cultural set comprise the frontier of that unit, the strategically important, coveted space that is inevitably shared with and vied for by other, nearby sets. The frontier incorporates the farthest, dissipating range of a core’s presence or influence. The concept of trans-frontier, then, includes not only those towns that lie along the boundary of a particular core but also other sites situated beyond that limited space and across the border/liminal zone itself—the area into which surrounding, competing cores find the easiest path to extending their influence and control. Thus the radial sway of any given cultural area or political entity does not end at a sharp, fixed point; rather, it fades out as the power radius from an adjacent core fades in. Unlike the periphery, which is conceptually unidirectional (that is, it relates primarily to its counterpart, the core), the liminal zone is at least bidirectional, Janus in nature. Sites within this zone must sort out their affiliations with two (or more) sometimes cooperative but often opposing cores. Such areas are especially vulnerable to seemingly erratic group behaviors embodying “complementary opposition.” Any core can, of course, share liminal zones with more than one adjacent cultural or political unit. In such cases, the borderlands can relate primarily to one of several cores or to multiple cores simultaneously. In the latter instance, the area quickly becomes a much more complex liminal zone. If a single northern set were added to the present Judah–Philistia illustration—for example, the imposing reach of a powerful city near the borderlands of both cores (such as Gezer)—the complexity of relationships would increase dramatically, particularly in the northern Shephelah. Judah itself provides a good example of this reality. In addition to managing its western front, this highland core undoubtedly had to administer similar areas heading in other compass directions, as evident in the back-and-forth maneuvers of Asa of Judah and Baasha of Israel in 1 Kgs 15:16–22. The short distance between Ramah and Miẓpah belonged to a rather narrowly defined liminal zone. When compared to the marginal zone in the Shephelah, it appears that in southern

The Archaeology and History of Tel Zayit

163

Canaan such areas typically spanned a 3–5 km stretch beyond the boundaries of the respective political centers. The reciprocal engagement between Asa and Baasha underscores the concept of shifting margins between otherwise fairly well-defined entities. There always existed an imperfect fit between, say, tightly spaced city-states such as those attested in the el Amarna correspondence from the Late Bronze Age or juxtaposed regional kingdoms documented in Iron Age Canaan. Regardless of the criteria along which any core developed its self-definition (e.g., ethnic, racial, linguistic, cultural, religious, or other forms of homogeneity), borderlands always gave the lie to this construct. For this reason, no core area or cultural set that managed to establish a presence or some degree of political control in any of the liminal zones around its boundaries could take that foothold for granted. For whatever continuity a core may project into the liminality beyond its boundary, the ostensible diversity that exists there may actually promulgate its own distinct local culture or sense of identity—a culture or identity that accepts only in a selective way the cores’ attempts to expand their influence through various types of symbols, interchanges, and kinship ties. In a liminal state, it may well serve the best interests of the local participants to maintain ties with both (all) of the contentious core cultures along their frontiers. In other words, “cultures have borders” but “borders have cultures” (as recognized by Rabinowitz 1998: 142; see Barth 1969: 10), and a liminal-zone culture that senses its own autonomy will act accordingly. Circumstances may, in fact, prompt the inhabitants of a liminal zone to feel torn between the competing cores that lie beyond their own imagined borders and, as a result, to express greater social and demographic unity among themselves. They may even come to resist attempts by any surrounding area to impose hard political boundaries within or through their region. Thus, it is not simply the institutional structures (e.g., regional governors) or symbols of state (embedded, for example, in monumental architecture, writing, and religious or cultic icons) of two or more cores that give meaning to borderland areas. Emergent local customs and the will of the various peoples who coinhabit these zones play vital roles in their own self-definition. Historically, geographically, and culturally, the area that lay between an imaginary (and movable) line running north from ancient Lachish to Beth Shemesh and another that extended from Tell el-Ḥesi through Tell el-ʿAreini and Tel Zayit to Tel Batash (Timnah), probably taking in at times even Philistine Gath (Tell eṣ-Ṣâfi), provides a perfect example of a līmen, or bidirectional threshold, between ancient Judah and Philistia. To the east of this area, larger fortified sites such as Beth Shemesh, Azekah, and Lachish (and perhaps Tel Burna) represented the western boundary of Judah. Each of these sites guarded a particular valley passageway into the hill country. Other sites, such as Tell Judeideh, Khirbet Qeiyafa, and Socoh, undoubtedly served to reinforce these more exposed boundary towns.

164

Ron E. Tappy

Fig. 4.  The Tel Zayit stone, showing the abecedary and a ground-out hollow (Z. Radovan, Jerusalem).

Inside this frontier edge, the Judahite core in the highlands would have experienced greater success at maintaining a more dominant role as arbiter of control and order, judicial organization, cultural and political affairs, kinship structures, and more—that is to say, of all the structured elements affording definition to the unit as a whole. On the other hand, in the liminal zone outside (west of) the boundary (that is, in the borderland, where the penetration of some of these same elements was less structured but still highly symbolic in nature), the arbitration powers of the Judahite and Philistine cores would have grown progressively more limited and challenged even as the highland and coastal plain centers sought to grow themselves. Viewed in this light, one should expect to encounter alternating trends or even a mix within the material culture of Shephelah sites lying along or within the liminal zone—a situation to which the archaeological record bears clear witness. By tracking changes in the archaeological record at key sites in the region, both the general history of the area and the specific nature of the liminal zone itself become more apparent (I have sketched out the general picture in Tappy 2008c: 26–37).

II.  Three Applications of the Liminal-Zone Model Let us return to Tel Zayit with my interpretive model in mind. Archaeological and historical considerations from widely separated phases in the settlement history of Tel Zayit shed light on the highland–coastal interplay in the intermediary lowlands area surrounding the site. (1) On July 15, 2005, the final day of that year’s season, excavators discovered a heavy limestone boulder with two lines of letters clearly incised on one side and a large, bowl-shaped hollow ground into the other side (figs. 4–5). The inscription is of an archaic alphabet found, notably, in a secure and well-defined archaeological context dating to the tenth century BCE, a much-debated period in the history of Israel. The provenance of this abecedary within a liminal zone that was about to

The Archaeology and History of Tel Zayit

165

Fig. 5.  The script of the Tel Zayit abecedary.

undergo significant cultural changes raises important issues related to the identification of the language behind the inscription and its place in the evolution of literacy within southern Canaan. Following the Late Bronze Age and the build-up of more than six meters of occupational levels at Tel Zayit, no substantial settlement at the site appears to have occurred during the Iron Age I period. Not until the Iron IIA period, sometime after 1000 BCE, did a new town arise on the summit of the mound. Occupation continued throughout the next 300 years (tenth–eighth centuries BCE), with a major conflagration occurring at the end of each century. When badly burned, disarticulated mud bricks appeared around the exposed top of a large monolith near the eastern shoulder of the tell, excavation there revealed architecture showing clear signs of a massive destruction by fire. Heavily burned debris smothered two adjacent chambers (one with a flagstone floor, the other with a beaten-earth surface; fig. 6) and associated walls. Although the eastern half of this structure had succumbed to heavy erosion, the western wall survived, and this feature incorporated the Tel Zayit Abecedary.

166

Ron E. Tappy

Fig. 6.  Two rooms from a tenth-century structure (circle indicates the position of the abecedary) (R. E. Tappy).

The ceramic assemblage associated with these rooms, though surprisingly meager, displays dark-red slip and coarsely spaced, irregular and discontinuous handburnishing, which are classic tenth-century traits. The best-preserved bowl (see Tappy 2006: 12, fig. 4:10) shows a combination of horizontal hand-burnishing around straight upper sidewalls and basket-weave pattern-burnishing across the interior base. Similar forms and treatment are well attested in Lachish Levels V–IV, from the late tenth–ninth centuries BCE (Zimhoni 1997: 88–97, 115–16, and figs. 3.15:5; 3.16:4; 3.17:13; 3.21:1, 7, 11, 15, 23). Ninety-two percent of the bowls from Lachish Level V display both slip and burnishing (versus slip only); by Level IV, the quantity reduces to 80% (Zimhoni 1997: 116, fig. 3.35). At Tel Zayit, a nearly whole cooking pot, filled with botanical debris, rested directly on the easternmost extension of the beaten-earth floor mentioned above. Both the ceramic form and a radiocarbon test of its contents confirm a tenth-century (in fact, mid-tenth-century) BCE date for this context (fig. 7). The best available parallel for this vessel comes from Stratum V at Beersheba (Aharoni 1973: pl. 54:10). Following the late tenth-century BCE crisis, the town rebounded quickly, and at least two new, thick floors of the early ninth century BCE covered and sealed the previous phase, including the findspot of the inscription. One cannot overvalue

The Archaeology and History of Tel Zayit

167

Fig. 7.  A cooking pot and radiocarbon dating of its contents.

this stratigraphic datum. That successive floor levels completely sealed the conflagration provides a firm terminus ante quem for the functional life of the preceding structure and thus for the inscription it contained. Because some stones in the wall partially overrode the letters of the abecedary, the incising of the stone occurred sometime prior to the construction of the building. Thus the script itself likely dates no later than the mid-tenth century BCE (a date supported by the cooking pot and its contents). In other words, the terminus ante quem for the actual engraving of the stone relates directly to the construction date of this stratum, not to the later time of its destruction, from which the bulk of the pottery described above would derive. Just how long before the construction of this tenth-century building someone engraved the alphabet into the stone remains difficult to determine, but the 200-year gap in occupation immediately prior to the tenth-century level provides sufficient control on that end. Regardless of whether the architectural structure that yielded the Tel Zayit Abecedary suffered heavy destruction by fire in the latter part of the tenth century or in the early ninth century BCE, the construction of the building and the incising of the inscription must have occurred earlier in the tenth century BCE, sometime around the midpoint of the century. No principle of investigation (stratigraphic or ceramic analysis, radiocarbon dating, historical reckoning, paleographic study) undermines this interpretation. But what about liminality? After the town’s rebuilding in the early ninth century, it thrived once again, only to end in another blaze of destruction in the late ninth century. A thick blanket of debris (up to 0.73 m) smothered all the principal features. But coastal ceramic industries now dominate the ceramic repertoire, though many of the forms were manufactured locally in the Shephelah. Surely, some of these traditions—for example, Ashdod Ware, one-handled cooking jars,

168

Ron E. Tappy

and short-necked coastal amphorae with sharp shoulder breaks—would have spread to Tel Zayit in the preceding tenth century BCE had the site already established ties to the coast. Their absence from the tenth-century-BCE stratum is, in this case, significant. It seems clear that a cultural (and probably a political) shift from one core to the other unfolded at Tel Zayit in the course of the ninth century. Given the borderlands location of the site, this development occasions no surprise. Radiocarbon dating places the ninth-century BCE destruction in the latter half of the century, apparently sometime after the 830s, when Assyrian pressure on the west lessened, thus allowing regional kingdoms to become more aggressive locally (see Tappy 2001: 516; 2006: 15, Table 4). The text of 2 Kgs 12:17–18 preserves notice of this event and links it to Hazael of Damascus. Tel Zayit itself may have become ensnarled in this conflict because, as preliminary ceramic analysis suggests, it had recently realigned itself more openly with coastal culture and, in fact, may have become a satellite of Philistine Gath. Both Edom and Libnah revolted against Judah once Jehoram succeeded his father, Jehoshaphat, as sole ruler in Jerusalem (2 Kgs 8:20–22a; 2 Chr 21:8–10a). From this point on, Naḥal Guvrin (the area of Judah’s lowlands District 4; see the discussion of Josh 15:42–44 above) and, undoubtedly, the city of Libnah maintained stronger ties with coastal polities than did any of the other valley districts or towns listed in Joshua 15. After the Israel-Judah-Edom coalition against Moab ended in failure and possibly also the murder of the Edomite crown prince, the Edomites apparently blamed their partners and proceeded to sever their political ties with Judah (Rainey 2006: 205). When Jehoram led his army to Judah’s eastern frontier to quell the Edomite uprising, he created an opportune historical moment for the erstwhile priestly city at Libnah, in Judah’s western liminal zone, to slip away from highland control. (Recall my earlier discussion of the multiple liminal zones that entities such as Judah had to manage along their various borderlands.) This shift likely occurred soon after the mid-ninth century BCE, in the 840s (see Rainey and Notley 2006: 205–6), and set the stage for the emboldened Hazael to vent his anger against both Gath and the newly affiliated Tel Zayit. The ninth-century-BCE pottery assemblage at Tel Zayit nicely parallels the repertoire discovered in the contemporary demolition of Stratum A3 at Tell eṣ-Ṣâfi (Maeir, ed. 2012: 11, 38ff.). Consequently, the larger city at Lachish garnered new status during the course of the ninth century BCE, and Judah’s principal representation shifted south and east to this recognized boundary city. Lachish never established commercial or cultural connections with the coastal plain. The differences that emerge between the material culture of Tel Zayit and that of Lachish in the ninth century BCE reflect both the location of Lachish along or within the more structured boundary of Judah and the transfrontier liminality of Tel Zayit. By the early-to-mid-tenth century BCE, then, Tel Zayit’s very existence in this area represented an early attempt by a cultural core that lay in the highlands to the east to establish a presence in its western lowlands. That someone in the new town wrote out a complete linear alphabet using a transitional script that in many

The Archaeology and History of Tel Zayit

169

Fig. 8.  The Tel Zayit abecedary: a liminal script (drawing by P. Kyle McCarter).

