133 34 172MB
English Pages 848 [846] Year 2014
K O M M O S An Excavation on the South Coast of Crete by the University of Toronto and the Royal Ontario under the auspices of the American School of Classical Studies at Athens Joseph W . Shaw Maria C. Shaw EDITORS
Volume I, Part 1
Museum
LIBYAN
SEA
'lAIN
234 9 10 o_ __===-~~~~5~~6~l7==!8 __!-_ 1988
km.
KOMMOS
I
The Kommos Region and Houses of the Minoan Town E D I T E D M A R I A WITH
BY C.
J O S E P H
W.
S H A W
A N D
S H A W
C O N T R I B U T I O N S
BY
John Bennet, Philip P. Betancourt, Harriet Blitzer, Giuliana Bianco, Peter J. Callaghan, Mary K. Dabney, Katherine A. Frego, John A. Gifford, Deborah K. Harlan, John W. Hayes, John McEnroe, Lucia Nixon, Michael Parsons, Sebastian Payne, David S. Reese, Mark J. Rose, Katherine A. Schwab, Catherine Sease, Joseph W. Shaw, Maria C. Shaw, Jennifer M. Shay, C. Thomas Shay, Richard Hope Simpson, L. Vance Watrous, Helene Whittaker, James C. Wright, Janusz Zwiazek, and others
Parti The Kommos Region, Ecology, and Minoan Industries
Princeton University Press
:
P R I N C E T O N ,
N E W
J E R S E Y
Copynght © 1995 by Pnnceton University Press PublIshed by Pnnceton University Press, 41 WillIam Street, Pnnceton, New Jersey 08540 In the United Kingdom. Pnnceton University Press, Chichester, West Sussex All Rights Reserved ThiS book has been composed In Llnotron Palatlno and Aldus by The ComposIng Room of Michigan, Inc. Pnnceton University Press books are pnnted on aCid-free paper and meet the gUidelInes for permanence and durability of the Committee on ProductIOn GUidelInes for Book Longevity of the CounCil on Library Resources Pnnted In the United States of Amenca 10 9 8 7 6 5 4 3 2 1 Library of Congress Cataloging-In-PublicatIOn Data (ReVised for volume 1, pt. 1) Kommos. Includes bibliographICal references and Indexes Contents: v. 1. The Kommos regIOn and houses of the Minoan town pt. 1. The Kommos regIOn, ecology, and MInoan Industnes - v. 2. The fInal neolIthIC through Imddle MInoan III pottery / Philip. Betancourt - v 3 The late bronze age pottery / LIVingston Vance Watrous. 1. Kommos Site (Greece) 2. Mmoans. I. Shaw, Joseph W II. Shaw, Mana C. III. Betancourt, Philip P., 1936IV. Watrous, LIVIngston Vance, 1943. V University of Toronto. VI. Royal Ontano Museum. VII. Amencan School of ClaSSical Studies at Athens. DF221.C8K66 1990 939'.18 89-10817 ISBN 0-691-03334-X (v. 1, pt 1) ISBN 0-691-03594-6 (v 2 : alk. paper) ISBN 0-691-03607 (v. 3)
TO
THE
MEMORY
Sir Arthur Evans (1851-1941) Discoverer of the Kommos Site A N D
James Walter Graham (1906-1991) A N D
Nicholas Platon (1909-1992) Outstanding Interpreters of Minoan Culture
OF
Contents Preface by Joseph W. Shaw and Maria C. Shaw Abbreviations List of Plates List of Tables
C H A P T E R
xi xvii xix xxxv
1
The Topography and Archaeological Exploration of the Western Mesara by Joseph W. Shaw
C H A P T E R
2
The Exploration and Excavation of the Kommos Site by Joseph W. Shaw
C H A P T E R
3
The Physical Geology of the Western Mesara and Kommos by John A. Gijford
C H A P T E R
91
5
The Minoan Fauna by David S. Reese, with contributions by Mark J. Rose and Sebastian Payne
C H A P T E R
30
4
The Modern Flora and Plant Remains from Bronze Age Deposits at Kommos by C. Thomas Shay and Jennifer M. Shay, with !Catherine A. Frego and Janusz Zwiazek
C H A P T E R
8
163
6
Soil and Land Use Studies at Kommos by Michael Parsons, with John A. Gijford
292 ix
Contents
X
C H A P T E R
7
The Archaeological Survey of the Kommos Area by Richard Hope Simpson, with Philip P. Betancourt, Peter J. Callaghan, Deborah K. Harlan, John W. Hayes, Joseph W. Shaw, Maria C. Shaw, and L. Vance Watrous C H A P T E R
8
Minoan Implements and Industries by Harriet Blitzer References Index Plates
325
403 537 561 571
Preface This publication is the third to appear in a series of volumes being published by Princeton University Press on the results of the excavations from 1976 through 1985 at the Minoan and Greek site of Kommos in southern Crete. Unlike the two volumes already published (on the Middle and Late Minoan pottery), written by single authors (Betancourt 1990 [Vol. II] and Watrous 1992 [Vol. Ill]), this installment combines chapters by more than twenty-seven individuals contributing their special expertise to an understanding of both the physical and the archaeological character of this immensely rich and complex ancient site as a whole. This volume appears as Volume I in the series, however, because of its introductory nature. Because of considerations of size and to some extent of theme, the volume has been divided into two parts, with this preface serving as the main introduction to both. Naturally, the two parts are designed to be viewed as a single unit. Volume I, Part 1, offers a general introduction to the site of Kommos with chapters on the history and character of its excavation seen within the context of earlier archaeological exploration in the Mesara Plain and specifically in the Kommos area (Joseph W. Shaw), studies on the geomorphology (John A. Gifford), the flora (C. Thomas and Jennifer M. Shay) and the fauna (David S. Reese and others) of the Kommos region, and ancient and modern land use (Michael Parsons, with John A. Gifford), as well as a catalogue and analysis of finds from a foot survey in the Kommos area led by Richard Hope Simpson. A final chapter on Minoan implements and industries (Harriet Blitzer) provides a transition to Part 2 of this volume as it examines the availability of local resources as well as artifacts derived from them. An appendix to Chapter 2, Part 1, provides an entry from the diary of Sir Arthur Evans, the archaeologist who above all others alerted scholars to the archaeological potential of Kommos. Part 2 of Volume I concentrates on the town of the Minoan settlement at Kommos, leaving what we consider to be the Civic Center in the Southern Area of the site to be examined in a separate volume devoted to that area, where further excavation is now under way (projected Vol. V in the series). After an introductory chapter to the town as a whole (Joseph W. Shaw), which in part synthesizes the material in Kommos I, 1, a large portion of Part 2 is devoted to the publication of the houses, organized in two groups by area: the Hilltop (Chapter 2, by Maria C. Shaw and Lucia Nixon) and the Central Hillside (Chapter 3, by James C. Wright and John McEnroe), followed by Chapter 4, reporting on various categories of finds from the Xl
XIl
Preface
town. Next there is a section on the town's arrangement and the characteristics of its domestic architecture (Chapter 5, by Maria C. Shaw), followed by one dealing with aspects of the inhabitants' domestic economy and the historical development of the town (Chapter 6, by Joseph W. Shaw). Included are appendices on the topographical grid and permanent survey points (J. W. Shaw and Giuliana Bianco) and object conservation practices (Catherine Sease). Throughout Volume I, Part 2, are final plans and sections made by our excavation architect, Giuliana Bianco. Their preparation for publication, taking almost a year, was done in consultation with Maria C. Shaw. Prior to this publication, a series of extensive preliminary reports on the excavations (nine in all) appeared in Hesperia (J. W. Shaw 1977b, 1978a, 1979a, 1980a, 1981a, 1982a, 1984a, 1986; J. W. and M. C. Shaw 1993). Although now in part superceded by the final publication in terms of certain theories or details, these reports proved most useful, even serving in part as a foundation for chapters in this volume. Thanks are due to the American School of Classical Studies at Athens for making its prestigious periodical available to us for reporting yearly discoveries from Kommos and to Marian H. McAllister, Hesperia's editor, in particular, for her excellent editorial and other assistance. As with any complex project, it was during the formative part of the excavation and when the odds seemed against its getting under way, that help from and cooperation on the part of many proved so crucial. Foremost of these was the then Ephor in Herakleion, Stylianos Alexiou (Pl. 2.27), who enthusiastically supported the excavation. He set a pattern to be followed over the years by subsequent Ephors, namely Manolis Bourboudakis, Angeliki Lembessi, Yannis Sakellarakis, and presently Charalambos Kritsas. To them and to their respective assistants we are deeply indebted. To be singled out are the archaeologists serving on different occasions as official observers: Antonis Vasilakis, George Rethymniotakis, Alexandra Karetsou, and Despina Vallianou. Further support and assistance was provided by Nikolaos Kondoleon, Dimitrios Lazarides, and Ioannis Tzedakis, former Director of Antiquities. Essential to the work undertaken was, unquestionably, the generosity of the American School of Classical Studies at Athens, in granting us one of its own few and precious permits for excavation. Conditional to the results of the first excavation season in 1976, the permit was subsequently extended, allowing for the uncovering of important remains of both the Greek and the Minoan periods over a large area. To be thanked especially, both for the role played in the granting of the permit and help and advice over the years regarding practical and other aspects of the excavation, are James McCredie, then Director of the American School, who was also Chairman of the Excavation Permits Committee; Homer Thompson, for recommending the excavation; C. W. J. Eliot, Assistant Director of the School, for useful advice; Steven Miller, Director of the School, who gave practical and moral support during some politically turbulent years; Henry Immerwahr, Director of the School; and the present Director, William Coulson. Legal matters, such as land acquisition, were efficiently handled by Harry Bikakis, the School lawyer. Nikolaos Kapetanakos, then working in the Topo-
Preface
XlU
graphic Office of the District of Herakleion, efficiently handled the legal demarcation of public land at Kommos, to ready matters for land expropriation. With regard to permits, it must be noted here that the question of an excavation by us would not even have been possible had it not been for the generosity and backing of the Italian School of Archaeology, which had been, up to the time we began, the only foreign school active in excavation in the Mesara area. For this we are deeply indebted to the late Doro Levi, then Director of the Italian School, as well as to Clelia Laviosa, an active member of that School. Indeed, subsequent working relations with other members of the School, such as the present director, Antonino Di Vita, and, in particular, our colleague Vincenzo La Rosa, Director of ongoing excavations at Aghia Triada, have been a source of cordial and scholarly exchange and cooperation. Funding for the excavation was provided primarily by a Canadian governmental agency, by American and Canadian corporations, by an institute, and by private individuals. Chief among these supporters over the years has been the Canada Council (later the Social Sciences and Humanities Research Council of Canada), which generously supported our efforts through grants: S74-0460; S74-1930; S76-1232-X1,2; 410-77-0565-Xl,2; 410-78-0590-Xl,2,3; 82-0042-Xl,2,3; 410-85-0379-Xl,2,3; and 411-88-0020-Xl,2,3,4,5,6. Council administrators to be thanked for their support for the project are Frank Milligan, Mireille Badour, Marion King, Patrick Mates, Peter Carruthers, and Janet Lautenschlager. Among corporate sponsors, Paul Elicker of the SCM Corporation, later taken over by Hanson Trust, maintained his company's support from 1976 through 1985 and always showed a keen interest in the aims and results of the excavation. During our first season, Olivetti Canada, Ltd., and Staedtler-Mars, Ltd., provided office and drawing equipment, respectively; Keuffel and Esser of Canada supplied drawing materials from 1976 through 1983; Polaroid Corporation of America gave us both camera and film (1976-1980); and Eastman Kodak Canada, Ltd., supplied us with film during all our earlier study and excavation seasons (1976-1989). With pleasure we acknowledge individual sponsors: Luther Replogle (1975); Louise Stone (1976); Rue Shaw (1976); Leon Pomerance, in cooperation with the American Institute of Nautical Archaeology (1976-1977, 1981); and Lome Wickerson (1992, 1993). There was also a particularly generous donor from Tasmania (1982-1984) who wishes to remain anonymous. During 1987, 1990, 1991, and 1993, the Institute for Aegean Prehistory (INSTAP), established by Malcolm Wiener, provided funds for additional land purchase and certain publication and excavation costs. Of the two Canadian sponsoring institutions, the Royal Ontario Museum consistently provided funding for the project. Warm appreciation is due to A. Douglas Tushingham, then Chief Archaeologist at the ROM, whose original backing of the project provided us with our first sure sponsorship and who generously arranged for a seminar session to inform the project before it began. Then and after his retirement, the Curators of the Greek and Roman
XlV
Preface
Department, Neda Leipen, John W. Hayes, and Alison H. Easson, made possible the continuation of the sponsorship. The second sponsoring institution, the University of Toronto, was also active, especially through the offices of the President of the University, John Evans, in the first years of our work, then later under James Ham, George Connell, and now James Prichard. James Keffer, Vice-President—Research, has been particularly helpful. The Department of Fine Art of the University of Toronto has also been instrumental in the forwarding of the project from as early as 1974 when Frederick E. Winter, then Chairman of the Department, accompanied me to Ottawa to make the first presentation to the Canada Council. Subsequently, during 1986 the Department of Fine Art, then under the direction of Robert Welsh, made a room available for the Kommos archives and computer; and during the 1987-1989 period, when Joseph W. Shaw was Chairman of that same Department, the Dean of the Faculty of Arts and Sciences, Robin Armstrong, provided a special grant for additional secretarial help. As part of the ongoing development of archives for Kommos, a database program was created by Janet Tenody. Maria C. Shaw has operated the system since with the help of Helene Whittaker and others. For its consistent interest in the Kommos publication series and the care shown to us by its staff, especially by Joanna Hitchcock, first as Executive Editor for the Humanities, and later by Elizabeth Powers, we are indebted to the Princeton University Press, which has greatly alleviated our burden with experienced, professional help over the years. Editing help was provided also by Eric Csapo, of the Department of Classics in the University of Toronto, and Mary K. Dabney, who undertook for the Press the editing of the Kommos Volumes III and I, part 1; Barbara Ibronyi has provided direction and efficiency in the preparation of Volume I with the help of Dawn Cain. Student assistants provided immense help. Of these should be singled out Rebecca Duclos, who helped with the setting up of the illustrations; Susan Downie; and Jacke Phillips, among others. Secretarial help was also furnished through the University of Toronto, and, in particular in the Department of Fine Art, by Nana Brundage, Lorna Thurston, and Use Wister. The backbone of any excavation is made up of its staff. Joseph W. Shaw has been the director from the beginning as well as the field photographer. In 1982 Maria C. Shaw, who had helped with the project from the days of its inception and in particular with the original expropriation arrangements and numerous aspects of the field work, was named Assistant Director. Of the many people who worked in various capacities in connection with the excavation not everyone can be mentioned here, but a table is provided in Volume I, Part 1 (Table 2.1), which details names of participants, the time of their participation, and their general responsibilities. Among the staff in the field, a special role was held by the trenchmasters, usually advanced graduate students and professors, in whose intelligent hands rested the responsibility of recording and untangling stratigraphy and preparing detailed reports. The bulkiest material recovered from the site was, as at so many excavations, the pottery. Over the years
Preface
xv
we were fortunate to have ceramic experts who remained members of the team throughout the main phase of the excavation. In this respect, Philip P. Betancourt, Peter J. Callaghan, Alan Johnston, and L. Vance Watrous should be singled out for special mention. John W. Hayes, who was at Kommos for shorter times, helped establish storeroom procedure, in addition to studying the Roman pottery. Aside from pottery, two other people helped with identifying artifacts and other materials: David S. Reese, analyzing the ancient faunal remains, and Harriet Blitzer, analyzing the stone tools and objects connected with Minoan industries. Occasionally there were visiting scholars aiding us with their particular expertise: Daniel Geagan, who examined our Greek inscriptions; Robert Koehl, who studied the rhyta and was knowledgable about Cypriot and other Eastern imports; and Peter Warren, who joined us the first season so that we might profit from his knowledge and excavation experience. There were also visits by Patricia Bikai, J. Nicolas Coldstream, J. A. McGillivray, Eliezer Oren, Mervyn Popham, Jeremy B. Rutter, and Hugh Sackett, all offering advice and help in their areas of expertise. Oren, in particular, confirmed the earlier suspected presence of Egyptian sherds among the Bronze Age pottery. Basic to the recording and other logistical matters was, of course, the cataloguing department, set in the excavation headquarters in the village of Pitsidia. The Chief Cataloguers Elizabeth Comstock, Mary K. Dabney, Katherine A. Schwab, Deborah K. Harlan, and Niki Kantzios—pleasant and efficient in a difficult task—should be singled out for special mention and thanks. Cataloguing and secreterial help was also provided by a number of student and other assistants in the Department of Fine Art at the University of Toronto (Table 2.1). We are indebted to our technical staff, among whom we would single out Giuliana Bianco, talented and tireless excavation architect and one of our artists during the entire excavation period; the skilled conservators Catherine Sease and Barbara Hamann; and a number of talented photographers responsible for artifacts (pottery and small finds) with occasional forays into the field to help the Director: R. K. ("Chip") Vincent, Jr., Taylor Dabney, Timothy De Vinney, Winn Burke, Alexander C. Shaw, and Robin A. Shaw. In Athens during the early seasons, the staff of the photographic office of D. Harissiades, especially Stergios Svarnas, helped with photographic work. Throughout the excavation period John Glover and Louisa Yick of the Photographic Facility of the Faculty of Arts and Sciences of the University of Toronto made slides and printed photographs in Toronto. Among the many artists, Joseph Clarke was present for many years. We thank him and the other artists for their fine drawings. During the early seasons, and in order to understand the cultural context of the Kommos site better, it became necessary to organize a survey of the local area, conducted by Richard Hope Simpson. Besides a number of students, who often doubled also as excavators, there were other members of the survey team: Michael Parsons, investigating land use; John A. Gifford, studying the geology; and Jennifer M. Shay and C. Thomas Shay, researching botanical history.
XVl
Preface
For all of our seasons, the late George Beladakis of Pitsidia (Pl. 2.29) was excavation foreman as well as part-time site guard. Patient, responsible, intelligent, and wise, our friend George was for us the perfect foreman. His opinion was valued by all and his calm persuasion molded cooperative and hardworking groups of site workmen, most of whom were from Pitsidia. I would mention those loyal ones who excavated with us every summer: Aristotle Fasoulakis, Iannis Fasoulakis, Siphis Fasoulakis, Manolis Kalogrydakis, Andreas Kontoudakis, and Nikos Spinthakis—workmen for a decade. We should also recall George Manousoudakis, who during the same period carefully and skillfully operated the front-loader that removed hundreds of thousands of cubic meters of sand as well as the already excavated earth from the site. Basic to all our efforts was also the splendid fare supplied by our cooks, among whom we would mention with thanks Maria Fasoulakis, Harekleia Kypraki, Kaddiani Kypraki, Ero Kypraki, Anastasia Spinthaki, and Athanasoula Tsavolaki. Indeed, help provided by various local individuals, often offered without expected reward, made all the difference in the day-to-day running of a complex project and in other more crucial matters. Of the multitude of names only a few can be mentioned. Of these, we wish to thank Zacharias Spyridakis, former Chief Guard of the Phaistos area, who first in 1964 showed Joseph W. Shaw how to get to the Kommos site and who assisted on several occasions, for instance, in setting in borders for the property or furnishing the topographers with material; Petros Kyprakis, whose advice and encouragement led to the establishment of our base in Pitsidia; Nikos Markakis, who advised us on appropriate stores in Herakleion for the purchase of equipment; Ioannis Sphakakis and his family; Manolis and Theoniphi Kadianaki, landlords and dear friends; Iakovos and Maria Kadianaki, who rented to us the house which became the excavation's storerooms and headquarters; and the late Petros, and Fofo, Spinthaki, also friendly and dependable landlords. During the entire period one of our chief suppliers of writing and drawing materials and xerox copies of documents was Phanourios Chrystodoulakis, a progressive and helpful merchant in Mires. The welcome, enthusiasm, and hospitality of the residents of the general Pitsidia-Matala region, despite the dislocations caused by expropriation and non-commercial aspects of our enterprise, have often made our work and stay there a pleasure. Indeed, in some ways, the excavation work has become a regional effort on the part of some of the inhabitants who are, after all, the closest heirs to Kommos itself. To all of these, and to many others, we are forever grateful. Joseph W. Shaw Maria C. Shaw Department of Fine Art University of Toronto
24 November 1993
Abbreviations A A B BM(NH) Bo Bo C C ca. CH cm CS d dim EB EH EM EN ER est F FN G G
g gr.b. GS h H ha HP HST HT I kg km L LB LH LM LN
Archaic (Chap. T) silver bronze or copper British Museum, Natural History bone or ivory utilitarian bone (Chap. 8) Classical (Chap. 7) clay or terra cotta (Chap. 8) circa Central Hillside centimeter chipped stone diameter dimension Early Bronze Age Early Helladic Early Minoan Early Neolithic Early Roman estimated faience Final Neolithic Geometric (Chap 7) gold gram greatest breadth ground stone height Hellenistic hectare House with the Press House with the Snake Tube Hilltop inscriptions kilogram kilometer lead Late Bronze Age Late Helladic Late Minoan Late Neolithic
LR m M M M max MB Med MH Mi min MinPDim mm MM MNI MPD MPDep MPDim MPH MPL MPTh MPW NC NH NISP O
o/c OH P RB SA SC S.D. Sh th Ven W
Wt
Late Roman meter metalworking remains (Chap. 8) Minoan (Chap. 7) molar (Chap. 5) maximum Middle Bronze Age Medieval Middle Helladic miscellany (objects) minimum minimum preserved dimension millimeter Middle Minoan minimum number of individuals maximum preserved diameter maximum preserved depth maximum preserved dimension maximum preserved height maximum preserved length maximum preserved thickness maximum preserved width Northern Cliffside North House number of identified specimens organic (objects) Ovis/Capra Oblique House plaster Round Building Southern Area Southern Cliffside standard deviation shell thickness Venetian width weight
All measurements are given in centimeters unless otherwise specified. XVIl
List of Plates Frontispiece The Mesara and adjacent areas to the south (by Thomas Boyd, see Chap. 1, n. 1). Chapter 1 Plate 1.1. View by Francesco Basilicata, 1615, of the western valley of the Mesara, from Plate 87 of the Basilicata Manuscript "Citta Fortezze Castelli Siti Fort Spiaggie Porti E Redoti Del Regno Di Candia." Photograph by Harriet Blitzer through generous study permission of Mrs. Kastrinogiannakis, Acting Director, Historical Museum, Herakleion, Crete. Plate 1.2. View by Francesco Basilicata, 1615, of the Matala Valley, from Plate 85 of the Basilicata Manuscript "Citta Fortezze Castelli Siti Fort Spiaggie Porti E Redoti Del Regno Di Candia." Photograph by Harriet Blitzer through generous study permission of Mrs. Kastrinogiannakis, Acting Director, Historical Museum, Herakleion, Crete. Plate 1.3. Portion of an early map of the western Mesara (courtesy of the Herakleion Historical Museum, Map 1117). Chapter 2 Plate 2.1. Portion of diary page by Sir Arthur Evans (1924, courtesy of the Ashmolean Museum). Plate 2.2. The Kommos area in 1924 (by Piet de Jong, after A. Evans 1928: fig. 42).
Plate 2.5. Kommos, the shoreline and site, before excavation, from the south (Shaw). Plate 2.6. A Middle Minoan wall west of the North House, before excavation (1975). Plate 2.7. A Second World War German antivehicle mine, general view (left) and detail of the detonator (right) (T. Dabney). Plate 2.8. Kommos, the northern hill and slopes as seen from the south after excavation has progressed (1980). Plate 2.9. Balloon photograph of Kommos before excavation, north on right (W. and E. Myers, 1976). Plate 2.10. As Plate 2.9, detail of the northern hill, north on left (W. and E. Myers, 1976). Plate 2.11. Balloon view of Kommos after excavation, north at upper left (W. and E. Myers, 1984). Plate 2.12. General site plan by Giuliana Bianco (1985), north at top. Plate 2.13. Plan of the Hilltop Houses by Giuliana Bianco (1988). Plate 2.14. Trench plan of the Hilltop Houses by Giuliana Bianco (1989). Plate 2.15. Plan of the Central Hillside Houses by Giuliana Bianco (1988). Plate 2.16. Trench plan of the Central Hillside by Giuliana Bianco (1989). Plate 2.17. General period plan of the Southern Area by Giuliana Bianco (1992).
Plate 2.3. Kommos, the well and the "sherd slope" north of the well before excavation and the building of modern retaining walls (1975).
Plate 2.18. Trench plan of the Southern Area, upper level, by Giuliana Bianco (1989).
Plate 2.4. Kommos, the northern hill and slopes as seen from the south before excavation (1975).
Plate 2.19. As Plate 2.18, lower level, by Giuliana Bianco (1989). XlX
XX
Plate 2.20. The blessing of the site in July 1976. Plate 2.21. Exploratory Trench 2A, as staked out on the first morning of excavation, with Lucia Nixon (left) and Maria C. Shaw (right). Plate 2.22. The Southern Area during sand removal in 1977. Plate 2.23. The staff during the second season (1977). Plate 2.24. The main storeroom in Pitsidia, before improvements, with Maria C. Shaw, Alexander C. Shaw, and Robin A. Shaw (1975). Plate 2.25. The Pitsidia storeroom, after improvements (ca. 1980). Plate 2.26. A seminar in Pitsidia on a Saturday. Plate 2.27. Left to right, Peter Warren, Joseph W. Shaw, Maria C. Shaw, M. Aposkitou Alexiou, Ephor Stylianos Alexiou (1976). Plate 2.28. The shore house of George Sphakakis, used for storage and recreation. Plate 2.29. The Director, Joseph W. Shaw, and Foreman George Beladakis (ca. 1980, R. Vincent). Chapter 3 Plate 3.1. Simplified rendition of Geologic Map of Greece, 1:50,000, Tymbakion Sheet, southwest quadrant (Bonneau et al. 1984). Plate 3.2. Map of the drainage basins of the Kalamaki, Pitsidia, and Matala streams, plus distribution of the recent sand cover in the survey area. Plate 3.3. Matala, slope on Miocene marls of the Ambelouzos Formation, from Matala Bay southeast to Matalokephala. No vertical exaggeration. Plate 3.4. Kalamaki valley sediments. Cumulative grain-size curves for three samples from survey sites 56 and 20. Plate 3.5. Portion of LANDSAT 1 (Band 5) photo of Kommos region (scene identification 8106208271500), showing the very abrupt change in coastal morphology at the Matala headland, where there is a change from a low-angle unconsolidated shore zone to a rocky shore, which is fault-controlled.