ways was distinct from earlier Phoenician letter forms (fig. 8) might well have represented a culturally symbolic act at a time when different entities were vying for presence in and influence over this area. In the next century, however, a shift toward coastal culture occurred at Tel Zayit, precisely at a time when a biblical notation indicates that Libnah rebelled against Jerusalem. Although the insurrection may have had many root causes, the event seems to have ignited a more extensive western threat from the Philistines and even certain other groups (possibly the Menuites; see 2  Chr 21:16–17). Although it had helped to open Judah’s western frontier, then, Tel Zayit ultimately provides a classic example of the precarious hold that the political center in Jerusalem wielded over its borderland communities. The Tel Zayit Abecedary is a vital link in an emerging series of epigraphic evidence from the tenth century BCE. In addition to a number of minor writings from this time, three significant inscriptions reflect the evolution of writing in southern Canaan. The series includes, from earliest to latest and starting in the late eleventh

170

Ron E. Tappy

century: the Khirbet Qeiyafa ostracon, the Tel Zayit Abecedary, and the Gezer Calendar (the recently discovered Jerusalem pithos inscription should stand at the early end of this sequence). Ironically, the Khirbet Qeiyafa inscription enjoys a clear archaeological context but remains quite difficult to read and interpret. The Gezer Calendar, on the other hand, is quite readable but lacks a datable findspot. Its placement within the series is based on paleographic comparisons. Although the Tel Zayit Abecedary does not present a narrative text with grammatical, morphological, syntactical, or historical detail, the inscription displays a fully developed, linear alphabet that derives from a very secure archaeological context and represents the inland script tradition of central and southern Canaan in the tenth century BCE. The letter-forms preserved on the stone provide a template against which to assess the script of other epigraphic finds, especially when those finds come from compromised contexts or lack any context. The paleography of the abecedary reflects an important moment in the evolution of alphabetic writing in southern Canaan (that is, in the emergent, patrimonial kingdom of Judah) at the outset of Iron Age IIA. Because some of the letter-forms echo older Phoenician styles while others both distinguish themselves from Phoenician and anticipate the emergence of the mature Hebrew national script of the ninth century BCE (fig. 8), the inscription reflects a transitional alphabet. In this way, both the script of the alphabet itself and the site of its discovery are liminal in nature. (2)  In the eighth century, both Uzziah and Hezekiah attempted to push sundry symbols of their culture as far as Timnah and, at times, even to Philistine Gath and possibly Ekron. But during the course of his third campaign, in 701 BCE, Sennacherib disrupted everything in the liminal zone of the Shephelah and the two cores that straddled it. The regional center at Lachish, which from its inception had displayed firm loyalty to its highland/Judahite core, suffered massive destruction, and Assyrian annals reveal that Sennacherib reassigned significant tracts of land in the nearby marginal zone from Judahite to neo-Philistine control. It is likely not a coincidence that the list of Shephelah districts in Joshua 15 follows the same order as Sennacherib’s line of march (Elah–Lachish–Guvrin—that is, north–south–central), judging from the best reconstructions of that somewhat elusive itinerary (compare Naʾaman 1974 and Younger 2003). The Assyrian leader surely knew that, in addition to Judah’s aggression against Padi at Ekron, cities in both the Elah and Lachish valleys either had come under increasing Judean pressure (Gath) or had shown steady loyalty to Judah (Lachish). Consequently, he had to deal with them first. The central Guvrin Valley, however, had maintained strong cultural ties with the more pro-Assyrian coastal plain at least since Libnah’s revolt against Judah roughly 150 years earlier. Sennacherib could afford, therefore, to leave that area last in his regional plan of attack. Ultimately, from the Assyrian point of view, Lachish emerged as the jewel of this campaign, judging from the magnificently carved wall slabs displayed in Room XXXVI of the royal palace in Nineveh. But what about the “forty-six .  .  . fortified walled cities and surrounding smaller towns” that Sennacherib claims to have

The Archaeology and History of Tel Zayit

171

conquered? (See Frahm 1997: 54, 59, lines 49–50; ARAB 2.§240 [compare §312]; ANET 288; Cogan 2000: 303). His royal annals record that he awarded these captured towns to the rulers of the very three Philistine cities that follow the Shephelah districts (and precede those of the hill country) in Joshua 15—that is, to Mitinti, king of Ashdod; Padi, king of Ekron; and Ṣillī-Bēl, king of Gaza (Frahm 1997: 54, 59, line 53; see also the ARAB 2, ANET, and Cogan 2000 references above). Younger (2003: 261) believes that some sort of Assyrian record,―perhaps the lost portion of the Azekah Inscription,―once existed, providing the names of the 46 towns. References to captives and foreign delegations appear in Assyrian records as early as the reign of Ashur-naṣirpal II (Banquet Stele; Wiseman 1952: 24–44, pls. 2–11; Grayson 1976: 172–76, no. 17, and, more recently, 1991: 288–93, no. 30), but the number of such groups increased considerably in the late eighth century BCE. Neo-Assyrian letters identify one of these groups as the Ṣīrāni (LÚ.MAḪ.MEŠ), “(tribute-bearing) emissaries,” who traveled annually to the Assyrian capital with gifts and mandatory payments from leaders in their native lands or cities. Both Ekron and Gaza appear in this correspondence, along with Judah, Moab, and Edom to the east (Postgate 1974: 124). Many of the place-names attested in these madattu (“tribute”) letters also appear in lists of wine rations sent to the Assyrian capital (Kinnier Wilson 1972; for a listing of critical reviews of this publication, see Russell 1991: 320 n. 31). Generally speaking, the Nimrud wine lists fall into an early group (ca. 800–780 BCE) and a later group (ca. 734–712 BCE; see Russell 1991: 234). During the last third of the eighth century BCE, the neo-Philistine triad of Ekron, Ashdod, and Gaza established and maintained contact with Assyria via unnamed officials who either delivered native wine from these cities to Nimrud or, given the relatively small quantities mentioned, themselves received rations of wine during their stay in the Assyrian capital (see wine list ND 10078 in Dalley and Postgate 1984: 246–47, no. 135; note also that the inclusion of Ekron is based on a reading of {an}-qa-ru!-⟨na⟩-a-a in line 13 of this list [1984: 247 n. 13]). Sennacherib’s dominance of the liminal zone, therefore, allowed him not only to isolate the hill country from the coast but also to reward the three specific Philistine centers (Ekron, Ashdod, Gaza) that had, for several decades, dispatched tribute- and wine-bearing envoys to Nimrud or Nineveh. Whether he also intended to lay waste the interior hill country (i.e., the Judahite core) at this time or simply wanted to blockade it while dealing with its western lowlands remains an interesting question. In the light of tactics employed by other neo-Assyrian kings, the latter scenario seems probable (Tappy 2007: 275–76). On the Judahite side of the ledger, following the events of 701, redactors responded by inserting these same three cities among the true districts of Judah as preserved in the older list of Joshua 15. The somewhat curious north–south–central ordering of the three original districts (Elah, Lachish, and then Guvrin) may even reflect knowledge of and adherence to Sennacherib’s line of march. Like the natural terrain and the events now unfolding, the biblical compilers ensconced the rebellious, pro-coastal Libnah district in

172

Ron E. Tappy

Fig. 9.  Architectural remains in Square K20.

a liminal position between the more established holdings of Judah and Philistia. These secondary textual changes probably transpired sometime after Sennacherib’s assault on the Shephelah borderlands, when a political realignment of unnamed towns occurred. But the historical grist for these emendations had been ground earlier by the emissaries of Sargon’s day and the field strategies and annalistic records of Sennacherib. In their counterclaim of ownership over towns and villages now lost to Ekron, Ashdod, and Gaza, the biblical redactors simply engaged in literary one-upmanship. (3)  As the centuries passed after Sennacherib’s devastating assault against the region in 701 BCE, Tel Zayit continued to live out a liminal life between larger cultural spheres. Following its occupation in the Persian and Hellenistic periods, the town experienced at least three phases of building during the Early Roman

The Archaeology and History of Tel Zayit

173

Fig. 10.  Segments of the Peutinger Map, from the Nile Delta to Antioch.

era. Then, in the Late Roman period, a huge building, apparently a fortress, commanded a view from the summit (fig. 9). The Tabula Peutingeriana (fig. 10; hereafter TP), a detailed but highly schematized presentation of the roads and staging posts of the Roman world, confirms again the strategic value of the borderlands around Tel Zayit. Considering the odd size of the TP (roughly 13 inches high by 22 feet in preserved length and possibly 28 feet in original length), it seems likely that this composite map served as a display piece in Rome sometime before 300 CE (Talbert 2010: 135–36, 144–45). Rainey and Notley (2006: 13) placed the map’s origin slightly earlier, in the second century CE; in any event, it seems to predate the spread of Christian influence under Constantine. The map presented a remarkable public exhibition of terrestrial routes that stitched together the imperium Romanum, that immense portion of the inhabited world that Rome dominated, from the Atlantic Ocean to India (or, as Ptolemy wrote in Geographia V.5.2, “that part of the earth which is inhabited by us . . .”). The display functioned as much more than a

174

Ron E. Tappy

utilitarian guide for travelers, couriers, and military officers (Talbert 2010: 142–57); the map itself surely stood as an impressive symbol of Roman power. Given the striking economy of vertical space and the dramatic compression of the entire Mediterranean basin, relatively tiny Palestine on the eastern littoral might easily have received only cursory attention. But by the time the Roman cartographer made his map, the province of Syria–Palaestina (formerly Iudaea) had garnered significant, albeit negative attention within Rome’s imperial system. Between the period of provincial prefects and procurators (ca. 6–70 CE) and the relatively peaceful but internally unstable Severan Dynasty (193–235 CE), at least five major historical cycles necessitated the development of an impressive road network in Palestine. These watershed events included: (1) the suppression of the First Jewish Revolt (66–70 CE); (2) Trajan’s (98–117 CE) establishment of the Provincia Arabia (106 CE); (3) the so-called Kitos War (115–117 CE); (4) Hadrian’s personal tour of Judaea (130 CE), when he rebuilt Jerusalem as Aelia Capitolina, prohibited Jewish citizens from entering the city, and issued decrees outlawing other Jewish practices; and, ultimately, (5) the Bar-Kokhba Revolt (132–135 CE). Thus, political and military affairs that unfolded over a sixty-year period and that focused on troubles in the eastern Mediterranean prompted the construction in Palestine of a well-developed and maintained system of roads, caravanserai, water facilities, guardposts, milestones, etc., by which the Romans both met their need for access and control and publicized their symbols of power (a goal admirably abetted by the milestones themselves; see Roll 1983: 153). Although at least nine major latitudinal routes traversed Palestine by the third century CE (Roll 1983: 145), the TP records only one that passed through the Judean Shephelah. By the early fourth century CE, Bishop Eusebius of Caesarea (Onomasticon) selected Eleutheropolis (Beth Guvrin) as the pivotal reference point from which to gauge distances to towns throughout the general region. The descent from Jerusalem to Beth Guvrin followed a longstanding route through the Elah Valley (between Socoh and Qeiyafa, southeast of Azekah, and southward to Beth Guvrin), and the most direct approach from Beth Guvrin out to Ashkelon still passed through the Tel Burna–Tel Zayit area—that is, through the former Libnah/Guvrin District 4 of Joshua 15. The Roman road itself is well preserved along a portion of this route. Since Tel Zayit represented the last (i.e., westernmost) town in the Naḥal Guvrin and lay near the transition from the Shephelah to the coastal plain, it would have provided the first and most exposed point of protection against any unwanted incursion from the coast. It is not surprising, then, that excavation has uncovered what appears to be a large Late Roman fortress on the western summit of the tell. As had all others before them, the Romans recognized the strategic importance of this liminal area. Preliminary analysis of the latest pottery recovered from beneath the principal floors of this stronghold suggests a construction date in the late second or early third century CE, certainly after the Bar-Kokhba rebellion and the rule of Hadrian and perhaps not until the rise of Eleutheropolis as a Christian center during the Council of Nicaea in 325 CE. In the decades following Hadrian’s

The Archaeology and History of Tel Zayit

175

rule, synagogues and, eventually, churches arose in this area, thus creating a liminality of religion, a meeting place for practitioners from related but different faiths.

III. Conclusion Across the longue durée, for nearly 4,000 years, those who lived and died at Tel Zayit enjoyed the benefits but also faced the challenges that resulted from calling a borderland community home. This kind of environment likely promoted a pragmatic acculturation more than rigid assimilation (e.g., Stone 1995; Faust and LevTov 2011). Those who lived there, who arrived from different and often opposing homeland cores, learned to feel at home with the familiar stranger. As the process of cultural fusion unfolded over long, disparate historical periods, the inhabitants of this liminal area exercised their hard-earned autonomy even as they nurtured a continuing, symbiotic relationship with the cultural cores that surrounded and interacted with them. Both actions typify liminal-zone groups. The archaeology of Tel Zayit helps bring to light such self-expression in the borderlands between Judah and Philistia. It also reveals times when imperial intruders (e.g., Egyptians, Assyrians, Romans) disrupted indigenous patterns and exploited this strategic region for their own purposes.

Bibliography Ackerman, S. 2005 When Heroes Love: The Ambiguity of Eros in the Stories of Gilgamesh and David. New York: Columbia University Press. Aharoni, Y., ed. 1973 Beer-sheba I: Excavations at Tel Beer-sheba, 1969–1971 Seasons. Publications of the Institute of Archaeology 2. Tel Aviv: Institute of Archaeology, Tel Aviv University. ANET = Pritchard 1969 ARAB = Luckenbill 1926–27 Assante, J. 2002 Sex, Magic and the Liminal Body in the Erotic Art and Texts of the Old Babylonian Period, Pp. 27–52 in Sex and Gender in the Ancient Near East, Proceedings of the 47th Rencontre Assyriologique Internationale, Helsinki, July 2–6, 2001, ed. S. Parpola and R. M. Whiting. Helsinki: University of Helsinki, The Neo-Assyrian Text Corpus Project. Barth, F., ed. 1969 Ethnic Groups and Boundaries: The Social Organization of Cultural Difference. Boston: Little Brown. Bellah, R. 2011 Religion in Human Evolution: From the Paleolithic to the Axial Age. Cambridge, MA: Harvard University Press. Bellah, R. N., and Joas, H., eds. 2012 The Axial Age and its Consequences. Cambridge, MA: Harvard University Press. Bloch, M. L. B. 1953 The Historian’s Craft. Trans. P. Putnam. New York: Knopf. 1974 Apologie pour l’histoire ou Métier d’historien. Paris: A. Colin. 1995 Histoire et historiens. Paris: A. Colin.

176

Ron E. Tappy

Braudel, F. 1972 The Mediterranean and the Mediterranean World in the Age of Philip II. New York: Harper and Row. 1980 On History. Trans. S. Matthews. Chicago: University of Chicago Press. Bunimovitz, S., and Lederman, Z. 1997a Six Seasons of Excavations at Tel Beth-Shemesh:―A City on the Border of Judah [Hebrew]. Qadmoniyot 30: 22–37. 1997b Beth-Shemesh: Culture Conflict on Judah’s Frontier. Biblical Archaeology Review 23/1 ( January/February): 42–49, 75–77. 2001 The Iron Age Fortifications of Tel Beth Shemesh: A 1990–2000 Perspective. IEJ 51/2: 121–47. Cogan, M. 2000 Sennacherib’s Siege of Jerusalem (2.119B). Pp.  302–3 in The Context of Scripture, Vol.  II: Monumental Inscriptions from the Biblical World, ed. W. W. Hallo and K.  L. Younger. Leiden: Brill. Cohen, A. 1985 The Symbolic Construction of Community. London: Ellis Horwood / New York: Tavistock. Cohen, A., ed. 1986 Symbolising Boundaries: Identity and Diversity in British Cultures. Manchester: Manchester University Press. Cole, J. W., and Wolf, E. R. 1974 The Hidden Frontier: Ecology and Ethnicity in an Alpine Valley. London and New York: Academic Press. Cross, F. M., and Stager, L. E. 1991 Ashkelon Discovered. Washington, D.C.: The Biblical Archaeology Society. Dalley, S., and Postgate, J. N. 1984 The Tablets from Fort Shalmaneser. London: The British School of Archaeology in Iraq. Dorsey, D. A. 1991 The Roads and Highways of Ancient Israel. Baltimore: The Johns Hopkins University Press. Eade, J. 2000 Order and Power at Lourdes: Lay Helpers and the Organization of a Pilgrimage Site. Pp. 51–76 in Contesting the Sacred: The Anthropology of Pilgrimage, ed. J. Eade and M. J. Sallinow. Urbana: University of Illinois Press. Faust, A., and Lev-Tov, J. 2011 The Constitution of Philistine Identity: Ethnic Dynamics in Twelfth to Tenth Century Philistia. Oxford Journal of Archaeology 30: 13–31. Fox, R. G. 1977 Urban Anthropology: Cities in Their Cultural Settings. Englewood Cliffs, NJ: Prentice-Hall. Frahm, E. 1997 Einleitung in die Sanherib-Inschriften. Archiv für Oreintforschung, Beiheft 26. Vienna: Institut für Orientalistik der Universität Wien. Jaspers, K. T. 1953 The Origin and Goal of History. London: Routledge and Kegan Paul. Luckenbill, D. D. 1926–27  Ancient Records of Assyria and Babylonia. Vols. 1–2. Chicago: University of Chicago Press.

The Archaeology and History of Tel Zayit

177

Maeir, A., ed. 2012 Tell es-Safi/Gath I: The 1996–2005 Seasons, Part I: Text. Wiesbaden: Harrassowitz. McCarter, P. Kyle 2008 Paleographic Notes on the Tel Zayit Abecedary, Pp.  45–59 in Literate Culture and Tenth-Century Canaan: The Tel Zayit Abecedary in Context, ed. R. E. Tappy and P. Kyle McCarter. Winona Lake, IN: Eisenbrauns. Mobley, G. 2006 Samson and the Liminal Hero in the Ancient Near East. Library of Hebrew Bible/Old Testament Studies 453. London: T & T Clark, 2006. Monroe, C. 2008 Liminality in the Eastern Mediterranean Maritime Sphere. Paper presented at the annual meeting of the American Schools of Oriental Research, Boston, MA. Naʾaman, N. 1974 Sennacherib’s “Letter to God” on his Campaign to Judah. Bulletin of the American Schools of Oriental Research 214: 25–39. Postgate, J. N. 1974 Taxation and Conscription in the Assyrian Empire. Rome: Biblical Institute Press. Pritchard, J. B. 1969 Ancient Near Eastern Texts Relating to the Old Testament. Princeton, NJ: Princeton University Press. Rabinowitz, D. 1998 National Identity on the Frontier: Palestinians in the Israeli Education System. Pp.  142–61 in Border Identities: Nation and State at International Frontiers, ed. T. M. Wilson and H. Donnan. Cambridge: Cambridge University Press. Rainey, A. F. 1980 The Administrative Division of the Shephelah. Tel Aviv 7: 194–202. 1983 The Biblical Shephelah of Judah. Bulletin of the American Schools of Oriental Research 251: 1–22. Rainey, A. F., and Notley, S. 2006 The Sacred Bridge: Carta’s Atlas of the Biblical World. Jerusalem: Carta. Reiner, E. 1967 City Bread and Bread Baked in Ashes. Pp. 116–20 in Languages and Areas: Studies Presented to George V. Bobrinskoy. Chicago: University of Chicago Press. Ritter, C. 1866 The Comparative Geography of Palestine and the Sinaitic Peninsula, trans.W. L. Gage. Edinburgh: T. & T. Clark. Roll, I. 1983 The Roman Road System in Judaea. Pp. 136–61 in The Jerusalem Cathedra: Studies in the History, Archaeology, Geography and Ethnography of the Land of Israel. Jerusalem: Yad Izhak Ben-Zvi Institute. Russell, J. M. 1991 Sennacherib’s Palace without Rival at Nineveh. Chicago: University of Chicago Press. Sallnow, M. J. 1981 Communitas Reconsidered: The Sociology of Andean Pilgrimage. Man ns16/2: 163–82. Stahl, N. 1995 Law and Liminality in the Bible. Journal for the Study of the Old Testament Supplement Series 202. Sheffield: Sheffield Academic Press.

178

Ron E. Tappy

Stone, B. 1995 The Philistines and Acculturation: Culture Change and Ethnic Continuity in the Iron Age. Bulletin of the American Schools of Oriental Research 298: 7–32. Talbert, R. J. A. 2010 Rome’s World: The Peutinger Map Reconsidered. Cambridge: Cambridge University Press. Tappy, R. E. 2001 The Archaeology of Israelite Samaria, Volume II: The Eighth Century BCE. Harvard Semitic Studies 50. Winona Lake, IN: Eisenbrauns. 2006 An Abecedary of the Mid-Tenth Century B.C.E. from the Judean Shephelah, Bulletin of the American Schools of Oriental Research 344: 5–46. 2007 The Final Years of Israelite Samaria: Toward a Dialogue between Texts and Archaeology, Pp. 258–79 in Up to the Gates of Ekron: Essays on the Archaeology and History of the Eastern Mediterranean in Honor of Seymour Gitin, ed. S. White Crawford, A. BenTor, J. P. Dessel, W. G. Dever, A. Mazar, and J. Aviram. Jerusalem: The W. F. Albright Institute of Archaeological Research and the Israel Exploration Society. 2008a Historical and Geographical Notes on the ‘Lowland Districts’ of Judah in Joshua 15:33–47. Vetus Testamentum 58: 381–403. 2008b East of Ashkelon: The Setting and Settling of the Judean Lowlands in the Iron Age IIA Period. Pp. 449–63 in Exploring the Longue Durée: Essays in Honor of Lawrence E. Stager, ed. J. David Schloen. Winona Lake, IN: Eisenbrauns. 2008c Tel Zayit and the Tel Zayit Abecedary in their Regional Context. Pp. 1–44 in Literate Culture and Tenth-Century Canaan: The Tel Zayit Abecedary in Context, ed. R. E. Tappy and P. Kyle McCarter. Winona Lake, IN: Eisenbrauns. 2012 The Tabula Peutingeriana: Its Roadmap to Borderland Settlements in Iudaea-Palestina, with Special Reference to Tel Zayit in the Late Roman Period. Near Eastern Archaeology 75/1: 36–54. Tigay, J. 1982 Pp. 212–13 in The Evolution of the Gilgamesh Epic. Philadelphia: University of Pennsylvania Press. Turner, V. 1964 Betwixt and Between: The Liminal Period in rites de passage. Pp. 4–20 in Symposium on New Approaches to the Study of Religion: Proceedings of the 1964 Annual Spring Meeting of the American Ethnological Society, ed. J. Helm. Seattle: American Ethnological Society. 1967 The Forest of Symbols: Aspects of Ndembu Ritual. Ithaca, NY: Cornell University Press. 1969 The Ritual Process: Structure and Anti-Structure. Ithaca, NY: Cornell University Press. 1977 Variation on a Theme of Liminality. Pp. 36–52 in Secular Ritual, ed. S. F. Moore and B. Meyerhoff. Assen: Van Gorcum. Van Gennep, A. 1960 The Rites of Passage, trans. M. B. Vizedom and G. L. Caffee. Chicago: University of Chicago Press. Orig., 1909. Younger, K. L., Jr. 2003 Assyrian Involvement in the Southern Levant at the End of the Eighth Century BCE. Pp.  235–63 in Jerusalem in Bible and Archaeology: The First Temple Period, ed. A. G. Vaughn and A. E. Killebrew. Atlanta: Society of Biblical Literature.

The Archaeology and History of Tel Zayit

179

Ussishkin, D. 2004 The Renewed Archaeological Excavations at Lachish [1973–1994], Vols. I–II. Tel Aviv: Emery and Claire Yass Publications in Archaeology of the Institute of Archaeology, Tel Aviv University. Kinnier Wilson, J. V. 1972 The Nimrud Wine Lists. London: The British School of Archaeology in Iraq. Wiseman, D. J. 1952 A New Stela of Aššur-naṣir-pal II. Iraq 14: 24–44, Pls. 2–11. Wolf, H. N. 1987 A Study in the Narrative Structure of Three Epic Poems: Gilgamesh, The Odyssey, Beowulf. New York: Garland. Zimhoni, O. 1997 Studies in the Iron Age Pottery of Israel: Typological, Archaeological, and Chronological Aspects. Tel Aviv: Tel Aviv University Press.

This page left intentionally blank.

Settlements and Interactions in the Shephelah during the Late Second through Early First Millennia BCE Ido Koch

1.  Introduction: Geography and Methodology The present volume is a welcome first attempt to amass current archaeological data on the Shephelah, including a dozen ongoing academic-affiliated excavations and numerous surveys and salvage excavations. The published data accumulated over the past decade is now at a point where it requires a revision of some of the commonly accepted historical reconstructions and an elaboration of others. But I am putting the cart before the horse. First, I would like to make some introductory remarks regarding geography and methodology. The term “Shephelah” is a Highlander’s perspective on the low hills located to the west of the Highland. It reflects the drop in the terrain of the western slopes of the Highland in a series of valleys (from the Ajalon Valley in the north to the Yaval Valley in the south) formed on the contact line with the lower hills to their west. This distinction, though, should not deter us from recognizing the fact that both regions are topographically connected, particularly in two parts: 1. The Valley of Ajalon separates the Shephelah from the spurs of the Highland (the Ramallah anticline) but also serves as a natural traffic route connecting both regions (via the “Beth Horon Ascent”) (Magen and Finkelstein 1993: 19–20) and the Lydda Valley to the northwest. 2. The numerus intermittent streams (Hebrew: naḥal) that drain the highland from south of Bethlehem to Hebron and lead to the Shephelah. This is especially clear to the east of HaʾElah Valley, where the ꜤArqob—a rugged, stony terrain that is lower than the Highland yet higher than the Shephelah hills—serves as a natural topographical stairway (Amiran 1948: 112–15).