List of Plates Plate 3.6. Scarp in the Kalamaki alluvial plain at site 56, showing the construction of a Hellenistic/Roman floor on muddy colluvium from nearby hill and then its burial by sandy colluvium (June 1975, J. A. Gifford). Plate 3.7. Vigles commercial sandpit, looking out to sea, showing an exposure of red colluvium with soil developed at surface under the sand cover (August 1978, J. A. Gifford). Plate 3.8. View of middle reach of Matala stream channel (north branch) near survey site 130, showing location of Middle Minoan pithos fragments approximately at a depth of 1.3 m in alluvial scarp (June 1979, J. A. Gifford). Plate 3.9. In the Matala Valley at the same locality as Plate 3.8, a close-up of pithos fragments in situ, showing one example of six large, unworn Middle Minoan pithos fragments included in cobbly alluvium (June 1979, J. A. Gifford). Plate 3.10. View, looking south towards Matala, of sand deposit under commercial excavation, south of Pitsidia ridge. The depth of sand here is 8 to 9 m and the deposit controls modern topography. The discontinuity at 3 to 4 m below the modern surface is indicated (June 1979, J. A. Gifford). Plate 3.11. Vertical air (balloon) photograph of the excavation site and the nearshore sandcovered bottom. Submerged in 1 to 2 m of water just seaward of the beach is a band of beachrock and, at the extreme edge of the photo, the eastern edge of the Papadoplaka, submerged in 6 m of water. North on top (W. and E. Myers, 1976). Plate 3.12. Top: Pre-excavation surface morphology of Kommos as seen from the southwest at an angle of 15°, based on ca. 1,300 data points from the 1976 Bandekas survey. Bottom: Estimate of bedrock surface morphology below sand, colluvium, and cultural strata, as seen from the southwest at an angle of 15°, based on ca. 250 data points from trench soundings and outcrops. Plate 3.13. Cumulative grain-size plot of Kommos kouskouras and lepidha samples. Plate 3.14. Cumulative grain-size plot of sand deposits at Kommos. Plate 3.15. Cumulative grain-size plot of Kom-
List of Plates
XXl
mos paleosols, showing correlation among Kommos site sediments, soils from survey site 75 (middle of B horizon), and the Pitsidia stream soil exposure (1 m deep).
Plate 4.7. Roadside with spring flowers between an olive orchard and a steep bank covered with shrubs and grasses. Note scattered cypress trees in the distance.
Plate 3.16. Cumulative grain-size plot of Kommos sandy colluvium samples.
Plate 4.8. Steep slopes of Vigles Hill overlooking the Libyan Sea. The low shrubs are separated by a network of sheep trails.
Plate 3.17. Cumulative grain-size plot of Kommos fill sediments. Plate 3.18. Correspondence analysis of 52 Kommos sediment samples and seven size variables. Plate 3.19. Ad hoc sea-level rise curve for Kommos-Matala area over the past 4,500 years (Not based on radiocarbon-dated material). Plate 3.20a-g. Schematic block diagrams of Kommos-Papadoplaka area, showing proposed coastal evolution over the past 4,500 years. Plate 3.21. X-ray diffractograms of clay from the Kommos excavation (13A/3:29) and from an outcrop at the base of Vigles Hill, south of the site. Chapter 4 Plate 4.1. Balloon photograph of Vigles Hill and sand dunes immediately south and east of Kommos, north at top (W. and E. Myers, 1976). It shows the locations of sampling sites: scattered and discontinuous shrub (sites 2, 5, 7, 18, 19, and 20), continuous shrub (6 and 17), scattered shrub on sand (3 and 4), and grassland with shrub (22). This is one of several balloon photographs that were used to draw the vegetation map of the area (Pl. 4.13). Plate 4.2. Olive orchards and fields in the Matala Valley. Shrubs cover the rocky hillside in the foreground and the distant hills. Plate 4.3. Beach and sea cliff below Kommos with a sparse cover of grasses and low shrubs. Plate 4.4. The shore of the Libyan Sea, the Kommos excavation, a grove of tamarisk trees, and the sand dunes with a sparse cover of ononis shrubs. Plate 4.5. The seaward slope of Vigles Hill with the sand dunes below. The mouth of the Pitsidia Valley is in the background. Plate 4.6. Fallow fields in the Matala Valley, clothed with spring flowers.
Plate 4.9. Large deciduous oaks near Zaros. Plate 4.10. The Kommos survey area, showing the 23 vegetation sampling sites and the routes travelled during botanical reconnaissance. Plate 4.11. Ordination plot of 23 vegetation sampling sites arranged by similarity of composition using the cover of shrubs and herbs. The circle encloses the twelve sites not shown by number. For site locations see Pl. 4.10. Plate 4.12. Ordination plot of 18 sites arranged by their similarity of composition using the cover of shrubs. Plate 4.13. Vegetation map of the immediate Kommos area based on balloon photographs taken before the Kommos excavations began. Plate 4.14. Pinus halepensis (pine) modern charcoal in tangential longitudinal section showing elongate tracheids (T), resin canals (C), and numerous rays (R) 3 to 7 cells high (SEM x ca. 90). Plate 4.15. Pine charcoal from Kommos of MM III-LM I age (43A/94) in tangential longitudinal section showing elongate tracheids, resin canals (C), and many rays 3 to 7 cells high (SEM x ca. 90). Plate 4.16. Cupressus sempervirens (cypress) modern charcoal in tangential longitudinal section showing elongate tracheids (T) and rays (R) 1 cell wide and 4 to 7 cells high (SEM x ca. 180). Plate 4.17. Cypress charcoal from Kommos of MM III-LM IA age (42A/50) in tangential longitudinal section showing elongate tracheids and rays 1 cell wide and up to 14+ cells high (SEM x ca. 90). Plate 4.18. juniperus (juniper) modern charcoal in tangential longitudinal section showing elongate tracheids (T) and rays (R) 2 to 3 cells high (SEM x ca. 190).
XXIl
Plate 4.19. Juniper charcoal from Kommos of Orientalizing-Classical age (43A/4) in tangen tial longitudinal section showing elongate tracheids and rays 1 to 3 cells high (SEM x ca. 170). Plate 4.20. Quercus coccifera (Kermes oak) mod ern charcoal in transverse section showing ves sels (V) in radial groups or flares (SEM x ca. 100). Plate 4.21 Evergreen oak charcoal from Kommos of MM III-LM IB age (42A/54) in transverse sec tion showing vessels in radial groups or flares; the charcoal surface is uneven (SEM x ca. 55). Plate 4.22. Ficus carica (fig) modern charcoal in transverse section showing large scattered ves sels (V) that are single or in pairs, and some times in threes. Tangential bands of thickwalled cells (B) alternate with bands of thinner walled cells. The rays (R) are several cells wide (SEM x ca. 45). Plate 4.23. Fig charcoal from Kommos of LM I age (53A/34) in transverse section showing large scattered vessels in pairs and threes. Tangential bands of thick-walled cells alternate with bands of thinner walled cells (SEM x ca. 60). Plate 4.24. Platanus orientalis (plane) modern charcoal in transverse section showing numer ous vessels (V) filling the space between the wide rays. The rays (R) are composed of large cells (SEM χ ca. 95). Plate 4.25. Plane charcoal from Kommos of LM I age (50A/69) showing numerous vessels filling the space between the wide rays (SEM x ca. 95). Plate 4.26. Crataegus sp. (hawthorn) modern charcoal in transverse section showing numer ous vessels (V) fairly evenly distributed between narrow rays (R) (SEM x ca. 95). Plate 4.27. Pyrus I Crataegus/ Malus charcoal from Kommos of Protogeometric-Geometric age (42A/ 74) in transverse section showing numerous vessels (V) fairly evenly distributed between narrow rays. The charcoal surface is uneven (SEM X ca. 95). Plate 4.28. Primus aulas (almond) modern char coal in transverse section showing large vessels (V), which are often in radial rows, and wide
List of Plates rays (R) of large cells. Much of the matrix in this photograph is composed of very thick-walled cells (fiber-tracheids, F). An annual growth ring (G) traverses the photograph (SEM x ca. 180). Plate 4.29. Almond charcoal from Kommos of Protogeometric-Classical age (43A/86) in slight ly oblique transverse section. The large vessels (V) are in radial rows; the rays (R) are of large cells. F indicates fiber-tracheids (SEM x ca. 135). Plate 4.30. Ceratonia siliqua (carob) modern char coal in transverse section showing large vessels (V), some single, some in groups of two or three. The vessels are widely spaced in a matrix of smaller cells. The rays (R) are distinct and several cells wide. There are several splits in the charcoal (*) (SEM x ca. 50). Plate 4.31. Carob charcoal from Kommos of MM III-LM IB age (42A/54) showing large ves sels. The vessels are widely spaced in a matrix of smaller cells. The rays are distinct. There is a short radial split in the charcoal (*) (SEM x ca. 70). Plate 4.32. Ononis natrix (ononis) modern char coal in tangential longitudinal section showing the ends of the numerous multicellular rays (R) among a narrow network of elongate vessels and tracheids (SEM χ ca. 45). Plate 4.33. Ononis charcoal from Kommos of Orientalizing age (42A/15) showing the ends of numerous multicellular rays among a narrow network of elongate vessels and tracheids (SEM x ca. 90). Plate 4.34. Pistacia lentiscus (lentisc) modern charcoal in transverse section showing large (V) and small vessels in groups or short radial rows. The rays (R) are 1 to 5 cells wide (SEM x ca. 95). Plate 4.35. Lentisc charcoal from Kommos of Orientalizing age (42A/24) in a slightly oblique transverse section showing large and small ves sels in radial rows. The rays are 1 to 5 cells wide (SEM x ca. 95). Plate 4.36. Acer sempervirens (maple) modern charcoal in transverse section. The wood is dif fuse porous with vessels (V) scattered through out, and there are conspicuous rays (R), up to 5 cells wide. An annual growth ring (G) crosses the photograph (SEM x ca. 95). Plate 4.37. Maple charcoal from Kommos of
List of Plates MM-LM I age (43A/90) in transverse section. Lhe wood is diffuse porous with scattered vessels and conspicuous rays (SEM x ca. 95). Plate 4.38. Rhamnus lycoides (buckthorn) modern charcoal in transverse section showing dendritic pattern of vessels (V) and two distinct growth rings (G) (SEM x ca. 95). Plate 4.39. Buckthorn/phillyrea charcoal from Kommos of LM I age (50A/69) showing dendritic pattern of vessels (SEM x ca. 95). Plate 4.40. Tamarix sp. (tamarisk) modern charcoal in transverse section showing large single vessels (V) and vessels in groups between the wide multicellular rays (R) (SEM x ca. 85). Plate 4.41. Tamarisk charcoal from Kommos of LM I age (53A/56) showing large vessels in groups between the wide multicellular rays (SEM x ca. 95). Plate 4.42. Erica arborea (heather) modern charcoal in transverse section showing numerous scattered single vessels (V) between rays (R) that are several cells wide (SEM x ca. 90). Plate 4.43. Heather charcoal from Kommos of MM-LM III age (43A/102) in transverse section showing scattered single vessels and a ray that is several cells wide (SEM x ca. 130). Plate 4.44. Olea europaea (olive) modern charcoal in transverse section showing single vessels and vessels in radial groups of two or three (V). The rays (R) are conspicuous and 2 to 3 cells wide (SEM x ca. 180).
XXlIl
ment; d, Columba livia, left scapula, proximal end. All from 12A3/47 (MM III-LM I). Bottom: e, Columba livia humerus; f, Apus tarsometatarsus; g, Columba livia tarsometatarsus. All from 12A1/81,83 (LM II). Plate 5.3. Fish cranial bone and otolith measurements (by M. J. Rose). Plate 5.4. Significant shell species at Kommos. A, Patella caerulea and Patella lusitanica (four examples); B, Monodonta turbinata; C, Cerithium vulgatum; D, Murex trunculus; E, Murex brandaris; F, Thais haemastoma; G, Glycymeris glycymeris with water-worn hole at umbo; H, water-worn Glycymeris "checker"; I, Columbella rustica; J, Pisania maculosa; K, Donax trunculus; L, Erosaria spurca; M, fossil oyster. (AU shells are from the LM II Space 7 dump in the Central Hillside, except for F, which is from the LM III deposits in the Southern Area.) Plate 5.5. Selected ornamental shells from Kommos. A, Bittium with hole opposite mouth (LM III); B, gastropod-bored Murex brandaris (MM III); C, water-worn Murex brandaris with hole on ventral side (LM IIIA2-B); D, water-worn Thais with hole opposite mouth (LM IIIA2-B); E, Euthria with hole opposite mouth (MM III); F, Conus holed on side of mouth (LM IIIA2); G, Natica holed on lower body (LM III); H, Arcularia with open dorsum and ventral hole (LM II); I, Monodonta with ground-down apex (LM IIIA1); J, Phalium lip fragment (MM III); K, Acanthocardia holed at the umbo (LM IIIA2-B); L, Dentalmm (MM III).
Plate 4.45. Olive charcoal from Kommos of MM III-LM age (IB, test in A/8A:7) showing single vessels and vessels in radial groups of two or three. The rays are conspicuous and 2 to 3 cells wide (SEM x ca. 185).
Plate 5.6. Charonia sequenzae found built inside the wall between Rooms 4 and 6, CH, of LM IHA date (Sh 4B).
Chapter 5
Plate 5.8. Fossil Strombos from a mixed LM III and later deposit in the Southern Cliffside (4A2/33).
Plate 5.7. Charonia sequenzae from the floors of Rooms N12 and N13 of the North House on the Hilltop, of LM IIIA2 date (Sh 1).
Plate 5.1. Diagram of bird skeleton. Forelimb: coracoid, scapula, humerus. Wing: radius, ulna, carpal, carpometacarpus. Upper leg: femur. Mid leg: tibiotarsus. Lower leg: tarsometatarsus, toes, claws.
Plate 5.9. Apodemus mystacinus from MM III Kommos. aa.l: M 2 dext., occlusal and occlusolabial views.
Plate 5.2. Bird bones. Top: a, Columba livia, left humerus, proximal end; b, Columba livia, left tibiotarsus shaft fragment; c, Falco sternum frag-
Plate 5.10. Apodemus cf. sylvaticus from Kommos. A, g.10 (Temple C); B,' zz.l (Temple Bl-2); and C, 111.4 (Temple B3): M 1 and M 2 dext., oc-
List of Plates
xxiv clusal views; D, g.5 (Temple C): M1 sin. occlusal and occluso-labial views; E, nn.l (Temple B3): M 1 and M 2 dext., occulsal and occluso-labial views; F, uu.2 (Temple B3): M 2 sin., occlusal view. Plate 5.11. Acomys mos. A, c. 12 and view; C, c.l: M 1 - 3 M 1 - 3 sin., occulsal
sp. from Temple C at KomB, c.13: M1 dext., occlusal dext., occulsal view; D, c.7: view.