In contrast to this, the distinction between the Shephelah and the Coastal Plain to its west is rather elusive, with no topographical obstacle visible in the local landscape (Dan 1988: 50; Stern 1988: 12). In other words, while the region can be characterized vis-à-vis the Highland or the coast, the connections between all three regions are evident. In what follows, I would like to reassess these connections from the Late Bronze Age II through the Iron Age IIA. 181

182

Ido Koch

My second introductory issue is related to methodology. I have chosen to present the immense stratigraphic information from the region (and relevant references) in Table 1 in the Appendix. In this table, I make a distinction between data from published excavation reports and data that was published in preliminary form and/or was gained from only limited excavation. The latter data is partial and fragmentary and makes it difficult to gain insight into the continuity of settlement and the precise chronology of various phases. Another problem in tables such as this and any other kind of presentation of stratigraphic data is that these are artificial and static assemblies of sites that existed in “archaeological periods”—which are a modern scholarly projection into the past—rather than a precise reflection of a particular historical reality. Some of the sites may have been founded at the beginning of a “period” and abandoned shortly thereafter, while others may have been established at the end of that period and existed mainly in the period that followed (Finkelstein 1996: 227). These remarks are even more critical when dealing with survey data (Table 2 in the Appendix). It is noteworthy that there is a fundamental difference between areas that have been comprehensively surveyed, producing survey maps, and those in which surveys were partial or were focused on specific sites. The difference manifests itself in the appearance of a cluster of sites in one survey area as opposed to a paucity of sites beyond an imaginary line between one survey area and a neighboring area. Differences can also be seen in the surveys themselves, usually stemming from differing research methods, intensity of research, surveyors’ prior knowledge, manner in which data was amassed and interpreted, and consideration of environmental elements, such as modern buildings, thick woodland or dunes covering the ancient surface—all of which make identification of early archaeological remains difficult (Kowalewski 2008: 249–57; Sharon and Dagan in Dagan 2011). The data that emerges from archeological surveys are in essence biased; obviously, not every sherd receives equal consideration, and settlement periods are therefore occasionally omitted from the documentation of some sites (Garfinkel and Ganor 2010: 69–72). Since surveys are superficial by nature, it is difficult to identify actual settlements, given that the starting point might be no more than a “spot” of sherds, sometimes attesting to temporary activity or agricultural activity dating from later periods (pottery brought from nearby sites in order to fertilize fields) (Faust and Katz 2011: 233 n. 4). In light of these considerations, discussion in this essay does not interpret survey data as attesting to a settlement pattern (and hence not to the extent of a settlement or its population). Instead, the data is interpreted as attesting to the distribution of human activity associated with settlements that have been identified in archaeological excavations. With the above in mind, I would like to review four issues relevant to the Late Bronze Age through Iron Age IIA in the Shephelah: (1) the rise of Tel Miqne as the chief center during the Iron Age I; (2) the contemporaneous changes in local settlement patterns; (3) the interregional interaction during the Iron Age I; and (4) the increase in number of sites in the region during the Iron Age IIA.

Settlements and Interactions in the Shephelah

183

2.  Tel Miqne and Its Neighbors during the Late Bronze Age and Iron Age I When plotting the archaeological data, it appears that the most prominent process in the entire region of the Shephelah is the continuous founding and refounding of local centers at the very same sites from the Middle Bronze Age II to the Iron Age II. Clearly, not all the centers were established simultaneously or even existed concurrently, and the migration of a center’s settlement gravity can be traced from one site to a neighboring site. This was the case of Tel Miqne and its neighbors, Tel Gezer to the north, Tell eṣ-Ṣafi to the south, and Tel Beth-Shemesh to the east (Bunimovitz and Lederman 2006; Maeir and Uziel 2007; Naʾaman 2010: 501; Uziel, Shai, and Cassuto 2014). In this seesawing, power shifted from Middle Bronze Age II Tel Miqne to its neighbors in the Late Bronze Age, concentrated at Tel Miqne during the Iron Age I (with a growth in the size of Tell eṣ-Ṣafi during the course of the period), shifted once again to the neighbors in the Iron Age IIA, and back to Tel Miqne in the later phase of the Iron Age IIB and predominantly in the Iron Age IIC. In what follows, I would like to elaborate on this shift from the Late Bronze Age IIB to the Iron Age I. Beginning in the Late Bronze Age, the archaeological evidence for settlement history of the region during the Late Bronze Age II is well known, and combined with written evidence—mainly the Amarna correspondence (for the most recent publication, see Rainey 2015)—it is possible to suggest a gradual development of several sites during the long period of Egyptian hegemony, predominantly Lachish, Gath, and Gezer (Naʾaman 2011; Finkelstein 2014). Individuals and groups from these locales and others were highly integrated in an Egyptian-oriented system that included intensive interaction with the Egyptian court and its representative at Gaza, Jaffa, and other installations (Koch 2014, 2016). Based on archaeological evidence from Tel Lachish and Tell eṣ-Ṣafi, combined with 19th Dynasty written attestations of Gezer (Morris 2005: 377–81), it can be suggested that all three were local centers of power during the 13th century BCE as well. By the Late Bronze Age IIB, most settlements in the region were destroyed or abandoned (fig. 1a). When plotting the stratigraphic information dating to the Late Bronze/Iron Age Transition 1 (fig. 1b) it is clear that a monumental change was taking place in the region: revitalization was partial at best, since only Tel Lachish Level VI, and perhaps also Tel Azekah, were considered as local centers of power. All other sites were settled only to a limited extent. These two were soon destroyed, most probably in the wake of the withdrawal of the Egyptians from the region and the collapse of the social structure that relied on them: it appears that settlement at Tel Lachish was disturbed during the period when a temple located on the summit of the mound was sacked and a public building along the western slope was colonized by squatters sometime prior to destruction of the town (Ussishkin 2004: 71). Current evidence from Tel Azekah indicates a sudden and violent destruction during 1.  Otherwise termed Iron Age IA or Late Bronze Age III (Martin 2011: 20)

184

Fig. 1.  The Shephelah and nearby regions: Settlement patterns during a. the Late Bronze Age III; b. Transition Late Bronze Age/Iron Age; c. Iron Age I.

Ido Koch

Settlements and Interactions in the Shephelah

185

the Late Bronze/Iron Age Transition. Both sites feature meager evidence of activity during the Iron Age I. It was only sometime later that Tel Miqne grew in size from the village located on the 4-ha mound (Stratum VIIIA) to a large center—a city, in local terms, covering the entire 20-ha site (Stratum VIB) 2 (fig. 1c). Various public buildings, cultic activity, workshops with specialized production that reflects nascent industries, and a varied material culture all attest to social differentiation. The excavators (e.g., Dothan and Gitin 1993: 1053–54) have also argued that the city was fortified from the start, surrounded by a massive mud-brick wall, but this conclusion has received scholarly criticism (Ussishkin 2005). The large settlement at Tel Miqne evolved during the Iron I (Strata VIB–IVA) and survived until its destruction in the early 10th century BCE (Dothan and Gitin 1993: 1056). Other settlements feature continuity and a gradual development (fig. 1c). These include Tel Batash Strata V, Tel Ashdod Strata XIIIA–XIA, Tel Beth-Shemesh Strata 6–4, Tel Yarmout Acr5–3, Tel Gezer Strata XIII–XI, and predominantly the expansion of Tell eṣ-Ṣafi, where remains dating to the late Iron Age I settlement are found on the entire mound. In light of the limited remains from a majority of sites in the region, it has been argued that Tel Miqne was the chief settlement (Finkelstein 2007: 520–21; Yasur-Landau 2007: 615). I would argue that the concentration of power at Tel Miqne should be seen against the breaking down of the previous social structure that characterized the region under Egyptian hegemony following the demise of the elites residing at Tel Gezer, Tell eṣ-Ṣafi/Gath, and possibly also Tel Beth-Shemesh. There was no revitalization of a similar social structure, as indicated by the rural character of these sites and their neighbors, during the Late Bronze/ Iron Age transition. Thus, the changes in the settlement pattern reflect the concentration of power in a center that was different from those flourishing in the days of Egyptian hegemony. The demise of the traditional elites led to the emergence of a new social network unrelated to the previous centers of power. The rise of new group(s) at Tel Miqne was possibly contemporaneous with the decline in Egyptian activity, yet the site’s real window of opportunity appeared following the final Egyptian withdrawal. This is not to say that these years were peaceful; on the contrary, four of Tel Miqne’s neighbors (Tel Gezer, Tel Ashdod, Tel Batash, and Tel Beth-Shemesh) were destroyed or abandoned at least once during the 11th century BCE; all except Tell eṣ-Ṣafi, including Tel Miqne, feature evidence of destruction or abandonment during the terminal phase of the late Iron Age I. 3

2. Preliminary publications of the excavations (e.g., Mazow 2005: fig. 3.2a–3.2i; 2014: figs. 3–5) indicate that, although the lower terrace at Tel Miqne was already settled in Stratum VII, architecture became dense from Stratum VI onward. 3.  For competition over resources as a possible factor contributing to rivalry among the towns in the region, see Uziel et al. 2014: 304.

186

Ido Koch

This settlement pattern was therefore completely different from that of the Late Bronze Age II, when power was distributed among several centers, but quite similar to the Middle Bronze Age II, when a large center at Tel Miqne (Dothan and Gitin 2008: 1953) was surrounded by smaller yet fortified settlements at Tel Gezer (Dever, Lance, and Wright 1970: 18–19, 41–45; Dever et al. 1974: 28–35; Dever 1993: 500–501), Tell eṣ-Ṣafi (Uziel 2008: 203–14; Maeir 2012a: 16), and Tel Beth-Shemesh (Bunimovitz and Lederman 2008: 1645; 2013). In this case, one would wonder: what is the “normal” structure in the region: several medium-sized power hubs, as in the Late Bronze Age II, or rather a large dominant settlement, as in the Middle Bronze Age II and again in Iron Age I?

3.  Settlement Patterns from the Late Bronze Age to the Iron Age I A second prominent feature of archaeological conversation regarding the Shephelah in the Bronze and Iron Ages claims that there was “an almost complete abandonment of the countryside” during the Iron I (Finkelstein 1996: 232). The greatest decline was in the hinterland of Tel Lachish and Tell eṣ-Ṣafi; but even there, surveys have recorded a handful of sites with Iron Age I remains. It may be presumed that these areas were heavily affected by the destruction of their local centers but were not yet entirely desolate; they were rather sparsely settled and probably were used mainly for grazing. 4 The limited number of sites were clustered in a few areas (see fig. 2). Both Tel Batash and Tell eṣ-Ṣafi, the areas closest to Tel Miqne, flourished during the Iron Age I. Tel Batash enjoyed a short period of prosperity until it was abandoned sometime in the late Iron Age I; Tell eṣ-Ṣafi grew during the transition to the Early Iron Age IIA. A survey south of Tell eṣ-Ṣafi recorded Iron Age I remains at a handful of sites. To the west and southwest, Tel Ashdod and Ashkelon remained settled much as they had been in the Late Bronze Age II; the decline in population in their vicinity is seen only around the former Egyptian-oriented installations at Tel Mor and Ashdod Yam (S) (Berman, Barda, and Stark 2005). The Iron Age I town of Tel Gezer developed gradually (HUC Strata XIII–XI), featuring large structures with several phases (and possibly several disruptions) and a fortification wall (Ortiz and Wolf, this volume). Following the abandonment of the site, it was reoccupied in the Late Iron Age I (HUC Strata X–IX) until it was destroyed. According to survey data, a decline in the hinterland around the Ajalon Valley (from 19 to 15 sites) was felt mostly in the rugged hills to the north. 5 Conversely, the number of sites in the hills around the eastern part of the valley remained static. Further eastward, long, narrow spurs of the Highland (the “Beth 4. A comparable regional phenomenon is well documented in the Ottoman period (Grossman 1994: 201–3). 5.  This decline was perhaps connected to the decline in activity farther to the north, in the Lydda Valley (Gophna and Beit-Arieh 1997), on the way from Tel Gezer to the now destroyed Jaffa. (On Jaffa during the Iron Age I see Kaplan and Ritter-Kaplan 1993: 656, 658.)

Settlements and Interactions in the Shephelah

Fig. 2.  Excavated sites and survey spots in Southwest Canaan: a. Late Bronze Age II and b. Iron Age I.