Plate 5.12. Mus cf. domesticus from Kommos. A, g.7 (Temple C); B, kkk.l; and C, uu.3 (both Temple B3): M 1 dext., occlusal view; D, d.4 (Temple B3): M1 and M 2 dext., occlusal view; E, 111.3 (Temple B3): M 1 - 3 dext. occlusal view; F, zz.2 (Temple Bl-2): M 1 and M 2 dext., occlusal view; G, tt.l (Temple B3): M 1 and M 2 dext., occlusal view; H, d.l (Temple B3): M 1 - 3 sin., occlusal view; I, ww.l (Temple A); J, u u . l (Temple B3); and K, c.4 (Temple C): M 1 - 3 dext., occlusal view; L, d.2 (Temple B3): MM1 and M 2 dext., occlusal view. Plate 5.13. Top: Mus cf. domesticus from Kommos. A, kkk.l; B, d.4; C, uu.3 (all Temple B3); and D, g.7 (Temple C): M 1 dext., labial and antero-occlusal views. Bottom left: Mus domesticus from the southern Argolis, Greece. E, SP 626; and F, SP 628: M1 dext., labial and anteroocclusal views. Bottom right: Mus abbotti from the southern Argolis. G, SP 721; and H, SP 770: M 1 dext., labial and antero-occlusal views. Chapter 6 Plate 6.1. Map of soil distribution in the Kommos region. Compare with Table 6.1. Plate 6.2. Sketch of Mesara paleosols profile.
Plate 6.7. Map of land use in the Kommos region. Plate 6.8. Map of present-day land use potential. Chapter 7 Plate 7.1. Acropolis of ancient Metallon from north. Plate 7.2. Southern Matala Valley from eastnortheast. Plate 7.3A. Upright stones at point 23 on MetalIon acropolis, from south. Plate 7.3B. Cuttings in quarry to west of Metallon acropolis, from south. Plate 7.4. North side of Matala Bay. Plate 7.5. Southern Matala Valley from northwest. Plate 7.6. Aqueduct at site 37, point D from northeast. Plate 7.7. Trypeto Spilaio from southeast. Plate 7.8. Orthes Petres from northeast. Plate 7.9. Arolithia from south. Plate 7.10. Ayios Antonis: architectural fragments. Plate 7.11. Ayios Antonis: architectural fragments. Plate 7.12. Arolithia from northwest. Plate 7.13. Site 51 from east, with Nisos in distance.
Plate 6.3. East-west cross section to show geomorphic position of paleosols north of Kommos. 1, Kommos hill (= site 75 locality on Plate 6.1); 2, Pitsidia stream valley.
Plate 7.14. Roman tomb at site 100 from eastsoutheast.
Plate 6.4. Soil loss for the Kommos region recalculated using the USLE (Wischmeier and Smith 1978).
Plate 7.16. Mouth of cistern to east of site 118, from east.
Plate 6.5. Cross section of typical massive terrace. "Added soil" was dug from below the retaining wall immediately upslope. Plate 6.6. Cross section of typical dry-wall terrace. "Added soil" was dug from below the retaining wall immediately upslope.
Plate 7.15. Charakas (site 18) from southeast.
Plate 7.17. Mouth of cistern to north of site 118, from north, with site 118 behind. Plate 7.18. Kommos and Vigles from south. Plate 7.19. Vigles (site 66) from north-northeast. Plate 7.20. Circuit wall on Vigles: point A from northeast.
List of Plates
xxv
Plate 7.21. Circuit wall on Vigles: point B from east-northeast.
Plate 7.36A. Late Orientalizing to Archaic sherds from sites 70 and 10 (Vigles).
Plate 7.22. Circuit wall on Vigles: point C from east-northeast.
Plate 7.36B. Late Orientalizing to Archaic sherds from sites 70 and 10. Scale 1:2.
Plate 7.23. Lower part of Kalamaki stream valley from southeast.
Plate 7.37A. Head of Daedalic figurine from site 10.
Plate 7.24. Sendones and Langos from southeast.
Plate 7.37B. Drawing of Daedalic figurine from site 10.
Plate 7.25A. Evangelistria chapel from southsoutheast.
Plate 7.38A. Late Orientalizing to Archaic sherds from site 66 (Vigles).
Plate 7.25B. Ashlar foundations to southwest of Evangelistria chapel.
Plate 7.38B. Late Orientalizing to Archaic sherds from site 66. Scale 1:3.
Plate 7.26A. Evangelistria: block with "swallowtail" clamp.
Plate 7.39A. Hellenistic pottery from sites 2, 8, 73 (ancient Metallon), and 86.
Plate 7.26B. Evangelistria: horizontal cornice block (incomplete).
Plate 7.39B. Classical and Hellenistic sherds from sites 2, 8, 33, 73, and 86. Scale 1:3.
Plate 7.26C. Evangelistria: marble block with relief decoration.
Plate 7.39C. Sherds from sites 33 and 49.
Plate 7.26D. Evangelistria: impost capital. Plate 7.26E. Inscribed block: side A. Plate 7.26F. Inscribed block: side B. Plate 7.26G. Side B: detail of inscription.
Plate 7.40A. Hellenistic and Roman sherds from site 38. Plate 7.40B. Hellenistic and Roman sherds from site 38. Scale 1:3. Plate 7.41A. Medieval sherds from site 13.
Plate 7.27. Ancient walls at site 89.
Plate 7.41B. Medieval sherds from site 13. Scale 1:3.
Plate 7.28. Ancient terrace wall below site 81 from south-southwest at point 4.
Plate 7.42A. Medieval sherds from site HO.
Plate 7.29. Kalamaki stream valley from east.
Plate 7.42B. Medieval sherds from sites 108 and 110. Scale 1:3.
Plate 7.30. Site 33 from east-northeast. Plate 7.31. Pitsidia from north, with site 81 in foregound. Plate 7.32. View to north from Pitsidia. Plate 7.33A. Minoan sherds from site 25. Plate 7.33B. Minoan sherds from site 25. Scale 1:3.
Plate 7.43. Medieval sherds from site 108. Plate 7.44A. Pegasus buckle relief from east of Pitsidia. Plate 7.44B. Pegasus buckle relief from east of Pitsidia. Plate 7.45A. Medieval sherds from site 51.
Plate 7.34A. Minoan sherds from site 50.
Plate 7.45B. Medieval sherds from site 51. Scale 1:2.
Plate 7.34B. Minoan sherds from site 50. Scale 1:4.
Plate 7.46. Location map of sites in the Kommos Survey Area.
Plate 7.35A. Minoan sherds from site 57 (57-6 is later).
Plate 7.47. Minoan sites in the Kommos Survey Area.
Plate 7.35B. Minoan sherds from site 57 (57-6 is later). Scale 1:3.
Plate 7.48. Post-Minoan sites in the Kommos Survey Area.
List of Plates
XXVl
Plate 7.49. The Kommos and Vigles areas.
Plate 7.75. Sketch plan of site 33.
Plate 7.50. Sketch map of ancient Metallon.
Plate 7.76. Sketch plan of foundations at site 55.
Plate 7.51. Sketch plan of site 94. Plate 7.52. Sketch plan of Matala Bay (1898, British Admiralty Navigational Chart). Plate 7.53. Sketch map of sites 37, 76, and environs. Plate 7.54. Sketch sections of aqueduct at site 37. Plate 7.55. Sketch plan of site 13 (Didyma Hadzibraga). Plate 7.56. Sketch plan of site 8 (Orthes Petres). Plate 7.57. Sketch map of Arolithia area. Plate 7.58. Sketch plan of site 38 (near Ayios Stephanos). Plate 7.59. Anta capital and voussoir block from site 110 (Ayios Antonis). Plate 7.60. Sketch plan of site 39. Plate 7.61. Sketch plan of knoll at site 51. Plate 7.62. Sketch plan of Roman tomb at site 100. Plate 7.63. Sketch plan of Roman foundations at site 118. Plate 7.64. Sketch plan of wall foundations at site 6 (Vigles). Plate 7.65. Curving foundations at site 70 (Vigles). Plate 7.66. Wall foundations at site 66 (Vigles). Plate 7.67. Plan of foundations at site 10 (Vigles). Plate 7.68. Plan of part of circuit wall on Vigles, at point B. Plate 7.69. Sketch plan of foundations (of tower?) at site 62. Plate 7.70. Sketch map of Evangelistria district. Plate 7.71. Plan of remains at site 56. Plate 7.72. Plan of Evangelistria (site 5). Plate 7.73. Sketches of swallowtail clamps at site 5. Plate 7.74. Sketch map of Langos-Sendones area.
Chapter 8 Plate 8.1. Detail of sedimentary raw materials used for ground stone implements at Kommos. A. Poorly sorted pebble to granule conglomerate. B. Granule conglomerate with admixture of coarse sand. C. Fossiliferous limestone. Plate 8.2. Detail of sedimentary raw materials used for ground stone implements at Kommos. A. Grey-green, well-cemented, fine-grained sandstone. B. Beige, well-cemented, fine-grained sandy limestone. C. Grey phyllite. D. Black, metamorphosed chert. Plate 8.3. Type 1 ground stone implements (pebble to cobble size). Top row: GS 10, GS 7, GS 64 2nd row: GS 16, GS 119, GS 75 3rd row: GS 77, GS 52, GS 73 4th row: GS 24, GS 1, GS 13 5th row: GS 130, GS 25, GS 23, GS 62 Plate 8.4. Type 1 ground stone implements (cobble size). Top row: GS 8, GS 5 2nd row: GS 36, GS 56 3rd row: GS 20, GS 33 4th row: GS 2, GS 88 Plate 8.5. Type 1 ground stone implements (cobble to boulder size). Top row: GS 57, GS 134 2nd row: GS 91, GS 34, GS 89 3rd row: GS 117, GS 54 4th row: GS 120, GS 67 5th row: GS 114, GS 76 Plate 8.6. Type Top row: GS 2nd row: GS 3rd row: GS 4th row: GS
2 ground stone implements. 152, GS 168, GS 147 184, GS 182, GS 175 177, GS 163 146, GS 141
Plate 8.7. Type 2 ground stone implements. Top row: GS 179, GS 183, GS 149 2nd row: GS 148, GS 150, GS 159 3rd row: GS 142, GS 165, GS 160
List of Plates
XXVIl
Plate 8.8 Type 2 ground stone (metamorphic rock). Top row: GS 174, GS 170 2nd row: GS 158, GS 157 3rd row: GS 167, GS 151
implement
Plate 8.9. Type 2 ground stone implements (metamorphic rock). Top row: (S 443), GS 178 2nd row: GS 138, GS 180 3rd row: GS 185, GS 172 Plate 8.10.Type Top row: GS 2nd row: GS 3rd row: GS
2 ground stone implements. 169, GS 166 145, GS 153 143, GS 140
Plate 8.11. Type 3 ground stone implements. Top row: GS 191 2nd row: GS 212, GS 190 3rd row: GS 192, GS 186 Plate 8.12. Type 3 ground stone implements. Top row: GS 189, GS 208 2nd row: GS 187, GS 199 3rd row: GS 202 Plate 8.13. Type 4 ground stone implements. Top row: GS 223, GS 222 2nd row: GS 217, GS 224 3rd row: GS 218, GS 220, GS 221
B. Top row: GS 246, GS 250 2nd row: unidentified, GS 276 3rd row: GS 252 Plate 8.18. Type 5 ground stone implements (purple/red argillaceous arkose). A. Top row: GS 225, GS 254, GS 256 2nd row: GS 257, GS 258 3rd row: GS 266, GS 268 B. Top row: GS 269, (S 706), (S 922) 2nd row: GS 504, GS 297, GS 296 Plate 8.19. Type 6 ground stone implements. Top row: GS 326, GS 327 2nd row: GS 319, GS 318, GS 338 3rd row: GS 320, GS 365, GS 335 4th row: (S 683), GS 329, GS 339 Plate 8.20. Type 6 ground stone implements. Top row: GS 316, GS 307, GS 325 2nd row: GS 312, GS 311, GS 328 3rd row: GS 308, GS 310, GS 309 Plate 8.21. Type 6 ground stone implements. Top row: GS 315, GS 317, GS 331 2nd row: GS 305, GS 346, GS 334 3rd row: unidentified, GS 314, GS 323, GS 324 4th row: GS 332, GS 321, GS 342 5th row: GS 330, GS 313
Plate 8.14. Type 5 ground stone implements (well-cemented green and grey sandstone). Top row: GS 298, GS 229, GS 271, GS 249, GS 236 2nd row: GS 261, GS 247, GS 283, GS 281 3rd row: GS 245, GS 241, GS 265, GS 277 4th row: GS 259, GS 228, GS 263, GS 227
Plate 8.22. Type 7 ground stone implements (handstones). Top row: GS 402, GS 353 2nd row: GS 367, GS 401, GS 394 3rd row: GS 373, GS 374, GS 361 4th row: GS 348, GS 397 5th row: GS 352, GS 357, GS 369
Plate 8.15. Type 5 ground stone implements (green sandstone). Top row: GS 279, GS 243 2nd row: GS 278, GS 272, GS 291 3rd row: GS 274, GS 287, GS 244 4th row: GS 264, GS 275, GS 288
Plate 8.23. Type 7 ground stone implements (handstones). Left, top to bottom: GS 375, GS 392, GS 356, GS 363. Right, top to bottom: GS 380, GS 362, (S 391), GS 351, (S 1576)
Plate 8.16. Type 5 ground stone implements (phyllite and other metamorphic rocks). Top row: GS 282, (S 497) 2nd row: GS 292, GS 242, GS 246, GS 294 3rd row: GS 295, GS 273, GS 234, GS 270
Plate 8.24. Type 7 ground stone implements (handstones). Top row: GS 389, GS 383 2nd row: GS 393, GS 385 3rd row: GS 386, GS 360 4th row: GS 358
Plate 8.17. Type 5 ground stone implements (hematite, phyllite, and other raw materials). A. Top row: GS 262, GS 233, GS 253 2nd row: GS 237, GS 240
Plate 8.25. Type 7 ground stone implements (handstones). Top row: GS 382, GS 370
List of Plates
XXVlIl
2nd 3rd 4th 5th
row: row: row: row:
GS GS GS GS
355, 391, 395, 400,
GS GS GS GS
377 378 381 349
Plate 8.33. A. GS B. GS C-D. GS
Type 10 ground stone implements. 466 467 456
Plate 8.26. Type 8 ground stone implements (plastering and pigment-burnishing tools). Top row: GS 403, GS 417 2nd row: GS 406, GS 404 3rd row: GS 420, GS 410, GS 414 4th row: GS 407, GS 415, GS 405 5th row: GS 423, GS 413, GS 418
Plate 8.34. Type 11 ground stone implements (side view). Top row: GS 510, GS 512 2nd row: GS 509, GS 513 3rd row: GS 487, GS 502 4th row: GS 496, GS 489 5th row: GS 486, GS 493
Plate 8.27. Type 9 ground stone implements. Top row: GS 435, GS 429 2nd row: GS 436, GS 438 3rd row: GS 427, GS 431 4th row: GS 432, GS 442 5th row: GS 452, GS 444
Plate 8.35. Type 11 ground stone implements (view showing percussion-severing scar on implements in Plate 8.34). Top row: GS 510, GS 512 2nd row: GS 509, GS 513 3rd row: GS 487, GS 502 4th row: GS 496, GS 489 5th row: GS 486, GS 493
Plate 8.28. Type 9 ground stone implements. Top row: GS 430, GS 447, GS 455 2nd row: GS 448, GS 441 3rd row: GS 446, GS 437, GS 451 Plate 8.29. Type 9 ground stone implements (detail of metamorphosed chert examples). A. GS 426 B. GS 428 C. GS 439 D. GS 450 E. GS 454 Plate 8.30. Type 9 ground stone implements (detail of ground ends and pecked mid-sections on long margins). A. GS 450 and GS 454, side view showing pecked mid-section B. (top to bottom) GS 429, GS 436, GS 451, showing ground beveled ends. C. GS 450 and GS 454, detail of ground ends. D. GS 445 and GS 431, detail of ground ends. E. (top to bottom) GS 451, GS 448, GS 445, showing ground ends.