187

188

Ido Koch

Horon Ascent”) led to the al-Jib Plateau, where intensification in human activity during the Iron Age I, compared to the Late Bronze Age II, is evident. 6 Tel Beth-Shemesh featured gradual growth and accumulation of wealth during the Iron Age I (Stratum 6); this was interrupted and followed by a renewed occupation that dates to the Late Iron Age I (Strata 5–4). Farther to the east, the Late Bronze Age village at Tel Yarmout experienced similar growth until it was abandoned in the Late Iron Age I (Strata Acr-6—3). Surveys in the region documented two Iron Age I sites. Farther southeast, the moderate ascents east of HaʾElah Valley allow direct access from the Shephelah to the Highland between Bethlehem and Halhul, where Kh. Tubeiqa was the sole Late Bronze Age II site (Sellers 1968; Funk 1993). Some 7 km to the southeast is Hebron (Tell er-Rumeida), where Late Bronze Age II pottery was recorded (Eisenberg and Nagorski 2002; Eisenberg and Ben-Shlomo 2016). A rich, contemporaneous burial cave was found at Kh. Jadur (Ben-Arieh 1981). Iron Age I pottery is recorded from these three sites as well as from eight additional survey sites, most of which are located on the narrow plateau and the western slopes of the Highland (Ofer 1993). The Late Bronze Age II town of Tel Beit-Mirsim, farther to the south, gradually declined through the Iron Age I (Strata B1–B3). Tel Eton was settled during the same period, but not much is known at this time about the settlement. These two should be described alongside Tel Ḥalif (connected to Tell Beit-Mirsim via the Yaval Valley), where settlement continued with no interruption from the Late Bronze Age IIB through the Iron Age I. To the east, the upper intermittent streams of Naḥal Shiqma and Naḥal Adorayim lead through the southwestern slopes of the Hebron Hills to Kh. Rabud (ca. 10 km southwest of Hebron), where limited excavations unearthed the remains of settlements dating to the Late Bronze Age IIB and the Iron Age I (Kochavi 1974). Four more sites southwest of Hebron yielded Late Bronze Age remains, and two other sites yielded Iron Age I remains (Ofer 1993). To conclude: the Iron Age I settlement pattern reveals the existence of several clusters of sites. The most prosperous of all were the settlements in the Lower Shephelah, Tell eṣ-Ṣafi and Tel Batash, in close proximity to Tel Miqne. Other sites in the region feature more moderate growth, while some experienced destruction. Apparently, almost all sites in the central and northern Shephelah except for Tell eṣ-Ṣafi suffered distress during the late Iron Age I. Inland clusters of sites in the Shephelah (east of Tel Gezer and Tel Beth-Shemesh) were located close to the two topographic transition zones that connect their valleys with the Highland (Beth 6. A disturbed Late Bronze Age II burial was unearthed at al-Jib (Pritchard 1963), less than 15 km from the sites adjacent to the valley (e.g., Yalu and Beit Ur at-Taḥta). The closest settlements were at Beitin (Finkelstein and Singer-Avitz 2009: 37) and Jerusalem (Finkelstein, Koch, and Lipschits 2011: 11; Maeir 2011: 179–82), and Late Bronze Age pottery was recorded at Deir el-ꜤAzar (Cooke 1923: esp. 115) and Qalunia (Kisilevitz et al. 2014). By the late Iron Age I, there were settlements at Beitin and Jerusalem and they were joined by dozens of sites around the previous settlements (Magen and Finkelstein 1993; Sergi 2015).

Settlements and Interactions in the Shephelah

189

Horon Ascent and the upper intermittent streams of Naḥal HaʾElah). In the Highland, human activity at the few sites that existed during the Late Bronze Age IIB— among them Beitin and Kh. Tubeiqa—intensified during the Iron Age I, part of a settlement process that reshaped the local landscape during the 11th century BCE. Pinpointing the proximity of these clusters of sites to the Highland, where intensification of human activity is recorded at sites located near the topographic transition zones, might assist in the search for the causes of the depopulation of the Shephelah: while difficult to prove archaeologically, the turmoil in the late 13th century BCE may have caused the weakening of local society and may have led peasants to search for refuge in other regions. Another scenario, based on the assumption that the Philistines resided in large urban centers of the “Pentapolis,” credited them for a policy of synoecism that forced the rural population to relocate and concentrate in a few sites (Bunimovitz 1998: 107–8; Shavit 2008: 160; Faust and Katz 2011: 235–36). This reconstruction has to be revised in light of current data pointing to a similar magnitude of Late Bronze Age and Iron Age I settlement remains at Tel Ashdod, Ashkelon, and Tell eṣ-Ṣafi, while the only settlement that dramatically grew was Tel Miqne. A revised scenario would be that the depopulation of the countryside reflects changes in land tenure, the extension of control over land by the elite located at Tel Miqne, and the subordination of the less powerful groups that were relocated close to their patrons. At the same time, the rise of a new social structure at Tel Miqne could have actually been a pull-factor, as people were attracted by the security and opportunities a large settlement could offer (cf. Yoffee 2005: 60).

4.  Interregional Interactions The excavations of Tel Batash and Tell eṣ-Ṣafi show that the Iron Age I material remains found there are similar to those found at Tel Miqne and other coastal sites. These remains have traditionally been understood as part of a monolithic Philistine material culture (Dothan 1982; Stager 1995). Nonetheless, when the various elements constituting this “culture” have been scrutinized (e.g., Hitchcock and Maeir 2013; Maeir, Hitchcock, and Horwitz 2013; Mazow 2014: 132–34, with literature), it has become clear that they exhibited a wide range of “homelands,” that they had appeared gradually at various sites, that they had been produced in different sizes and shapes, and that they had been used, sometimes, in diverse ways. Consequently, these remains can be understood in light of the overall changes that are visible in the archaeological record—reflecting a regional interaction network that included the large center at Tel Miqne, the harbors along the coast, and the towns in the lower Shephelah. The intensification of the local economy, best seen in the specialization in animal exploitation 7 and perhaps also in the 7.  Faunal assemblages from Tel Miqne (Lev-Tov 2010: 96 table 7.3), Tell eṣ-Ṣafi (Lev-Tov 2012: 594 table 28.2), Tel Batash (Panitz-Cohen and Mazar 2006: 311 table 81), and Tel Ashdod (Maher 2005: 285 table 8.3) reflect intensification in specialization trends in animal

190

Ido Koch

development of a textile industry (Mazow 2006–2007, 2013), reflects another aspect of the regional interaction that encompassed the coast and the lower Shephelah. Thus, the shared material remains can be interpreted as the remains of intensive regional interaction, reflecting a process in the making, an amalgamation of local developments alongside innovations in production and consumption shared by several communities, some possibly newcomers, that selectively adopted and adapted practices and concepts. Much attention has been given in scholarship to the most visible remains of this interaction, the “Philistine” Bichrome pottery (Dothan 1982; Ben-Shlomo 2006; Dothan, Gitin, and Zukerman 2006; Gilboa, Cohen-Weinberger, and Goren 2006; Ben-Dor Evian 2012). It has been recorded at numerous sites from Tel Qasile and Tel Aphek along the Yarkon River to Tel Jemmeh along Naḥal HaBesor, in all excavated sites on the Coastal Plain and in the Shephelah, with a handful of sites in the Highland (Tell en-Naṣbeh, Beitin, and Kh. Tubeiqa). The highest frequencies of use of the Bichrome pottery are recorded at Tel Miqne, Tel Ashdod, Tell eṣ-Ṣafi, Tel Batash, and Ashkelon, alongside Tel Qasile. This family of forms and decorative styles was developed through interaction between local traditions and foreign styles of forms and decorations, first in workshops located at Tel Miqne and Tel Ashdod (Gunneweg et al. 1986), later spreading to additional sites: one workshop was located at Tel Qasile (Yellin and Gunneweg 1985), producing vessels similar in form to those produced at Tel Miqne or Tel Ashdod, yet featuring some dissimilarities in decorative styles (Mazar 1985: 103–4). A second was located at Tel Jemmeh, producing a limited range of types (mainly bowls, kraters, and a small number of jugs) decorated with a limited range of motifs, compared to assemblages from Tel Miqne or Tel Ashdod (Ben-Shlomo and Van Beek 2014: 722–27, 793). A third was located at or near Tell en-Naṣbeh (Gunneweg et al. 1994), producing a similar, restricted range of types but only for a short phase before the Late Iron Age I (Dothan 1982: 55). It has been suggested that the emphasis placed in the Bichrome assemblage on serving and eating/drinking vessels reflects its role in communal feasting (Maeir 2008; Hitchcock and Maeir 2013; Mazow 2014: 146–47; Faust 2015). Consequently, its limited appearance at inland sites has been interpreted as the remains of a local elite that possibly associated itself with coastal feasting practices (Faust 2015: 180). This appealing hypothesis must, however, wait for a detailed study of a coherent ceramic assemblage from a secured archaeological context before it can be accepted. For the time being, the distribution pattern of the Bichrome pottery can be seen as reflecting interregional interaction networks whose focal points were the prosperous towns located on the Coastal Plain and in the Lower Shephelah. economy through the Late Bronze Age II to the Iron I (also Sapir-Hen, Gadot, and Finkelstein 2014: 715) that included herding for secondary products and further reliance on efficient use of cattle in cultivation of nearby fields.

Settlements and Interactions in the Shephelah

191

Thus, the development of the settlements in the Lower Shephelah and on the Coastal Plain brought about an intense web of economic contacts, leading to a wide distribution of shared material remains among major sites in these regions. The patterns of consumption and production of the Bichrome pottery attests to this intense network alongside other modes of interaction with neighboring regions, among them the inland Shephelah sites. Along these routes of interaction, ideas and practices were exchanged through unknown mechanisms of knowledge-transfer (relocation of potters from the “homeland” to the margins, exchange of ideas and know-how, or from a combination of other factors), including a short phase of production of Bichrome pottery on the al-Jib Plateau.

5.  The Growth in Number of Sites in the Iron Age IIA Another modification in the common historical reconstruction is required as a result of the mounting archaeological data amassed in recent decades: the interpretation of the increase in sites in the Iron Age IIA in the Shephelah as reflecting migration from the Highland, from east to west, as part of the westward expansion of the Kingdom of Judah toward the fertile lands of the Shephelah (Dagan 2000: 201, 2007: 29). The argument has also been put forward that settlement in the Shephelah increased so drastically that it was clearly not a natural process—that is, the additional settlements cannot be explained by a natural increase in the meager local population (Faust 2013a: 209). Some doubts obviously arise regarding these conclusions in light of the above-mentioned development of the settlement in the lower Shephelah on the one hand and in the upper Shephelah and the Highland on the other. For example, it is indeed possible that the clusters of sites in the upper Shephelah and in the Highland expanded due to natural growth or migrations. Yet, it is equally likely that demographic pressure and pursuit of supplementary agricultural products would have been factors in the much more heavily populated lower Shephelah. The extensive remains at Tell eṣ-Ṣafi reflect the concentration of people and wealth at this single site sometime after the destruction of Tel Miqne in the late Iron Age I, and eventually it became the primary hub in the region until its destruction in the late Iron Age IIA. Thus, the role of groups residing at Tell eṣ-Ṣafi during the Iron Age IIA should not be underestimated in historical reconstructions of this period in the Shephelah. There are additional methodological considerations that impede an overarching scenario of Judahite colonization of the Shephelah. First, the reconstruction of settlement processes based on generalized information such as archaeological survey data might be essentially mistaken. A case in point is Kh. Qeiyafa. This fortified settlement existed for a brief period at the beginning of the Iron Age IIA; if not for excavation of the site, it might have been concluded from the archaeological survey that a “fortified city” existed there throughout the Iron Age II (Dagan 2000:

192

Ido Koch

Site 12-14/62/1). Furthermore, ceramic finds at survey sites do not necessarily attest to the existence of settlements in any specific period. A clear example of this is Tel Yarmout, which was documented in the Shephelah survey as a fortified city in the Iron Age II (Dagan 2000: Site 14-12/74/1) but was shown by excavation to lack occupation at that time. The conclusion is that quite a few survey sites attest to agricultural activity only or, possibly, that the “spots” of sherds found reflect only later soil fertilization efforts. Other difficulties stem from the manner in which excavated sites are discussed and conclusions are drawn. Artificial geographical separation between excavated sites and neighboring, still-unexcavated sites could result in arriving at misleading conclusions. 8 Automatically associating the sites of the Shephelah with the Kingdom of Judah because of the “similarity between the material culture of the relevant strata and that of Judah, and the sharp differences between it and the coastal plain” (Faust 2013b: 210) is contrary to the facts; only two sites have been fully published (Tel Lachish and Kh. Qeiyafa). Discussions of ceramic types (Shai and Maeir 2003; Gitin 2006; Sergi et al. 2012), writing traditions (Byrne 2007; Finkelstein, Sass, and Singer-Avitz 2008; Finkelstein and Sass 2013), and burial practices (Fantalkin 2008) indicate continuity of practices unique to the Shephelah through the Iron Age IIA (Naʾaman 2010; Lehmann and Niemann 2014). Hence, the association of the early settlements from the Iron Age IIA with the Kingdom of Judah should be seen as a historical assumption rather than an archaeological fact. Another aspect is the variety of possibilities in the fluctuations in settlement patterns over a single archaeological period. For example, the destruction of Tell es-Ṣafi/Gath and nearby settlements during the late Iron Age IIA (e.g., Tel Zayit, Tel Ḥarasim, Tel Goded; see Sergi 2013: 228) could have led to the uprooting of the rural population in this district to move to the eastern Shephelah and the Coastal Plain. Fluctuations in settlement patterns no doubt also stemmed from other timeand site-sensitive causes; without detailed written sources, these causes will remain a mystery. An example of this is the expansion of settlement during the Roman period in the broad, fertile plain of ʿAmuq (“the plain of Antioch”), which apparently stemmed from changes in land-ownership traditions; a hierarchy controlled by a central government, documented in the Alalakh archive from the Late Bronze Age, grew into a privatized system that allowed discrete settlement throughout ʿAmuq and which is described in written sources from the Roman period in Antioch (Casana 2007: 212–14). Another test case is the settlement pattern in Palestine in 8. This is particularly obvious in the case of the southeastern valleys and the Hebron Hills. The three mounds in the valleys have been excavated and mostly published. In contrast, the information in hand regarding the nearby Highland sites is meager: both Hebron and Kh. Rabud have been partly excavated and only published in preliminary form; no coherent stratigraphic sequence has been suggested and no detailed pottery comparisons have been made. Thus, the frequent distinction between these sites and the sites located in the southeastern Shephelah valleys based on assumed differences in material remains rests on shaky ground.

Settlements and Interactions in the Shephelah

193

the Ottoman Period (similar in time-span to the Iron Age II), which, for a variety of reasons, changed more than once. Thus, for example, rural settlers abandoned their homes and migrated to nearby hills and distant regions seeking closer proximity to water sources or fertile fields, economic opportunities afforded by a growing urban center, attempting to evade taxes, or to avoid rivalries with other villages, raids, or environmental problems such as the creation of marshes (Grossman 1994). In conclusion, the data I have presented require detailed discussions of the settlement processes in the Shephelah in the Iron Age IIA. These discussions must encompass the data from numerous excavations, which will enable a comparison between the material culture in the Shephelah and sites far and near. The archaeological data, together with anthropomorphic, ethnological, and environmental considerations, will then serve as a foundation on which the historical reconstruction will also be based.

6.  Concluding Remarks Taking these data and this analysis into consideration, it is possible to suggest the following reconstruction. Fluctuations in the extent of human activity indicate the demise of the previous social system and the emergence of a new one. The Egyptian-oriented structure of several local centers located in the lower Shephelah—Tel Gezer, Tell eṣ-Ṣafi, Tel Lachish, and possibly also Tel Beth-Shemesh—was destroyed during the Late Bronze Age IIB and was eventually replaced by a sparse, fragmented structure with one major site: Tel Miqne. Other parts of the Shephelah were less prosperous, demographically and economically. These include the hinterlands of Tel Gezer and Tel Beth-Shemesh, both connected to the Highland, where settlement was intensified during Iron Age I. The processes leading to this fragmented landscape could have been the removal of the local centers of power (thus impeding the rural population from staying in these areas), the relocation of locals to other, perhaps safer regions, and push-and-pull factors circulating around the large center created at Tel Miqne. Yet, the gaps between one settled area and the other are small, no more than 20  km and usually much less. Thus, it is not surprising that interregional interaction is evident across this part of the landscape. Remains from Tel Miqne and nearby Tell eṣ-Ṣafi and Tel Batash suggest that their inhabitants interacted through a dense network with the entire Coastal Plain from Tel Qasile to Tel Jemmeh. Other Shephelah sites were engaged in this network more sporadically, as indicated by the limited appearance of coastal traits such as the Bichrome pottery. The more limited distribution of such remains at Tell en-Naṣbeh, Beitin, and Kh. Tubeiqa suggests that the highlander population was also part of this interaction, although on a more diminished level. This network of interaction was traumatized several times during the Iron Age I by the destruction of several sites. It came to an end in the late Iron Age I with the devastation of various sites in the region but was revived in the Iron IIA as

194

Ido Koch

evidenced by the emergence of new structures. Sites affected include Tell eṣ-Ṣafi, the largest settlement in the country, and smaller sites that were resettled, as well as dozens of spots that appear in surveys all across the Shephelah, reflecting an intensification of human activity. All of these data, except for the emergence of Tell eṣ-Ṣafi, have been connected in one way or another with the Kingdom of Judah or at least with colonization of the region by a highland population. Although it is indeed possible that some sites were founded by highlanders looking for better agricultural products, any reconstruction of the settlement process in the Shephelah should also include the possible role played by residents of Tell eṣ-Ṣafi. To conclude, there are several consistent trends in the Late Bronze Age IIB through Iron Age IIA period that emphasize the place of the Shephelah as part of a macro-region termed “Southwest Canaan”: the persistence of the local centers at the same sites; the centrality of those located in the lower Shephelah and the relatively limited population in the upper Shephelah; and the interaction between the Shephelah and neighboring regions, mainly the networks focused on the Coastal Plain and the lower Shephelah with the more limited participation farther inland as far as the Highland that lasted, with various modifications, throughout these years (and to some extent also in the Iron Age IIB–C 9). Thus, an overarching historical scenario suggested for the Shephelah during the Late Bronze and Iron Ages must recognize the archaeological remains from Southwest Canaan as a whole while acknowledging the particulars of each region.

Acknowledgements I would like to thank the editors, Aren Maeir and Oded Lipschits, for inviting me to contribute to this volume. This paper derives from my Ph.D. dissertation, Southwestern Canaan during the Late Bronze Age and the Early Iron Age: Empire, Elites and Colonial Encounters, written under the supervision of Nadav Naʾaman and Oded Lipschits of Tel Aviv University. I extend to both of them my gratitude for commenting on earlier drafts of this paper. Many thanks also go to Myrna Pollak. Work on this paper was conducted while I was a postdoctoral fellow at the Religions­ wissenschaftliches Seminar, University of Zurich (2015–2016). Special thanks are extended to Christoph Uehlinger. 9.  See, for example, the interaction between Ashkelon, the Shepehlah, and the Highland in the Iron Age IIC (Faust and Weiss 2005; Lipschits, Sergi, and Koch 2011: 27–28)

VIIIB Acr-6

Tel Miqne (Mazow 2005; Meehl, Dothan, and Gitin 2006; Dothan and Gitin 2008; Killebrew 2013; Mazow 2014)

Tel Yarmuth (Miroschedji 2008)

Kh. Qeiyafa (Garfinkel and Ganor 2009; Garfinkel, Ganor and Hasel 2014)

Tel Azekah (Lipschits, Gadot, and Oeming 2012: Lipschits and Gadot, this volume)

Tel Ḥarasim (Givon 2008) Settlement

V

Settlement

8

Tel Beth-Shemesh (Bunimovitz and Lederman 2006, 2008, 2009; Lederman and Bunimovitz 2014; this volume)

Ḥ. Nina (Yannai 2008; Golan 2013)

VIB

XV

Settlement?

Tel Batash (Panitz-Cohen and Mazar 2006)

Tel Malot (Shavit 1993; Ory and Shmueli 2006)

Tel Gezer (Dever, Lance and Wright 1970; Dever, et al. 1974; Dever 1986, 1993, 2003; Ortiz and Wolf 2012; this volume)

Tel Hamid (Shavit 2003; Wolff and Shavit 2008)

Tel Shaalavim (Parnos 2008)

LBA IIB

1200 BCE

6

XIII

V

1100 BCE

Settlement

Acr-5

Gap

Gap

4

IVB–A

Gap

IV

IV

III

3

Gap

Gap

VII

Gap

Late IA IIA

Settlement

IV

3

VIII

Early IA IIA

Settlement

Gap?

Settlement

II

2

III

Gap?

VII

VI

Settlement?

Iron Age IIB

1000 BCE    900 BCE    800 BCE

X   IX

Acr-3 Gap?

4

XI

Gap?

IA I

VB-A

5

Sherds

XII

Gap

Acr-4

VIIIA   VIIB   VIIA   VIB–A

7

VIA

XIV (= Gap?)

TBI

Table 1.  Stratigraphic Information (partial information is marked by a light-gray shade)

Appendix

Settlements and Interactions in the Shephelah 195

IA I

B2

B-5

VIII

B3

1

Settlement

V

B-4

IV

II

VII Settlement

Sherds

III

Gap

VIII

VIIB

VIIA

a.  Based on Late Iron Age IIA indicative types from Stratum B3 (Herzog and Singer-Avitz 2004: 221).

C2

Tel Beit-Mirsim (Greenberg 1987)

Tel Ḥalif (Seger 1983, 1993; Jacobs and Seger 2007)

A1 VIB

A-4, B-3, D-2

B-6

Gap

III

Settlement Areas A1, A2

Settlement

A2, F8-F7

Iron Age IIB

Settlement?

B1

VI

Gap

?

Late IA IIA

   A-4, D     A-3, D, E2, F9

Early IA IIA

1000 BCE    900 BCE    800 BCE

Tel Eton (Faust 2011; Faust, et al. 2014)

Gap?

1100 BCE

E-4a, Lower F-2   A-5, E-3, Lower F-1

TBI

Tel Maresha (Kloner 2003)

Kh. el-Qom (Holladay 1971)

Tel Ḥesi (Fargo 1993)

B-7

VII

Tel Lachish (Ussishkin 2004)

Kh. Umm el-Baqar (Nahshoni and Talis 2015)

IV

Area B

Tel Zayit (Tappy 2008; 2011; this volume)

Tel Burna (Shai, et al. 2012; Shai, McKinny and Uziel 2015; Shai, this volume)

Tel Goded (Bliss and Macalister 1902; Gibson 1994)

Tell eṣ-Ṣafi (Maeir 2012a; 2012b; this E-4b, Lower volume) F-3

LBA IIB

1200 BCE

Table 1.  Stratigraphic Information (partial information is marked by a light-gray shade)

196 Ido Koch

2

3

6

2

2

1

6

9

3

2

4

4

4

10

4

4

Hulda

Kefar Uriah

Shaʿar Ha-Gay

Beʾer Tuviya

Kefar Menahem

Beth-Shemesh

Nes Harim

Qomemiyut

Gath

Beth-Guvrin

Ruḥama

Lachish

Amatzia

Upper HaʾElah Basin and northwestern Hebron Hillsc

14

3

Late Bronze

Gedera

Survey

Beit Sira

Gezer

Rehovot

12

12

2

3

3

1

1

0

3

4

1

2

2

1

1

3

10

1

Iron Age I

Remarks

Two excavated sites (Kh. Tubeiqa and Hebron) and limited surveys (Ofer 1993; Dagan 2000)

One excavated site (Tel Eton) and an extensive survey (Dagan 2007)

48

One excavated site (Tel Lachish) and extensive survey (Dagan 1992)

130b

Two excavated sites (Tel Ḥesi and Tel Nagila) and a limited survey (Shavit 2003)

Two excavated sites (Tel Goded and Tel Maresha) and a limited survey (Dagan 2000)

Two excavated sites (Tel Burna and Tel Zayit) and limited surveys (Dagan 2000; Shavit 2003)

Limited surveys (Dagan 2000; Shavit 2003)

Extensive survey (Weiss, Zissu and Solimany 2004)

Five excavated sites (Tel Beth Shemesh, Tel Yarmuth, Kh. Qeiyafa, Tel Azekah, Tel Socoh) the eastern half underwent an extensive survey (Dagan 2010, 2011) and the western half, a limited survey (Dagan 2000)

Two excavated sites (Tell eṣ-Ṣafi and Tel Ḥarasim, and limited surveys (Dagan 2000; Shavit 2003)

Two surveys at specific points (Anonymous 1966; Shavit 2003)

Limited surveys (Dagan 2000; Shavit 2003)

Two excavated sites (Tel Batash and Beqoʿa) and limited surveys (Dagan 2000; Shavit 2003)

Two excavated sites (Tel Miqne and Ḥ. Nina) and limited surveys (Dagan 2000; Shavit 2003)

Comprehensive survey (Barda and Zbenovich 2012)

Comprehensive survey (Shavit 1992; Magen and Finkelstein 1993)

Three excavated sites (Tel Gezer, Tel Shaʿalabim, Tel Hamid) and comprehensive survey (Shavit 2013)

Salvage excavations (Ramle, Zarnuqa, Tel Malot) and partial survey (Paz, Lalkin and Danino 2014)a

29

6

6

9

3

50

76

21

2

3

3

4

12

7

32

7

Iron Age II

Table 2.  Survey Data

Settlements and Interactions in the Shephelah 197

3

6

Dvira

Upper Naḥal Adorayim and southwestern Hebron Hills

4

14

2 6 (33)

1

3

Iron Age II

? (33)d

1

Iron Age I

Remarks

Two excavated sites (Tell Beit Mirsim and Kh. Rabud) and a limited survey (Ofer 1993)

Extensive survey (Zissu, Ganor and Kheati 2015)

Surveys of specific points (Gophna 1963; Shavit 2003)

Survey (Huster 2015)

a.  The intensive survey encompassed about half the area of the survey map. b.  Dagan reported about 260 sites, but about half of those were located adjacent to other sites and therefore, for the sake of convenience, adjacent sites have been conflated in this table. c.  West of longitude 220000 and south of latitude 600000 near the watershed between the Lachish and Shiqma basins, about 2 km south of Hebron. d.  The surveyors note that these 33 sites could not be dated to a sub-period in the Iron Age.

1

Shuval

Late Bronze

3

Survey

Sderot

Table 2.  Survey Data

198 Ido Koch

Settlements and Interactions in the Shephelah

199

References Amiran, D. 1948 The Western Border of the Hebron Mountains. Bulletin of the Palestine Exploration Society 14: 112–18, 130. (Hebrew) Anonymous 1966 Shaphir. Hadashot Arkheologiyot 17: 4. (Hebrew) Barda, L., and Zbenovich, V. 2012 Archaeological Survey of Israel: Map of Gedera – 85. Jerusalem: Israel Antiquities Authority. Ben-Arieh, S. 1981 Tell Jedur. Eretz-Israel 15: 115–28. Ben-Dor Evian, S. 2012 Egypt and Philistia in the Iron Age I: The Case of the Philistine Lotus Flower. Tel Aviv 39: 20–37. Ben-Shlomo, D. 2006 Decorated Philistine Pottery: An Archaeological and Archaeometric Study. Oxford: Archaeopress. Ben-Shlomo, D., and Van Beek, G. 2014 The Smithsonian Institution Excavation at Tell Jemmeh, Israel, 1970–1990. Washington, D.C.: Smithsonian Institution Scholarly Press. Berman, A.; Barda, L.; and Stark, H. 2005 Archaeological Survey of Israel: Map of Ashdod – 84. Jerusalem: Israel Antiquities Authority. Bliss, F. J., and Macalister, R. A. S. 1902 Excavations in Palestine during the Years 1898–1900. London: Committee of the Palestine Exploration Fund. Bunimovitz, S. 1998 Sea Peoples in Cyprus and Israel: A Comparative Study of Immigration Processes. Pp. 103–13 in Mediterranean Peoples in Transition: Thirteenth to Early Tenth Centuries BCE, In Honor of Trude Dothan, ed. S. Gitin; A. Mazar; and E. Stern. Jerusalem: Israel Exploration Society. Bunimovitz, S., and Lederman, Z. 2006 The Early Israelite Monarchy in the Sorek Valley: Tel Beth-Shemesh and Tel Batash (Timnah) in the 10th and 9th Centuries BCE. Pp. 407–27 in “I will Speak the Riddles of Ancient Times”: Archaeological and Historical Studies in Honor of Amihai Mazar on the Occasion of His Sixtieth Birthday, ed. A. M. Maeir and P. de Miroschedji. Winona Lake, IN: Eisenbrauns. 2008 Beth-Shemesh. Pp. 1644–48 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. 2009 The Archaeology of Border Communities: Renewed Excavations at Tel Beth-​ Shemesh, Part 1: The Iron Age. Near Eastern Archaeology 72: 114–42. 2013 Solving a Century-Old Puzzle: New Discoveries at the Middle Bronze Gate of Tel Beth-Shemesh. Palestine Exploration Quarterly 145: 6–24. Byrne, R. 2007 The Refuge of Scribalism in Iron I Palestine. Bulletin of the American Schools of Oriental Research: 1–31.