Plate 8.36. Type 11 (side view). Top row: GS 501, 2nd row: GS 518, 3rd row: GS 505, 4th row: GS 500, 5th row: GS 508,
ground stone implements GS GS GS GS GS
515 514 492 499 485
Plate 8.37. Type 11 ground stone implements (view showing percussion-severing scar on implements in Plate 8.36). Top row: GS 501, GS 515 2nd row: GS 518, GS 514 3rd row: GS 505, GS 492 4th row: GS 500, GS 499 5th row: GS 508, GS 485 Plate 8.38. Type 11 ground stone implements. A. Side view of GS 498 (upper left), GS 490, GS 517, GS 497 (lower left). B. View showing percussion-severing scar of implements in (A): GS 498 (upper left), GS 490, GS 517, GS 497 (lower left).
Plate 8.31. Type 10 ground stone implements. Top row: GS 464, GS 460, GS 456 2nd row: GS 465, GS 461, GS 471, GS 474 3rd row: GS 477, GS 469, GS 483, GS 475 4th row: GS 462, GS 482, GS 472, GS 459
Plate 8.39. Type 11 large-scale, ground stone implements. Top row: GS 507, GS 488, GS 504 2nd row: GS 516, GS 506 3rd row: GS 511, GS 495, GS 503
Plate 8.32. Type 10 ground stone implements and one type 1/10 implement. Top row: GS 466, GS 478, GS 467 2nd row: GS 476, (S 1561), GS 463, (S 417) 3rd row: GS 480, GS 473, GS 470
Plate 8.40. Type 11 ground stone implements. A. GS 490, side view B. GS 490, view showing percussion-severing scar
List of Plates
XXlX
C. GS 489, side view D. GS 489, view showing percussion-severing scar
C. Detail of GS 597A, showing abraded circumference and pecked central depression.
Plate 8.41. Type 11 large-scale, ground stone implements. A. GS 488 B. GS 506
Plate 8.48. Type 15 large-scale ground stone implements. A. Top row: GS 590, GS 597 2nd row: GS 582 B. Top row: GS 601, GS 599 2nd row: GS 600 C. GS 586, GS 587
Plate 8.42. Type 12A implements (naturally perforated pebbles and cobbles). Top row: GS 521, GS 535, GS 527 2nd row: GS 533, GS 543, GS 539, GS 528, GS 537 3rd row: GS 526, GS 525, GS 519, GS 540 4th row: GS 532, GS 520, GS 538, GS 544 5th row: GS 542, GS 523, GS 522, GS 536, GS 529 6th row: GS 534, GS 545, GS 541, GS 524, GS 531, GS 530 Plate 8.43. Type 12B ground stone implements (perforated andesite and phyllite, water-worn cobbles). Top row: GS 550 2nd row: GS 548, GS 546 3rd row: GS 549, GS 547 Plate 8.44. Type 12B and type 12C ground stone implements (perforated, limestone slabs). Top row: GS 552 2nd row: GS 554, GS 553 3rd row: GS 555, GS 551 Plate 8.45. Type 14 ground stone implements. Top row: GS 574 2nd row: GS 572, GS 573 3rd row: GS 571, GS 570 Plate 8.46. Type 15 ground stone implements (percussion-flaked and ground discs). Top row: GS 591, GS 585 2nd row: GS 592, GS 580, GS 588 3rd row: GS 594, (S 1570), GS 593 4th row: (S 1573), GS 596, GS 577 5th row: GS 584, GS 578, (S 920), GS 575, (S 1572) Plate 8.47. Type 15 ground stone implements. A. Top row: GS 575, GS 576, GS 577, GS 578 2nd row: GS 579, GS 580, GS 581, GS 583 3rd row: (S 740, vessel lid), GS 584, GS 596, (uncat.) B. Detail of GS 580, showing percussionflaked scars and abraded margins.
Plate 8.49. Type 16A stone implements (naturally formed, differentially weathered pebbles). Top row: GS 615, GS 619, GS 618, GS 603 2nd row: GS 609, GS 625, GS 621, GS 623 3rd row: GS 611, GS 626, GS 602, GS 620, GS 624 4th row: GS 604, GS 612, GS 617, GS 608, GS 605 5th row: GS 613, GS 607, GS 616, GS 614, GS 610 Plate 8.50. Type 16B ground stone implements (pestle-like tools and markers). A. GS 631, GS 215 B. Top row: GS 636, GS 628 2nd row: GS 635, GS 627 C. Top row: GS 632, GS 633 2nd row: GS 630, GS 629 D. GS 634 Plate 8.51. Type 17 ground stone implements (querns). Top to bottom: GS 659, GS 652, GS 658, GS 664 Plate 8.52. (querns). Top row: 2nd row: 3rd row:
Type 17 ground stone implements
Plate 8.53. (querns). Top row: 2nd row: 3rd row:
Type 17 ground stone implements
GS 650, GS 651 (S 1350), GS 662 GS 660
GS 655, GS 661 GS 657, (uncat. 30A/3:22) GS 647, GS 656
Plate 8.54. Type 17 ground stone implements (querns). A. GS 650 B. GS 655 C. (S 1299) D. GS 661
List of Plates
XXX
E. GS 664 F. (S 1141) Plate 8.55 Type 18 ground stone implements (mortars). Top row: (S 613), GS 675 2nd row: GS 685, GS 680 3rd row: GS 687, GS 692 Plate 8.56. Type 17 and type 18 ground stone implements (mortars). Top row: GS 686 2nd row: GS 653, GS 688 3rd row: GS 676, GS 677, GS 679 Plate 8.57. Type 18 ground stone implements (mortars). Top row: GS 684 2nd row: GS 690, GS 683 Plate 8.58. (mortars). Top row: 2nd row: 3rd row:
Type 18 ground stone implements GS 678 GS 689 (S 83), GS 668
Plate 8.59. Type 2OB ground stone implements (molds/anvils). A. GS 702 B. GS 700 C. GS 701 D. GS 703 Plate sins) A. B. C. D.
8.60. Type 18 (mortars) and type 19 (baground stone implements. GS 674 GS 695 GS 693 GS 696
Plate 8.61. Type 18 (mortars) ground stone implements. A. GS 669 B. GS 694 C. (S 615) D. GS 671 E. GS 682 F. GS 683 Plate 8.62. Large-scale, ground stone implements (type 18 and type 20D). A. GS 670 B. GS 681 C. GS 672 D. GS 681
E. GS 673 F. GS 706 Plate 8.63. Type 2OC ground stone implements (spouted press beds). A-B. GS 704, in situ C. GS 705, after repositioning Plate 8.64. Tool group found in MM HB dump. Top row: GS 89, GS 127 2nd row: GS 204, GS 176, GS 175 3rd row: GS 223, GS 388, GS 336, GS 390 4th row: (S 986), GS 94, GS 387, GS 95 5th row: GS 90, GS 283, GS 539, GS 133 Plate 8.65. Tool group found in LM IIIA1-2 sottoscala, North House (one large-scale, pebble conglomerate quern missing in photo). Top row: GS 668 2nd row: (S 191), GS 351, GS 352 3rd row: (S 183), GS 460, (S 1314), GS 459 4th row: GS 473, GS 432, GS 323 Plate 8.66. Late Minoan tool groups. A. LM IA tool group, trench 5B. Top row: GS 456, GS 137 2nd row: GS 348, GS 186 B. Tool group found on LM IIIA2 floor of trench H B l . Left: GS 652 Middle, top to bottom: GS 365, GS 364 Right, top to bottom: GS 26, GS 312, GS 27 Plate 8.67. Tool groups. A. Tools found in MM III deposit, trench 28B/3:62. Left, top to bottom: (S 551), (S 556), (stone under C 2216) Middle, top to bottom: (S 554), GS 64, (S 557), (S 561) Right, top to bottom: (S 555), GS 264, GS 412, (S 553) B. Tools found in the LM II dump, House with the Snake Tube. Left, top to bottom: GS 219, GS 115, GS 211 Middle left, top to bottom: GS 181, (S 539), GS 311 Middle right, top to bottom: GS 116, GS 156, GS 155 Right, top to bottom: GS 240, GS 627, (S 1444), (S 1199)
List of Plates
XXXl
Plate 8.68. Details of whetstones (type 5) and handstones (type 7). A. GS 299 B. GS 377 C. GS 280 D. GS 240 E. GS 349
Plate 8.74. Chert implements and debitage. Top row: CS 14, CS 31, CS 36, CS 37, CS 49, CS 58, CS 59 2nd row: CS 66, CS 65, CS 72, CS 83, CS 91 3rd row: CS 82, CS 56, CS 52, (S 2105) 4th row: CS 41, CS 89, CS 74, CS 53 5th row: CS 26, CS 64, CS 67
Plate 8.69. A. Water-worn volcanic cobbles from Late Minoan contexts at Kommos. Top row: (S 1551C), (S 1568), (S 1551D) 2nd row: (S 1550), (S 1553), (S 1551E) 3rd row: (S 1551F), (uncat. 51Al/3:66), (S 1551G), (uncat.) B. Abraded clay implements from Late Minoan contexts at Kommos (C 6 and C 5).
Plate 8.75. Utilitarian bone implements. Top row: Bo 2, Bo 23, Bo 25, Bo 13, Bo 3 2nd row: Bo 11, Bo 15, Bo 1, Bo 26, Bo 10 3rd row: Bo 6, Bo 7, Bo 19, Bo 9, Bo 12 4th row: Bo 18, Bo 14, Bo 22, Bo 16, Bo 21, Bo 4, Bo 5, Bo 28, Bo 27 5th row: Bo 8, Bo 17
Plate 8.70. Grooved objects and objects with multiple pecked depressions.
A. (S 1417) B. GS 573 C. GS 345 D. GS 566 E. GS 567 Plate A. B. C. D.
8.71. Grooved and perforated objects. GS 697 GS 302 GS 569 GS 565
Plate 8.72. (A) Perforated discs (type 12D) and (B) naturally polished pebbles (type 16C) plus one type 1. A. GS 557 (bottom left), GS 556 (top left), GS 558 (right) B. Top row: GS 26, GS 643, (S 1515), GS 640 2nd row: GS 46, GS 31, GS 47, GS 638, (S 465) Plate 8.73. Obsidian implements and debitage. Top row: CS 1, CS 2, CS 3, CS 8, CS 10, CS 12, CS 13, CS 19, CS 20, CS 21, CS 22 2nd row: CS 28, CS 32, CS 33, CS 40, CS 43, CS 46, CS 86, (uncat.), CS 96 3rd row: CS 4, CS 5, CS 7, CS 11, CS 23, CS 24, CS 27 4th row: CS 30, CS 35, CS 47, CS 42, CS 43, CS 81, CS 48 5th row: CS 6, CS 9, (S 1591) 6th row: CS 16, CS 25, CS 94
Plate 8.76. ments. Top row: 2nd row: 3rd row: 4th row:
MM III-LM I clay crucible fragM M M M
8, M 17, M 15, M 13 9, M 14 16, M 12, M 10, M 11 7
Plate 8.77. MM III-LM I crucible fragments, clay raw material, and LM III crucible fragments. A. M 7, from side, showing aperture for wooden handle. B. M 7, detail showing exterior wall and handle aperture. C. M 41, clay raw material found in association with Late Minoan crucibles. D. M 21, LM III crucible fragment. E. M 21, LM III crucible fragment, side view. F. M 21, LM III crucible fragment, front view. Plate 8.78. LM III crucibles and clay investment molds. A. M 19 B. M 23 C-D. M 19 E. M 31 F. M 37 G. M 30 Plate 8.79. LM III clay investment mold (M 30) for production of double axe by lost-wax process. A. Side view. B. Top view, showing preserved funnel. C. Detail of clay and chaff admixture in outer envelope of mold. Plate 8.80. Clay pot bellows. A. M 42
XXXIl
B. M 43, detail of nozzle aperture at join with wheelmade body of vessel. C. M 43, side view of bellows nozzle showing slip and inclusions. D. M 43, oblique view. E. M 43, base of nozzle. F. M 43, detail of nozzle aperture at join with bowl. Plate 8.81. Pumice tools. Top row: M 54, M 52, M 50, M 46 2nd row: M 51, M 48, M 47 3rd row: M 49, M 55, M 56, M 53 4th row: M 57, M 45 Plate 8.82. Metal raw material. A. Copper ingot fragments. Top row: M 2, M 3, M 4 2nd row: M 1, M 5, M 6 B. Sheet, lump, and prill fragments. Plate 8.83. Metal implements, casting remains, wires, etc. Top row: M 169 2nd row: M 128, M 166 3rd row: M 144, M 63 4th row: M 168, M 62, M 132, M 121 5th row: M 61, M 97, M 171 6th row: M 86, (B 194), M 85 7th row: M 135, M 95, M 133, M 60, M 170, M6 Plate 8.84. Metal objects. Top row: M 128 2nd row: M 63 3rd row: M 144 4th row: M 83, M 124, M 81 5th row: (B 194); M 82 Plate 8.85. Metal fishhooks and wire fragments. Top row: M 72, M 94, M 92, M 70 2nd row: M 71, M 78, M 149, M 170 3rd row: M 60, M 76, M 145, M 96 4th row: M 69, M 75, M 127, M 106 5th row: M 84, M 74, M 142, M 103 Plate 8.86. Metal rods, wires, strips, and other fragments. Top row: M 132, M 89, M 58 2nd row: M 167, M 116, M 104 3rd row: M 86, M 85, M 151 4th row: M 125, M 133, M 159, M 146, M 131 5th row: M 139, M 111, M 135, M 155 6th row: M 129, M 136, M 134, M 80, M 66 7th row: M 59
List of Plates Plate 8.87. Bronze double axe, chisels, and a large-scale needle. A. M 154, side view. B. M 154, bottom view. C. M 154, top view D. Left to right: Needle M 141 and chisels M 147 and M 150 Plate 8.88. Metalworking furnaces and hearths. A. Hearth, House with the Snake Tube B. Furnace bed, east of the House with the Snake Tube. Plate 8.89. Type 1, type 2, and type 7 ground stone implements. Scale 1:2. Plate 8.90. Type 2, type 4, and type 5 ground stone implements. Scale 1:2. Plate 8.91. Type 5 ground stone implements. Scale 1:2. Plate 8.92. Type 6 and type 7 ground stone implements. Scale 1:2. Plate 8.93. Type 8 ground stone implements. Scale 1:2. Plate 8.94. Type 8 ground stone implements. Scale 1:2. Plate 8.95. Type 7, type 9, and type 10 ground stone implements. Scale 1:2. Plate 8.96. Type 11, type 12B, type 13, and type 15 ground stone implements. Scale 1:2. Plate 8.97. Type 17 and type 18 ground stone implements. Scale 1:2; scale of GS 652, 1:4. Plate 8.98. Type 18 ground stone implements. Scale 1:3; scale of GS 678, 1:6; of GS 670, 1:9. Plate 8.99. Chipped stone implements: obsidian. Scale 1:1. Plate 8.100. Chipped stone implements: obsidian. Scale 1:1. Plate 8.101. Chipped stone implements: chert. Scale 1:1. Plate 8.102. Utilitarian worked bone implements. Scale 1:1. Plate 8.103. Utilitarian worked bone implements. Scale 1:1. Plate 8.104. MM III/LM I and LM III clay crucibles. Scale 1:3.
List of Plates Plate 8.105. LM III clay pot bellows and LM III clay investment molds. Scale 1:3; scale of M 42 and M 43, 1:6; of M 44, 1:9. Plate 8.106. LM III metal double axe and LM III clay investment mold. Scale 1:2.
xxxui Plate 8.107. Metalworking: primary products and waste. Scale 1:2. Plate 8.108. Metal implements. Scale 1:2. Plate 8.109. Pumice implements. Scale 1:1.
List of Tables Chapter 2 Table 2.1. List of staff during the years 1976-90, indicating the year, home institution, and role(s) on the Kommos excavation. Chapter 3 Table 3.1. Rainfall records of two stations nearest Kommos, Mires (for 1968-69) and Tymbaki (for 1966-70), in millimeters/month (Platakis 1971: 55). Table 3.2. Neogene rocks of the Aghia VarvaraAmbelouzos area (Meulencamp et al. 1977: 14546, table 4) and provisional correlations with the 1:50,000 Tybakion geological map, a simplified version of which appears as Plate 3.1. Table 3.3. General stratigraphy at Kommos survey site 56 (Kalamaki stream bank). Table 3.4. Characteristics of alluvial plains in the survey area. Table 3.5. Excavated soil sediments derived from a Kommos paleosol. Table 3.6. Earthquakes of magnitude greater than 5.5 that occurred within 50 km of Kommos during the period 1902-76 (data from Galanopoulos 1977: table 1). Ordinal depth of shock categories are S: shallow (0-65 km) and I: intermediate (65-300 km). Table 3.7 Fossils from Kommos. Chapter 4 Table 4.1. Environment vs. human adjustments and influences on the landscape of southern Crete.
Table 4.2. Kommos area soil characteristics. Values are mean ±1 S.D. Table 4.3. Correlations between soil characteristics and ordination scores for 23 sites. Values are Pearson's r. Ordinations were performed by reciprocal averaging on untransformed vegetation cover data (* = significant at > 0.05). Table 4.4. Flora of Kommos survey area, collected 1980-84 and compiled from herbarium specimens that were verified by W. Greuter. Table 4.5. Habitats of the Kommos survey area showing the association of land forms, soil parent material, land uses, and plant communities. Table 4.6. Plant cover and frequency in selected communities near Kommos as shown in Plate 4.1. Table 4.7. Potential use of plants found in the Kommos survey area. Table 4.8. Summary of environmental characteristics, human adjustments, and their influences on the Kommos landscape. Table 4.9. Percent frequency and abundance of selected charcoal taxa from Bronze Age levels at Kommos (derived from Appendix 4.5). Table 4.10. Percent frequency and abundance of selected charcoal taxa recovered from Bronze Age levels at Kommos by major period (derived from Appendix 4.6). Table 4.11. Charred seed remains in Bronze Age levels at Kommos. Table 4.12. Percent frequency and abundance of charcoal taxa recovered from Kommos and the nearest occurrence of each tree and shrub in modern habitats in the western Mesara. xxxv
List of Tables
XXXVl
Table 4.13. Modern uses of woody plants recov ered as charcoal from Bronze Age levels at Kommos (based on references listed in Table 4.7). Table 4.14. Physical properties of some Mediter ranean woods (from Grzeczynskiego 1972). Table 4.15. Summary of the wood economy at Kommos during Minoan times. Chapter 5 Table 5.1. Distribution of major mammal bones at Kommos. Table 5.2. NISP of the major domestic mam mals.
Table 5.18. Fish remains from pottery vessel contexts. Table 5.19. MNI of fish remains from Space 7, CH. Table 5.20. Fish taxa from Minoan Kommos and contemporary sites. Table 5.21. Identified cranial/facial bone mea surements. Table 5.22. Identified otolith measurements. Table 5.23. Context summaries for fish remains. Table 5.24. Marine invertebrates from Kommos. Table 5.25. Marine invertebrates from Kommos (by period).
Table 5.3. Minimum number of individuals of Ovis/Capra, Sus, and Bos.
Table 5.26. Major marine molluscs at Kommos.
Table 5.4. Major accumulations of mammal bone fragments.
Table 5.27. Major accumulations of Patella and Monodonta.
Table 5.5. Burnt mammal bones.
Table 5.28. Major deposits of GIycymens.
Table 5.6A. Age of sheep/goat.
Table 5.29. Major deposits of Murex trunculus (individuals).
Table 5.6B. The oldest and youngest ovicaprid individuals.
Table 5.30. Holed Murex and Thais.
Table 5.7. Deposits with pig, their age and sex.
Table 5.31. Arcularia from Kommos.
Table 5.8. Deposits with cattle or cattle-sized bones.
Table 5.32. sa «ία.
Table 5.9. Bird species present.
Table 5.33. Holed Bittium.
Table 5.10. Distribution of bird remains.
Table 5.34. Cockles.
Table 5.11. Estimated size of Epinephelus sp. at Kommos.
Table 5.35. Cowries.
Table 5.12. Sparidae and Centracanthidae verte brae. Table 5.13. M S P for fish taxa identified at Kom mos. Table 5.14. Actual and expected counts of sea bream skeletal elements. Table 5.15. Percentage of fish NISP by anatomi cal category and taxon, and presence by per centage of contexts.
Holed Euthria, Columbella, and Pi
Table 5.36. Dentalium. Table 5.37. Land snail species present at Kom mos. Table 5.38 Distribution of the major land snails. Table 5.39. Apodemus mystacinus. Table 5.40. Apodemus cf. sylvaticus. Table 5.41. Acomys. Table 5.42. Mus. Table 5.43. Measurements of shrews.
Table 5.16. Ratios of paired recovered cranial/ facial elements.
Table 5.44. Cretan small mammal faunas.
Table 5.17. Anatomical distribution of Kommos fish vertebrae.
Table 5.45. Pooled counts for the Kommos ro dent teeth.
xxxvii
List of Tables Chapter 6 Table 6.1. Soil properties in the Kommos region. Table 6.2. Properties of paleosol (B horizons) in the vicinity of Kommos. Table 6.3. Typical agricultural calendar of a middle-income Pitsidia farmer. Table 6.4. Example of land use survey form. Table 6.5. Summary of land use and production statistics. Table 6.6. Summary of areal extent of land use potential classes. Chapter 7 Table 7.1. Kommos survey sites listed numerically.
Table 7.2. Kommos survey sites listed alphabetically. Chapter 8 Table 8.1. Concordance of Kommos numbers and publication catalogue numbers. Table 8.2. Sandblasted ground stone implements from Kommos. Table 8.3. Material remains as evidence for onsite working of metal. * = Objects that occur at Kommos. Table 8.4. Relation of chronological exposure and recovery of ground stone tools.