200

Ido Koch

Casana, J. 2007 Structural Transformations in Settlement Systems of the Northern Levant. American Journal of Archaeology 111: 195–221. Cooke, F. T. 1923 The Site of Kirjath-Jearim. The Annual of the American Schools of Oriental Research 5: 105–20. Dagan, Y. 1992 Archaeological Survey of Israel: Map of Lakhish – 98. Jerusalem: Israel Antiquities Authority. 2000 The Settlement in the Judean Shephelah in the Second and First Millennium B.C.: A Test Case of Settlement Processes in a Geographical Region. Ph.D. dissertation, Tel Aviv University. (Hebrew with English Summary) 2007 Archaeological Survey of Israel: Map of Amaẓya – 109. Jerusalem: Israel Antiquities Authority. 2010 The Ramat Bet Shemesh Regional Project: The Gazetteer. Jerusalem: Israel Antiquities Authority. 2011 The Ramat Bet Shemesh Regional Project: Landscapes of Settlement, from the Paleolithic to the Ottoman Period. Jerusalem: Israel Antiquities Authority. Dan, Y. 1988 The Soils of the Southern Shefelah. Pp. 50–58 in Man and Environment in the Southern Shefelah: Studies in Regional Geography and History, ed. E. Stern and D. Urman. Givatayim: Masada. (Hebrew) Dever, W. G. 1986 Gezer IV: The 1969–71 Seasons in Field VI, the “Acropolis”. Jerusalem: Nelson Glueck School of Biblical Archaeology. 1993 Gezer. Pp. 496–506 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. 2003 Visiting the Real Gezer: A Reply to Israel Finkelstein. Tel Aviv 30: 259–82. Dever, W. G.; Lance, H. D.; Bullard, R. G., Cole, Dan P. ; Seger, J. D.; and Wright, G. E. 1974 Gezer II: Report of the 1967–70 Seasons in Fields I and II. Jerusalem: Nelson Glueck School of Biblical Archaeology. Dever, W. G.; Lance, H. D.; and Wright, G. E. 1970 Gezer I: Preliminary Report of the 1964–1966 Seasons. Jerusalem: Nelson Glueck School of Biblical Archaeology. Dothan, T. 1982 The Philistines and Their Material Culture. New Haven: Yale University Press. Dothan, T., and Gitin, S. 1993 Miqne, Tel Pp. 1051–59 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. 2008 Miqne, Tel (Ekron). Pp. 1952–58 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. Dothan, T.; Gitin, S.; and Zukerman, A. 2006 The Pottery: Canaanite and Philistine Traditions and Cypriote and Aegean Imports. Pp. 71–175 in Tel Miqne-Ekron Excavations 1995–1996: Field INE East Slope Iron Age I (Early Philistine Period), ed. M. W. Meehl; T. Dothan; and S. Gitin. Jerusalem: Albright Institute of Archaeological Research. Eisenberg, E., and Ben-Shlomo, D. 2016 Tel Hevron. Hadashot Arkheologiyot: Excavations and Surveys in Israel 128.

Settlements and Interactions in the Shephelah

201

Eisenberg, E., and Nagorski, A. 2002 Tel Ḥevron (Er-Rumeidi). Hadashot Arkheologiyot: Excavations and Surveys in Israel 114: 91*–92*. Fantalkin, A. 2008 The Appearance of Rock-Cut Bench Tombs in Iron Age Judah as a Reflection of State Formation. Pp. 17–44 in Bene Israel: Studies in the Archaeology of Israel and the Levant during the Bronze and Iron Ages in Honour of Israel Finkelstein, ed. A. Fantalkin and A. Yasur-Landau. Leiden and Boston: Brill. Fargo, V. M. 1993 Ḥesi, Tel. Pp.  630–34 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. Faust, A. 2011 The Excavations at Tel ʿEton (2006–2009): A Preliminary Report. Palestine Exploration Quarterly 143: 198–224. 2013a The Shephelah in the Iron Age: A New Look on the Settlement of Judah. Palestine Exploration Quarterly 145: 203–19. 2013b From Regional Power to Peaceful Neighbour: Philistia in the Iron I–II Transition. Israel Exploration Journal 63: 174–204. 2015 Pottery and Society in Iron Age Philistia: Feasting, Identity, Economy, and Gender. Bulletin of the American Schools of Oriental Research: 167–98. Faust, A., and Katz, H. 2011 Philistines, Israelites and Canaanites in the Southern Trough Valley during the Iron Age I. Egypt and the Levant 21: 231–47. Faust, A.; Katz, H.; Ben-Shlomo, D.; Sapir, Y.; and Eyall, P. 2014 Tēl ʿĒṭōn/Tell ʿĒṭūn and Its Interregional Contacts from the Late Bronze Age to the Persian-Hellenistic Period: Between Highlands and Lowlands. Zeitschrift Des Deutschen Palastina-Vereins 130: 43–76. Faust, A., and Weiss, E. 2005 Judah, Philistia, and the Mediterranean World: Reconstructing the Economic System of the Seventh Century BCE. Bulletin of the American Schools of Oriental Research 338: 71–92. Finkelstein, I. 1996 The Philistine Countryside. Israel Exploration Journal 46: 225–42. 2007 Is the Philistine Paradigm Still Viable? Pp. 517–23 in The Synchronization of Civilisations in the Eastern Mediterranean in the Second Millennium B.C. III, ed. M. Bietak and E. Czerńy. Vienna: Verlag der Osterreichischen Akademie der Wissenschaften. 2014 The Shephelah and Jerusalem’s Western Border in the Amarna Period. Egypt and the Levant 24: 267–76. Finkelstein, I.; Koch, I.; and Lipschits, O. 2011 The Mound on the Mount: A Possible Solution to the “Problem with Jerusalem”. Journal of Hebrew Scriptures 11: 2–24. Finkelstein, I., and Sass, B. 2013 The West Semitic Alphabetic Inscriptions, Late Bronze II to Iron IIA: Archeological Context, Distribution and Chronology. Hebrew Bible and Ancient Israel 2: 149–220. Finkelstein, I.; Sass, B.; and Singer-Avitz, L. 2008 Writing in Iron IIA Philistia in the Light of the Tel Zayit/Zeta Abecedary. Zeitschrift Des Deutschen Palastina-Vereins 124: 1–14. Finkelstein, I., and Singer-Avitz, L. 2009 Reevaluating Bethel. Zeitschrift des Deutschen Palastina-Vereins 125: 33–48.

202

Ido Koch

Funk, R. W. 1993 Beth-Zur. Pp.  259–61 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. Garfinkel, Y., and Ganor, S. 2009 Khirbet Qeiyafa Vol. 1: Excavation report 2007–2008 Seasons. Jerusalem: Israel Exploration Society. 2010 Khirbet Qeiyafa in Survey and in Excavations: A Response to Y. Dagan. Tel Aviv 37: 67–78. Garfinkel, Y.; Ganor, S.; and Hasel, M. G. 2014 Khirbet Qeiyafa Vol. 2: Excavation Report 2009–2013: Stratigraphy and Architecture (Areas B, C, D, E). Jerusalem: Israel Exploration Society. Gibson, S. 1994 The Tell ej-Judeideh (Tel Goded) Excavations: A Re-appraisal Based on Archival Records in the Palestine Exploration Fund. Tel Aviv 21: 194–234. Gilboa, A.; Cohen-Weinberger, A.; and Goren, Y. 2006 Philistine Bichrome Pottery: The View from the Northern Canaanite Coast. Pp. 303– 34 in “I will Speak the Riddles of Ancient Times”: Archaeological and Historical Studies in Honor of Amihai Mazar on the Occasion of His Sixtieth Birthday, ed. A. M. Maeir and P. de Miroschedji. Winona Lake, IN: Eisenbrauns. Gitin, S. 2006 The lmlk Jar-Form Redefined: A New Class of Iron Age II Oval Shaped Storage Jar. Pp. 505–24 in “I will Speak the Riddles of Ancient Times”: Archaeological and Historical Studies in Honor of Amihai Mazar on the Occasion of His Sixtieth Birthday, ed. A. M. Maeir and P. de Miroschedji. Winona Lake, IN: Eisenbrauns. Givon, S. 2008 Ḥarasim, Tel. Pp.  1766–67 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. Golan, D. 2013 Khirbat el Maghara/Khirbat Nina. Hadashot Arkheologiyot: Excavations and Surveys in Israel 125. Gophna, R. 1963 “Ḥāṣērim” Settlements in the Northern Negev. Bulletin of the Israel Exploration Society 23: 173–80. (Hebrew) Gophna, R., and Beit-Arieh, I. 1997 Archaeological Survey of Israel: Map of Lod – 80. Jerusalem: Israel Antiquities Authority. Greenberg, R. 1987 New Light on the Early Iron Age at Tell Beit Mirsim. Bulletin of the American Schools of Oriental Research 265: 55–80. Grossman, D. 1994 Expansion and Desertion: The Arab Village and its Offshoots in Ottoman Palestine. Jerusalem: Yad Ben Zvi. (Hebrew) Gunneweg, J.; Asaro, F.; Michel, H. V.; and Perlman, I. 1994 Interregional Contacts between Tell en‐Nasbeh and Littoral Philistine Centres in Canaan during Early Iron Age I. Archaeometry 36: 227–39. Gunneweg, J.; Dothan, T.; Perlman, I.; and Gitin, S. 1986 On the Origin of Pottery from Tel Miqne-Ekron. Bulletin of the American Schools of Oriental Research 264: 3–16. Herzog, Z., and Singer-Avitz, L. 2004 Redefining the Centre: The Emergence of State in Judah. Tel Aviv 31: 209–44.

Settlements and Interactions in the Shephelah

203

Hitchcock, L. A., and Maeir, A. M. 2013 Beyond Creolization and Hybridity: Entangled and Transcultural Identities in Philistia. Archaeological Review from Cambridge 28: 43–65. Holladay, J. S. 1971 Khirbet el-Qom. Israel Exploration Journal 21: 175–77. Huster, Y. 2015 Ashkelon 5: The Land behind Ashkelon. Winona Lake, IN: Eisenbrauns. Jacobs, P. F., and Seger, J. D. 2007 Glimpses of the Iron Age I at Tel Halif. Pp. 146–65 in “Up to the Gates of Ekron”: Essays on the Archaeology and History of the eastern Mediterranean in Honor of Seymour Gitin, ed. S. White Crawford. Jerusalem: Albright Institute of Archaeological research. Kaplan, J., and Ritter-Kaplan, H. 1993 Jaffa. Pp. 655–59 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. Killebrew, A. E. 2013 Early Philistine Pottery Technology at Tel Miqne-Ekron: Implications for the Late Bronze–Early Iron Age Transition in the Eastern Mediterranean. Pp. 77–129 in The Philistines and other “Sea Peoples” in Text and Archaeology, ed. A. E. Killebrew and G. Lehmann. Atlanta: SBL. Kisilevitz, S.; Eirikh-Rose, A.; Khalaily, H.; and Greenhut, Z. 2014 Moza, Tel Moza. Hadashot Arkheologiyot: Excavations and Surveys in Israel 126. Kloner, A. 2003 Maresha Excavations Final Report I: Subterranean Complexes 21, 44, 70. Jerusalem: Israel Antiquities Authority. Koch, I. 2014 Goose Keeping, Elite Emulation and Egyptianized Feasting at Late Bronze Lachish. Tel Aviv 41: 161–79. 2016 Revisiting the Fosse Temple at Tel Lachish. Journal of Ancient Near Eastern Religions 16. Kochavi, M. 1974 Khirbet Rabûd = Debir. Tel Aviv 1: 2–33. Kowalewski, S. A. 2008 Regional Settlement Pattern Studies. Journal of Archaeological Research 16: 225–85. Lederman, Z., and Bunimovitz, S. 2014 Canaanites, “Shephelites” and Those who will become Judahites. Pp. 61–71 in New Studies in the Archaeology of Jerusalem and Its Region, ed. G. Stiebel; O. Peleg-Barkat; D. Ben-Ami; and Y. Gadot. Jerusalem: Israel Antiquity Authority. (Hebrew with English summary) Lehmann, G., and Niemann, H. M. 2014 When Did the Shephelah Become Judahite? Tel Aviv 41: 77–94. Lev-Tov, J. S. E. 2010 A Plebeian Perspective on Empire Economies: Faunal Remains from Tel MiqneEkron, Israel. Pp. 90–104 in Anthropological Approaches to Zooarchaeology: Colonialism, Complexity and Animal Transformations, ed. D. V. Campana; A. Choyke; P. J. Crabtree; S. D. deFrance; and J. S. E. Lev-Tov. Oxford: Oxbow Books. 2012 A Preliminary Report on the Late Bronze and Iron Age Faunal Assemblages from Tell es-Safi/Gath. Pp. 589–612 in Tell es-Safi/Gath I: Report on the 1996–2005 Seasons, ed. A. M. Maeir. Wiesbaden: Harrassowitz.