C H A P T E R
1
The Topography and Archaeological Exploration of the Western Mesara Joseph W. Shaw
1. Introduction 2. The Western Shore
1. Introduction The fertile Mesara Plain (Frontispiece, PIs. 1.1-1.3),1 densely populated since the Neolithic Period, is the largest plain in Crete, some 50 km long (east to west) and 10 km wide (maximum). It is bordered on the northwest by Mount Ida (Psiloritis) and its foothills and on the west by the Libyan Sea. Visible from the coast on the west are two small uninhabited islands (ancient Litoa) known today as "Paximadhia," the modern name for segments of baked, dried village bread which they are thought to resemble. South of the plain are the rugged Asterousia Mountains, a center of the unusual Early Minoan remains with their characteristic large tholos tombs. On the east the hills rise up to the Viannos area and the mountains of Dicte. At approximately the midpoint of the plain the river Anapodaris (ancient Potherios) drains to the east, although today it is usually dry in summer. To the west, however, the Geropotamos River (its northern branch, the ancient Lethaios) drains the valley throughout the year and empties into the sea. No doubt this river was wider in ancient times when less groundwater was used for irrigation. The river's navigability from the sea remains doubtful for any historical period. As is often the case, exploration first by enlightened administrators, ecclesiastics, and travelers laid the ground for excavation. In the Mesara (and southern Crete in general) archaeological activity 2 began with Frederico Halbherr's most important discovery of the "Great Inscription" at Gortyn (1884). The impetus for Halbherr's exploration stemmed partly from the 1
2
Western Mesara
enthusiasm of his teacher, Domenico Comparetti, and the discovery of a fragment of the inscription in 1879 by B. Hassoulier, from the French School of Athens. 3 Halbherr subsequently continued his work in various parts of Crete: the Idean Cave of Zeus, Prinias, and, especially, Gortyn, where, in 1897, he began purchasing land for excavation. At about the same time he examined certain remains, probably once in prehistoric tombs, at Ayios Onouphrios, which he afterwards showed to Arthur (later Sir Arthur) Evans. Ayios Onouphrios was not far from the site of ancient Phaistos, which the English Captain T. A. B. Spratt first correctly identified during his exploration of the island (1851-53). Not long afterwards, Halbherr and his student Antonio Taramelli began work at Mesara sites such as Kourtes, Phaistos, Miamou, and Kamares, as well as the sanctuary of Asklepios at Lebena along the southern shore (Frontispiece). This was done with the support of the Archaeological Institute of America. This support, however, was soon withdrawn in favor of the excavations at Corinth on the Greek mainland. Backing for Halbherr's work was then provided from his native country and, with a formal permit, the Italians began in 1898 to excavate in the Mesara. In the following year the Italian Archaeological Mission was founded in Crete. This was also the year of Crete's independence from the Turkish yoke borne since the fall of Herakleion (then Candia) in 1669. The spring of 1900 saw the beginning of a new, fruitful era of exploration, one which led to the discovery of the Cretan prehistoric culture that was shortly to be dubbed "Minoan." On 23 March, Evans began work at Knossos where a Greek merchant, Minos Kalokairinos, had some years before (1878-79) excavated pithoi within storerooms of what was later recognized as the Palace of Knossos. In the same spring, at the end of May, Halbherr and his pupil, Luigi Pernier, started to reveal the Palace of Phaistos. The central court of that palace was cleared in 1901. Shortly afterward Halbherr, with the help of Enrico Stephani and Roberto Paribeni, excavated the luxurious residence at Aghia Triada, a site west of Phaistos and at the end of the same ridge projecting from the Mesara Plain (Paribeni 1904).4 Later the Italian School of Archaeology was founded (1910), with Pernier as its director, not long before Crete's political union with Greece (1913). At intervals, up to the First World War, Italian archaeologists worked with the Greek archaeologist Stephanos Xanthoudides in order to excavate numerous Early Minoan tholoi (published in 1924). Xanthoudides also excavated a series of Middle Minoan houses at Kalathiana. In the meantime, excavation of ancient Greek, and, especially, Roman monuments continued apace at Gortyn. After the First World War, the Italian Mission was re-established under Halbherr's direction, with a new generation of archaeologists going to Crete, including Luisa Banti, who would help write, among other works, the second of the two volumes dealing with the later Phaistos Palace (Pernier and Banti 1951) and the recent study of Aghia Triada (Halbherr et al. 1977). There was also Margherita Guarducci, who would complete Halbherr's epigraphical work in her monumental Inscriptiones Creticae (1935-50), and Doro Levi. Levi's work would continue almost unabated for over sixty years, beginning with his studies of Minoan seal-
Introduction
3
ings 5 and continuing with his excavations in 1924 of Archaic Greek Arkades in the easternmost reaches of the Mesara. After the Second World War he began a new series of excavations of Minoan settlements at Middle Minoan Patrikies and Aghia Fotini, also Late Minoan Kannia near Mitropolis at Gortyn and Middle to Late Minoan Chalara on the eastern slope of the Phaistos ridge as well as the Minoan tholos tombs at nearby Kamilari. Most important, he began excavating part of the western wing of the Middle Minoan palace of Phaistos which still lay below the great dump that had accumulated during Halbherr's and Pernier's excavations. A series of richly illustrated volumes, describing these projects and summing up another ambitious phase of Italian excavation in Crete, has now been published (Levi 1976-82). Noteworthy also are the important excavations (1954-61) conducted by Doro Levi, and then Giovanni Rizza, on the acropolis at Gortyn, where the latter excavated early Greek levels, the later Greek Temple, and the basilica (Rizza and Scrinari 1968). Vincenzo La Rosa has been particularly active in stratigraphic research as well as conducting new excavations at Aghia Triada (La Rosa 1977, 1985, 1989). While large-scale excavation was concentrated at Phaistos, Aghia Triada, and Gortyn, smaller sites were investigated elsewhere by both Italians and Greeks. Among them were one of Spyridon Marinatos's earliest investigations at Stou Kouse (1924; between Kouses and Siva, now filled with fieldstone) and, later, at Apodoulou along the northwestern border of the Mesara (1930-34; Marinatos 1933, 1935). During the Nazi occupation of Crete, German archaeologists, such as Frederick Matz and Ernst Kirsten, excavated at Apodoulou and nearby Monastiraki, and a tholos tomb at Apesokari due south of Aghii Deka (Matz et al. 1951). After the war Greek archaeologists excavated more tholos tombs, a series by Costis Davaras and Stylianos Alexiou (1958-67) in the Asterousia Mountains; one by Yannis Sakellarakis (1965, 1967) at Aghia Kyriaki; and another by Antonis Vasilakis (1981) at Odigitria and then Trypiti (1989). Protogeometric tombs at Rotassi were excavated by Nicholas Platon (1957), and others at Petrokephali in the following year by Giovanni Rizza (later to be published by Luigi Rochetti [Rochetti 1967-68]). A single LM III tomb from Goudies, near Mires, was published by Clelia Laviosa (Laviosa 1970). Near Kamilari, at Selli, Vicenzo La Rosa excavated a few poorly preserved Late Minoan houses (1972-73), not far from where Stylianos Alexiou had found a possible shrine to Demeter in 1957 (Platon 1957: 335). Most recently rescue excavations of Early Minoan remains took place at Kalamaki (by Despina HatziVallianou—loannidou-Karetsou 1978: 357; Hatzi-Vallianou 1979: 383; Vallianou 1992) and of Roman buildings at Matala (by Despina Vallianou and Antonis Vasilakis). As one can see from the foregoing, the pattern of excavation in the Mesara has been one partially concentrated in tombs and in particular in Early Minoan tholoi and Late Minoan chamber tombs, both with their many rich finds. On the other hand there was the important excavation of the Minoan palace at Phaistos and rich Minoan houses (e.g., those at Mitropolis and Aghia Triada) destroyed at the end of the LM I period. At Gortyn excavation has generally dealt with civic structures of the Roman period, although only a small part of this
4
Western Mesara
enormous site has yet been excavated and much that has been exposed remains unpublished. Excavation of Dark Age material has been less common but it includes some tombs (for instance, at Rotassi and Petrokephali), the temple strata on the Gortynian Acropolis and now the Iron Age layers at Kommos. Part of the Geometric/Archaic settlement at Phaistos was also cleared by Doro Levi. Neolithic architecture and Early Minoan habitation sites (as opposed to tombs) have been neglected, as have Greek and Roman levels outside of Gortyn. A few Christian basilicas received attention in the Gortyn area; Turkish remains have been ignored (e.g., the fort at Koules, near the village of Gregoria). Despite a few regional surveys, 6 there is a great need for further topographical work, especially in the vast eastern Mesara, which remains almost unknown archaeologically.
2. The Western Shore We now focus on the shore of the Libyan Sea, where Kommos lies. From the small fishing and tourist town of Aghia Galene (ancient Soulia) on the north to Matala (ancient Metallon) on the south, the distance is about 12 km, a good walk that lengthens as one trudges in the dry sand along a curving shore. Because of the military base north of the mouth of the Geropotamos River, however, the entire distance is rarely walked today. Walking from the rocky, cobble-strewn beach of Aghia Galene south, one reaches first the high, reddish and eroding cliffs of Kokkinos Pyrgos, a small settlement along low cliffs with its back to the northwest wind (meltemi) that prevails during the summer months. Kokkinos Pyrgos is the most convenient shoreside outlet for the inland town of Tymbaki, a spreading market town that has flourished with the recent establishment of hundreds of hothouses for growing vegetables, especially tomatoes and melons, sold at the more northerly markets on the mainland. South of Tymbaki, just across the Geropotamos, lies Aghia Triada, the Minoan site that was certainly important during the Middle Minoan period, before the major LM I residences were established there. After complete destruction by fire at the end of LM I, a fate shared by many other Minoan establishments on Crete, Aghia Triada was resettled. The spreading new town with its stoas and other buildings grew once more in importance, playing a leading role in the regional economy. No doubt, prior to its abandonment as a settlement in LM III, there was a road running west from it, parallel to the river, which now flows into a low, marshy area before it reaches the sea. South of Aghia Triada, and within the area covered by the Kommos survey (Chap. 7), there is an uninterrupted shoreline that continues all the way to the projecting rocky peninsula of Nisos, separating the Kommos region from that of Matala. In the northern area, not far from the mouth of the Geropotamos, is the small shoreside town of Kalamaki, the expanding sea outlet for the inland town of Kamilari. At points here, and for almost a kilometer inland, the sand has accumulated, in places many meters deep, making it difficult to grow certain types of crops, such as wheat, although beans can flourish on sunny, sandy slopes in the spring.
The Western Shore
5
To the south, beyond Kommos, is the Nisos peninsula, a dry and almost treeless expanse of stony fields, with little soil cover, minimal sand accumulation, and little sign of any real occupation over the millennia. South of Nisos, the Matala valley was extensively settled during Greco-Roman times and is presently a thriving and bustling tourist resort. Early visits by Europeans to this coastal strip, especially to the area south of the Geropotamos, were rather limited. Most travelers preferred to pass further inland, if they came by sea, in order to investigate the ruins of Gortyn and the great quarry ("The Labyrinth") there in the hills. Moreover, the wide, sandy, almost treeless shore or the stagnant swamps near the mouth of the Geropotamos River were hardly welcome points to rest. Also, for seafarers otherwise wishing to visit the coast, the great Bay of Mesara is particularly dangerous when gale-force winds descend the slopes of Ida and then sweep with surprising force out to sea, as Captain T. A. B. Spratt, mapping the coast of Crete for the Admiralty, learned in 1851 during a high wind when crossing the bay (Spratt 1865, 1: 11-23). Most early comments on the shoreline here have been written by people who landed at Matala, the small town discussed in detail later in this volume. The earliest and, surprisingly, the most wide-ranging account is that of the intrepid Florentine ecclesiastic Cristoforo Buondelmonte, who in 1415 first inspected Matala and then, leaving it by foot, discovered a large temple of Artemis in the plain to the north before being picked up by his ship just south of the Geropotamos River. Here was "Piriotissa," the site of a Genoese tower (in Greek: pyrgos) that was burned in 1558 by corsair pirates and which gave its name to the region (now Pyrgiotissa). 7 Near the tower Buondelmonte was accosted by an old woman who cried out to him: Ah, my children, please leave quickly from here. Leave this beach abandoned since antiquity, for if one is in good health and drinks the water or eats the fish along the shore one is no longer able to travel. (Van Spitael 1981: 112; translation by J. W. Shaw) "At these words," he continues, "we left the shore, and setting our oars in place, we started to raise anchor. We cursed the place and made haste to get away from its inhabitants." In 1630 Francesco Basilicata, a Venetian engineer in charge of public works in Crete, set about evaluating the possibilities of defending the island. Of the Mesara shoreline (PIs. 1.1, 1.2) he (like Buondelmonte) also remarked on the unhealthiness of the river, "the water isn't good and, moreover, is wasted because of the rotting of the linen works. This is the cause of the poor climate in the Mesara." Here is his verdict, as a military engineer, concerning the shoreline as a site for a possible invasion from the south: Any fleet with small boats can make landings on the shore, even though there are rocks in a few places and the water is so shallow that the ships must stand far offshore, and even small boats will approach the shore with difficulty. . . . This shore
6
Western Mesara is 45 miles from Candia. The road, almost complete, passes t h r o u g h t h e plain a n d t h r o u g h very delightful hills, a n d is the best a n d easiest for o n e to take from the s o u t h e r n side of the k i n g d o m to Candia. There the ships can stay w i t h safety, u n l e s s s o u t h e r n or w e s t e r n w i n d s blow. They can fetch w a t e r from the villages, e v e n if they are 3, 4, a n d 5 miles away. With t w o t h o u s a n d soldiers a n d 100 h o r s e s , h o w ever, it is possible to defend the shoreline. 8
O v e r a h u n d r e d years later Richard Pococke p a s s e d t h r o u g h the area, s t o p p i n g at Tymbaki ("Tribachi") w h e r e h e saw "the extraordinary c e r e m o n y of a Greek m a r r i a g e " (Pococke 1745: 249). H e t h e n crossed over the plain to Matala, w h i c h (as b o t h B u o n d e l m o n t e a n d Basilicata h a d d o n e before him) h e e x a m i n e d in s o m e detail, as w o u l d C a p t a i n Spratt a n d , m u c h later, a traveler, A. Trevor-Battye (1913: 285).
Notes 1. The most complete source for investigation of the topography and history of the Mesara is Margherita Guarducci, Inscriptiones Creticae, especially Volume 1, Tituli Cretai Mediae Praeter Gortynios (1935) and Volume 4, Tituli Gortynii (1950). Firsthand evaluation of literary evidence for ancient toponyms has been made by Paul Faure in a series of articles (1960a, 1960b, 1961). See also I. F. Sanders 1976 and 1982 and, for further information on internal political structure and the observation of religious rites in Crete, see Willetts 1962. The Frontispiece, of the Mesara area and southern coastline, is partially based on plans published by Faure (1960a: fig. 2 opposite 196) and Guarducci (1935, I), with topographical detail drawn by Thomas Boyd from a German survey (Griechenland 1:100,000, Sonderausgabe IX, Athens 1941, Grid O-P, Sections 12 and 13). When possible, ancient names of sites (many of which continue to remain unsure) are given in brackets. We are also obliged to Paul Faure and L. Vance Watrous for their advice on toponyms. 2. The following brief description is based partly on a study of Italian, Greek, and other archaeological work in the Mesara made by our excavation architect, Giuliana Bianco, when a student at the University of Toronto and partly upon an area bibliography begun by the author and then improved upon by Ian Begg. Only the major publications mentioned in the text are cited for this volume's bibliography. For more
information about the excavations up to shortly before the Second World War, see Pernier 1947 and, for the Mesara specifically during the Roman period, see I. F. Sanders 1976 and 1982. Pendlebury's The Archaeology of Crete (1939) remains useful; Hiller (1977) has recently published a bibliography covering the ten-year period of ca. 1965-75 in Crete. Also, a most informative history of Italian-sponsored excavation is to be found in Creta Antica (Di Vita et al. 1984). 3. Comparetti and the German philologist Ernst Fabricius subsequently published the inscription, of which the most recent translation and commentary is that by Willetts (1967). 4. Unfortunately, there was no complete publication of Aghia Triada until recently (Halbherr et al. 1977; La Rosa 1977, 1985, 1989). 5. See Levi 1925-26 for his publication of the sealings found earlier at Aghia Triada. 6. A few of these "surveys" are really reports of what was seen when certain routes were taken, as when Sir Arthur Evans visited the area in 1924 (see text below) or when J. D. S. Pendlebury and others walked through ten years later (Pendlebury et al. 1932-33). Some areas are considered by Margherita Guarducci and/or Paul Faure from the point of view of their ancient toponyms (Guarducci 1935-50; Faure 1960a, 1963); I. F. Sanders considered the range of GrecoRoman inhabitation (1976, 1982). Rather more complete, real searches, however, were made in
Notes limited areas. There is that, for instance, by Sinclair Hood, Peter Warren, and Gerald Cadogan (Hood et al. 1964) of the far northwestern edge of the Mesara; also that by David Blackman and Keith Branigan, as well by J. L. Bintliff, of the Agiofarango Valley far to the southeast of Matala (Blackman and Branigan 1975, 1977; Bintliff 1977a: 605-665; 1977b); and now that by ourselves in the immediate Kommos area. A major project that will place our own work in better perspective is the recent survey of the western Mesara, centered on Phaistos, conducted by a Greek-American team under the direction of L. Vance Watrous and Despina Vallianou (Watrous et al. 1993). 7. For Buondelmonte's account see now Van Spitael 1981: 47, 110-11, 226-29. For the tower see Gerola 1905: 260-61. 8. Francesco Basilicata, as in Spanakis 1969: 41, 195, 196. No doubt the Genoese tower at Piriotissa was built earlier with similar thoughts in mind. During the occupation of Crete in the Second World War, the entire shoreline south (and probably north) of the Geropotamos was heavily defended by the Nazis against a possible Allied landing from North Africa. Rightly so, for, before leaving Crete in 1941, the English
7 had examined the beaches at Matala, Kokkinos Pyrgos, and Aghia Galene as sites for landing stores; Kokkinos Pyrgos was considered at the time the only suitable spot (Hammond 1982: 141). At Kommos the southern approaches to the seaside hills, as well as the upper beaches, were protected from amphibious assaults by antivehicle mines (Pl. 2.7) and the vertical cliffs immediately south of Kommos were fortified with guns set into concrete hillside positions reached by tunnels. A bunker was set on the Vigles Hill (Pl. 7.49 at D). The most extensive fortification along the shoreline was at Asphendilias, just southeast of the shoreside town of Kalamaki, where heavy artillery positions, still impressive for their size and extent, were constructed along with underground galleries for storage and shelter from bombardment by planes and/or naval guns. The villagers of Pitsidia recall with grim faces the slave labor they furnished the Nazis during the construction of the extensive barracks there on the northern side of the hill and also the daily chore, sometimes accompanied by beatings, of carrying water from town to Asphendilias.