204

Ido Koch

Lipschits, O.; Gadot, Y.; and Oeming, M. 2012 Tel Azekah 113 Years After: Preliminary Evaluation of the Renewed Excavations at the Site. Near Eastern Archaeology 75: 196–206. Lipschits, O.; Sergi, O.; and Koch, I. 2011 Judahite Stamped and Incised Jar Handles: A Tool for Studying the History of Late Monarchic Judah. Tel Aviv 38: 5–41. Maeir, A. M. 2008 Aegean Feasting and other Indo-European Elements in the Philistine Household. Pp.  347–52 in Dais—The Aegean Feast ed. L. A. Hitchcock; R. Laffineur; and J. L. Crowley. Liège and Austin, Texas: Université de Liège; University of Texas at Austin. 2011 The Archaeology of Early Jerusalem: From the Late Proto-historic Period (ca. 5th Millennium) to the End of the Bronze Age (ca. 1200 B.C.E.). Pp. 171–87 in Unearthing Jerusalem: 150 Years of Archaeological Excavations in the Holy City, ed. K. Galor and G. Avni. Winona Lake, IN: Eisenbrauns. 2012a Tell es-Safi/Gath 1: The 1996–2005 Seasons. Wiesbaden: Harrassowitz. 2012b Insights on the Philistine Culture and Related Issues: An Overview of 15 years of Work at Tell eṣ-Ṣafi/Gath. Pp.  345–404 in The Ancient Near East in the 12th–10th Centuries BCE Culture and History: Proceedings of the International Conference held at the University of Haifa, 2–5 May, 2010, ed. G. Galil; A. Gilboa; A. M. Maeir; and D. Kahn. Münster: Ugarit-Verlag. Maeir, A. M.; Hitchcock, L. A.; and Horwitz, L. K. 2013 On the Constitution and Transformation of Philistine Identity. Oxford Journal of Archaeology 32: 1–38. Maeir, A. M., and Uziel, J. 2007 A Tale of Two Tells: A Comparative Perspective on Tel Miqne-Ekron and Tell es-Safi/ Gath in Light of Recent Archaeological Research. Pp. 29–42 in “Up to the Gates of Ekron”: Essays on the Archaeology and History of the eastern Mediterranean in Honor of Seymour Gitin, ed. S. White Crawford. Jerusalem: Albright Institute of Archaeological research. Magen, Y., and Finkelstein, I. 1993 Archaeological Survey of the Hill Country of Benjamin. Jerusalem: Israel Antiquities Authority. Maher, E. 2005 The Faunal Remains. Pp.  283–90 in Ashdod VI: The Excavations of Areas H and K (1968–1969), ed. M. Dothan and D. Ben-Shlomo. Jerusalem: Israel Antiquities Authority. Martin, M. A. S. 2011 Egyptian-Type Pottery in the Late Bronze Age Southern Levant. Vienna: Verlag der Österreichischen Akademie der Wissenschaften. Mazar, A. 1985 Excavations at Tell Qasile, Part 2: The Philistine Sanctuary – Various Finds, the Pottery, Conclusions, Appendices. Jerusalem Institute of Archaeology, Hebrew University of Jerusalem. 2007 Myc IIIC in the Land of Israel: Its Distribution, Date and Significance. Pp. 571–82 in The Synchronization of Civilisations in the Eastern Mediterranean in the Second Millennium B.C. III, ed. M. Bietak and E. Czerńy. Vienna: Verlag der Osterreichischen Akademie der Wissenschaften.

Settlements and Interactions in the Shephelah

205

Mazow, L. B. 2005 Competing Material Culture: Philistine Settlement at Tel Miqne-Ekron in the Early Iron Age. Ph.D. Dissertation, University of Arizona. 2006–2007  The Industrious Sea Peoples: The Evidence of Aegean-style Textile Production in Cyprus and the Southern Levant. Scripta Mediterranea 27–28: 291–321. 2013 Throwing the Baby Out with the Bathwater: Innovations in Mediterranean Textile Production at the End of the 2nd/Beginning of the 1st Millennium BCE. Pp. 215–23 in Textile Production and Consumption in the Ancient Near East: Archaeology, Epigraphy, Iconography, ed. M.-L. Nosch; H. Koefoed; and E. Anderson Strand. Oxford and Oakville: Oxbow Books. 2014 Competing Material Culture: Philistine Settlement at Tel Miqne-Ekron in the Early Iron Age. Pp.  131–63 in Material Culture Matters: Essays on the Archaeology of the Southern Levant in Honor of Seymour Gitin, ed. J. R. Spencer; R. A. Mullins; and A. J. Brody. Winona Lake, IN: Eisenbrauns. Meehl, M. W.; Dothan, T.; and Gitin, S. 2006 Tel Miqne-Ekron Excavations 1995–1996: Field INE East Slope Iron Age I (Early Philistine Period). Jerusalem: W. F. Albright Institute of Archaeological Research and Institute of Archaeology, Hebrew University. Miroschedji, P. de 2008 Jarmuth. Pp. 1792–97 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. Morris, E. F. 2005 The Architecture of Imperialism: Military Bases and the Evolution of Foreign Policy in Egypt’s New Kingdom. Leiden and Boston: Brill. Naʾaman, N. 2010 Khirbet Qeiyafa in Context. Ugarit-Forschungen 42: 497–526. 2011 The Shephelah according to the Amarna Letters. Pp. 281–99 in The Fire Signals of Lachish, ed. I. Finkelstein and N. Naʾaman. Winona Lake, IN: Eisenbrauns. Nahshoni, P., and Talis, S. 2015 An Iron Age Site at Khirbat Umm el-Baqr (Naḥal Adorayim). ʿAtiqot 81: 69–105. (Hebrew with English Summary) Ofer, A. 1993 The Highland of Judah during the Biblical Period. Ph.D. dissertation, Tel Aviv University. (Hebrew with English Summary) Ortiz, S., and Wolf, S. R. 2012 Guarding the Border to Jerusalem: The Iron Age City of Gezer. Near Eastern Archaeology 75: 4–19. Ory, J., and Shmueli, O. 2006 Tel Malot. Hadashot Arkheologiyot: Excavations and Surveys in Israel 118. Panitz-Cohen, N., and Mazar, A. 2006 Timnah (Tel Batash) 3: The Finds from the Second Millennium BCE. Jerusalem: Institute of Archaeology, Hebrew University of Jerusalem. Parnos, G. 2008 Tel Sha‘alvim (East). Hadashot Arkheologiyot: Excavations and Surveys in Israel 120. Paz, Y.; Lalkin, N.; and Danino, D. 2014 Archaeological Survey of Israel: Map of Rehovot – 76. Israel Antiquities Authority. http:// www.antiquities.org.il/survey/new/; accessed June 10, 2016

206

Ido Koch

Pritchard, J. B. 1963 The Bronze Age Cemetery at Gibeon. Philadelphia: University Museum, University of Pennsylvania. Rainey, A. F. 2015 The El-Amarna Correspondence: A New Edition of the Cuneiform Letters from the Site of El-Amarna based on Collations of all Extant Tablets. Leiden and Boston: Brill. Sapir-Hen, L.; Gadot, Y.; and Finkelstein, I. 2014 Environmental and Historical Impacts on Long Term Animal Economy: The Southern Levant in the Late Bronze and Iron Ages. Journal of the Economic and Social History of the Orient 57: 703–44. Seger, J. D. 1983 Investigations at Tell Halif, Israel, 1976–1980. Bulletin of the American Schools of Oriental Research 252: 1–23. 1993 Ḥalif, Tel. Pp. 553–59 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. Sellers, O. R. 1968 The 1957 Excavations at Beth-Zur. Cambridge: American Schools of Oriental Research. Sergi, O. 2013 Judah’s Expansion in Historical Context. Tel Aviv 40: 226–46. 2015 The Emergence of Judah between Jerusalem and Benjamin. New Studies in the Archaeology of Jerusalem and its Region 9: 50–73. (Hebrew with English Summary) Sergi, O.; Karasik, A.; Gadot, Y.; and Lipschits, O. 2012 The Royal Judahite Storage Jar: A Computer-Generated Typology and Its Archaeological and Historical Implications. Tel Aviv 39: 64–92. Shai, I.; Cassuto, D. R.; Dagan, A.; and Uziel, J. 2012 The Fortifications at Tel Burna: Date, Function and Meaning. Israel Exploration Journal 62: 141–57. Shai, I., and Maeir, A. M. 2003 Pre-lmlk Jars: A New Class of Iron Age IIA Storage Jars. Tel Aviv 30: 108–23. Shai, I.; McKinny, C.; and Uziel, J. 2015 Late Bronze Age Cultic Activity in Ancient Canaan: A View from Tel Burna. Bulletin of the American Schools of Oriental Research 374: 115–33. Shavit, A. 1992 The Ayalon Valley and its Vicinity During the Bronze and Iron Ages Tel Aviv University. (Hebrew, with abstract) 1993 Tel Malot. Excavations and Surveys in Israel 12: 49–50. 2003 Settlement Patterns in Israel’s Southern Coastal Plain during the Iron Age II. Ph.D. dissertation, Tel Aviv University. (Hebrew with English Summary) 2008 Settlement Patterns of Philistine City-States. Pp. 135–64 in Bene Israel: Studies in the Archaeology of Israel and the Levant during the Bronze and Iron Ages in Honour of Israel Finkelstein, ed. A. Fantalkin and A. Yasur-Landau. Leiden: Brill. 2013 Archaeological Survey of Israel: Map of Gezer – 82. Israel Antiquities Authority. Stager, L. E. 1995 The Impact of the Sea Peoples in Canaan (1185–1050 BCE). Pp. 332–48 in The Archaeology of Society in the Holy Land, ed. T. E. Levy. London: Leicester University. Stern, E. 1988 The Southern Shefelah–Boundaries and Geographical Definition. Pp. 12–16 in Man and Environment in the Southern Shefelah: Studies in Regional Geography and History, ed. E. Stern and D. Urman. Givatayim: Masada. (Hebrew)

Settlements and Interactions in the Shephelah

207

Tappy, R. E. 2008 Zayit, Tel. Pp. 2082–83 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. 2011 The Depositional History of Iron Age Tel Zayit: A Response to Finkelstein, Sass, and Singer-Avitz. Eretz-Israel 30: 127*–43*. Ussishkin, D. 2004 The Renewed Archaeological Excavations at Tel Lachish (1973–1994). Tel Aviv: Institute of Archaeology of Tel Aviv University. 2005 The Fortifications of Philistine Ekron. Israel Exploration Journal 55: 35–65. Uziel, J. 2008 The Southern Coastal Plain of Canaan during the Middle Bronze Age. Ph.D. dissertation, Bar Ilan University. Uziel, J.; Shai, I.; and Cassuto, D. R. 2014 Ups and Downs of Settlement Patterns: Why Sites Fluctuate. Pp. 295–308 in Material Culture Matters: Essays on the Archaeology of the Southern Levant in Honor of Seymour Gitin, ed. J. R. Spencer; R. A. Mullins; and A. J. Brody. Winona Lake, IN: Eisenbrauns. Weiss, D.; Zissu, B.; and Solimany, G. 2004 Archaeological Survey of Israel: Map of Nes Harim – 104. Jerusalem: Israel Antiquities Authority. Wolff, S. R., and Shavit, A. 2008 Ḥamid, Tel. Pp. 1762–63 in New Encyclopedia of Archaeological Excavations in the Holy Land, ed. E. Stern. Jerusalem: Israel Exploration Society. Yannai, E. 2008 Khirbat Nina. Hadashot Arkheologiyot: Excavations and Surveys in Israel 120. Yasur-Landau, A. 2007 Let’s Do the Time Warp Again: Migration Processes and the Absolute Chronology of the Philistine Settlement. Pp. 609–20 in The Synchronization of Civilisations in the Eastern Mediterranean in the Second Millennium V.B. III, ed. M. Bietak and E. Czerńy. Vienna: Verlag der Österreichischen Akademie der Wissenschaften. Yellin, J., and Gunneweg, J. 1985 Provenience of Pottery from Tell Qasile Strata VII, X, XI and XII. Pp.  111–17 in Excavations at Tell Qasile, Part 2: The Philistine Sanctuary – Various Finds, the Pottery, Conclusions, Appendices, ed. A. Mazar. Jerusalem: Institute of Archaeology, Hebrew University of Jerusalem. Yoffee, N. 2005 Myths of the Archaic State: Evolution of the earliest Cities, States, and Civilizations. Cambridge: Cambridge University Press. Zissu, B.; Ganor, S.; and Kheati, R. 2015 Archaeological Survey of Israel: Map of Devira – 120. http://www.antiquities.org.il/survey/new/ ; accessed June 10, 2016