C H A P T E R
2
The Exploration and Excavation of the Kommos Site Joseph W. Shaw
1. Early Exploration 2. Later Exploration and the Excavation 3. Synopsis of Excavation (1976-1990) 4. Site Logistics 5. Site Recording 6. Recording and Cataloguing in Pitsidia Appendix 2.1. Diary Entry Recounting the Exploration of the Kommos Area by Sir Arthur Evans
1. Early Exploration Antonio Taramelli, who joined Halbherr in Crete and explored many areas of the island, was the first person to link the area southwest of the Geropotamos with activity during prehistoric times. Specifically, Taramelli noted the remains of Middle Minoan Kamares sherds south of Tymbaki and suggested that a main landing spot for Phaistos was not far south of there. He was also the first to connect the northern cliffs of Cape Nisos, immediately south and west of Kommos, with the possible site of the shipwreck which is said to have befallen part of Menelaos's contingent (Odyssey III.293-99) as the victorious warriors made their way back from Troy (Taramelli 1899: col. 296). The details of this theory, now adopted by a number of scholars (e.g., Aposkitou 1960: 157-58, Crile and Davaras 1963), and the possibility that there may be a connection between the possible wreck site and the Greek sanctuary found by us will be examined in detail at a later point; it suffices to say that, although not empirically provable, an argument to this effect can be extremely convincing. In the early summer of 1924 Sir Arthur Evans journeyed south from Knossos, first investigating portions of the shore south of the Asterousia Mountains and, then, that in the Matala 8
Early Exploration
9
area. Perhaps because he was aware of Taramelli's earlier suggestion, he also investigated thoroughly the area directly north of Nisos (Pl. 2.2, for Evans's plan) and first pointed to nearby Kommos as an important site of the Minoan period (A. Evans 1928: 88-92). With the help of his travelling companions, Duncan Mackenzie, Piet de Jong, and the Knossos foreman, Manolis ("lynx-eyed," as Evans would refer to him) Akumianakis as well as Myron Spithakis, who acted as mule boy, Evans conducted a brief but thorough archaeological survey of the area. His examination became the basis for his subsequent accounts of the site and for his interpretation of its potential significance. In his diary he scrawled with scarcely legible zeal (Pl. 2.1 and Appendix 2.1), "The end of the great South Road! and to great Minoan haven on the Libyan Sea!" Perhaps he wrote this while in his tent, said to have been yellow, while he camped that evening near the spring in the town of Pitsidia. His visit was still remembered and recounted by certain very old men in the area when interviewed during the first years of the excavation (M. C. Shaw 1980-81). Evans, with his earlier training as a newspaperman and his instinct for a good story, must have filed his report immediately upon his return to Knossos, because the first published account of which I am aware (dated by a correspondent in Athens to 15 July) was printed soon afterwards in the Morning Post on 23 July 1924. Other accounts were to follow, some being published as far afield as Lausanne, Adelaide, and Ceylon, but the basic dispatch concerning Kommos appeared in its most complete form in The London Times of 16 October 1924 (Evans 1924). Evans apparently incorrectly transliterated the modern name "Kommos" as "Komo". 1 In both accounts Evans's main theme is that Kommos was the chief port of the Mesara, connected with distant Knossos to the north by means of a paved pathway protected by occasional guard stations. Evans's view of the hegemony of Knossos as an initiating and controlling force in Crete was recently reinforced, at least for the LM II-IIIA1 period, by John Chadwick, on the basis of his gleanings from the Linear B tablets from Knossos. 2 Aside from a possible small excavation north of the Kommos site at Sphakoriako, however, Evans did not excavate, although in 1976 it was still said locally that he had been preparing to do so. 3 The closest that one comes in his accounts to excavation on the site is the mention of rows of pithoi reported by a local landowner, but not, as far as one can tell, seen by Evans himself. Their discovery, nevertheless, formed the basis for his picturesque label of the "Teloneion" or "Customs House" (Pl. 2.2), presumably established to accommodate the cargoes of Minoan ships returning from sea voyages abroad. Evans, however, did not elaborate on why this particular storeroom (and storerooms are typical of most large Minoan buildings) necessarily had a commercial, rather than a purely domestic, role. Evans also pointed out traces of what may be a raised pathway, still visible, which follows the top of the southern ridge, rather unlike many modern roads that follow the contours of the hills and more, we think, like a fortification wall of a later era (see Chap. 7). A road matching the description of one so presciently envisioned by Evans has been uncovered in the valley north of Vigles Hill, north of Minoan Building T and below the Greek Sanctuary at
10
Kommos Site
Kommos (J. W. Shaw 1982a: 192-93; 1982b). On the other hand, the intimate political connections he imagined to exist between Kommos and Knossos are, in our view, exaggerated (except, perhaps, during the period shortly after LM I, after the destruction of both Phaistos and Aghia Triada), for there is every likelihood that, if the construction of the road is to be connected with any palatial site, the more reasonable connection is to be found with the neighboring palatial settlement at Phaistos. This is not to disclaim, however, the valuable evidence supplied by Evans, who during the same trip had followed carefully the pattern of Minoan remains south from Knossos, up the Platyperama Valley and then down into the Mesara. 4 From the time of Evans's visit in 1924 to the beginning of our excavation in 1976, most archaeologists paid only moderate attention to the Kommos area. Before the Second World War there seems to have been no effort to excavate on the still somewhat inaccessible site and, after the war, the one project considered did not get under way because of the presence of a still lethal minefield. 5 Another factor contributing to this lack of attention may have been an increasing accumulation of sand that masked from view any walls visible on the surface east of the actual cliffside. Evans's plan (Pl. 2.2) notes quite a few walls, but none were visible there during the 1965-75 period when I was first exploring the area. The site, nevertheless, was visited and written about by others. Luigi Pernier, for instance, went there in 1930, perhaps in preparation for writing his survey of the Mesara region in his first volume on the Phaistos palace (Pernier 1935: 6-7). He accepted Evans's identifications and interpretation and collected some Middle and Late Minoan sherds. Not long afterward, in 1934, Pendlebury visited Kommos and confirmed Evans's views, although he questioned his identification of a tholos tomb on the southern hill, preferring to think of it as an apsidal building (Pendlebury et al. 1932-33: 80-81). In his justly famous The Archaeology of Crete (1939), he relies mainly on the evidence originally furnished by Evans. Failing to note any LM III sherds, however, he used the site as an example of those settlements deserted prior to this period by a populace seeking safer areas for occupation elsewhere (Pendlebury 1939: 239, n. 1). As excavation later revealed, the LM III period is unusually well represented at Kommos. Pendlebury was followed in 1937 by R. W. Hutchinson, who is said to have found Neolithic sherds there (Pendlebury 1939: 45), but they are perhaps similar to those recently published by Betancourt (1990: 22-26). Other recorded visits were made by Myron Zacharakis, an energetic high school teacher from Pobia (Zacharakis 1968: 16, 28), and (on the basis of the records at the Stratigraphic Museum of the British School of Archaeology at Knossos) in August 1955 by Sinclair Hood who collected fine ware, including some EM I and LM III. In the spring of 1963 William McDonald and Richard Hope Simpson brought back to Knossos Middle and Late Minoan sherds from the general area of the site including Early Minoan light on dark ware which they found in the vicinity of Evans's "ossuary" shown to the south on his sketch plan (Pl. 2.2).
Later Exploration I Excavation
11
Offshore and seaside observation was on the mind of some visitors. Nicholas PIaton, for instance, came to the conclusion that there was a "man-made construction in limestone set parallel to the shore" (Platon 1955: 564-65). Such a construction offshore, of which we have seen nothing during many snorkeling trips with mask, is, nevertheless, said by the locals, especially by the fishermen, to stretch north-south from a huge projection of stone (the νόlakas or boulder to the southwest) to a reef (the "Papadoplaka" or Priest's Slab, Pl. 3.11) still visible about 300 m off the shore. Apparently this mysterious wall is visible in some 10 m of water only when currents have scoured away the sand. Not long after Platon's visit, G. Crile and C. Davaras explored the nearby waters for evidence of wrecks (Crile and Davaras 1963). In the same year Honor Frost published a report on traces of harbor works under the sands on the shore (Frost 1963: 111-12), as did Paul Faure a decade later (Faure 1973: 45). So far, we have not seen such structures; a part of Building J may have appeared then, however, in a sandy scarp. 6 - 7
2. Later Exploration and the Excavation The author's own involvement with Kommos began during an informal visit in 1965, follow ing work at the Roman port town of Kenchreai in the Corinthia with Robert Scranton (then at the University of Chicago), and later with Nicholas Platon (then at the University of Thessalonika) at the Minoan harbor town and palace of Kato Zakros. The obvious links of these two sites with known trade routes had aroused an intense interest on my part in ancient harbors and seafaring. 8 During my first visit to Kommos it became clear to me that the site, although largely hidden by sand (Pl. 2.5), some of it quite deep, would nevertheless repay excavation, especially on the northern hill (Pl. 2.4), where excavation was later to begin. (For before and after views of the site taken from balloons by W. Myers and E. Myers, see PIs. 2.9-2.11.) Here, not far from a modern well, was a slope strewn with Middle and Late Minoan sherds (Pl. 2.3) and rubble that could only be from eroded walls and floors of a settlement. A wall, perhaps ancient, and a drinking trough for sheep and goats, cut out from an ashlar block, could be seen near the well (Pl. 2.3). Within a nearby deep pit, dug by the well's excavators in order to recover slabs to line the sides of the well, one could see ancient walls (Pl. 2.3), sug gesting that the fill to the east might be deep and, perhaps, stratified. This hope was subse quently borne out here on the hillside where a Late Minoan house was found neatly super posed on earlier, Middle Minoan remains. The pit had been dug down within what later came to be recognized as a room (Room 2) of the Late Minoan house, called "The House with the Snake Tube" in Part 2 of this volume. Far to the north, near the present border of the expropriated area, a heavy ashlar wall surrounded by Middle Minoan sherds could be seen in the almost vertical, overgrown sandy scarp (Pl. 2.6).
9
Originally the field work at Kommos was envisioned as a cooperative Greek-Canadian en terprise with Ephor Stylianos Alexiou (PL 2.27) and the author as co-directors, partly because
12
Kommos Site
of the difficulty in arranging for a normal excavation permit through a foreign school of ar chaeology. But when this could not be arranged, it became necessary to initiate work at Kom mos under the auspices of a foreign school. There being no Canadian archaeological base at the time (although one was established not long afterward, in 1976), after prolonged, friendly negotiations the American School of Classical Studies at Athens extended its hospi tality and generosity to include the granting of one of its three excavation permits. In the meantime a financial basis for the Kommos excavation was slowly being formed. Acquisition of the land, in particular a substantial part of the northern hill where the fill might be particularly deep, was a complex problem. The property had been acquired from the original owners in the 1950s by another party, and, despite some friendly exchange be tween the author and the owners (land speculators planning a hotel-bungalow complex), the latter did not give their permission to conduct trial excavations on the land, although it would have been possible to inspect the foundation trenches for the various structures to be built. Thus the only recourse was to attempt to arrange for the expropriation of the most promising section of land. Using as a basis a preliminary topographical plan made in 1972, formal expropriation procedures were initiated by the Greek Archaeological Service through an announcement in the government newspaper (5 October 1973). 1 0 1 1 The ancient site of Kommos actually consists of an area of some 40 stremmata (40,000 m 2 , approximately 10 acres) along the shore. 1 2 Along its southern edge, approximately even with the line of Nisos, there are sparse remains, chiefly sherds, of the Roman period. North of Nisos is the high "Vigles" ridge with its Early Minoan sherds (probably also walls) and sig nificant evidence of Archaic Greek occupation. To the north and still partly underlying the deep accumulation of sand is the Middle and Late Minoan town. The extent of the town to the south here remains unknown. The same is true for Greek settlement, although we know that Greek occupation stopped just north of the sanctuary. On the west, however, the spread of both town and sanctuary was naturally delimited by the cliffside scarp and upper shore line. On the northeast, at least on the Hilltop where the soil (rather than only sand) is visible in places, the town probably extended no more than 20 m beyond the line of expropriation. To the north, judging from the plentiful Minoan sherds visible in the cliffside scarp, the town must continue at least as far as the next hill, some 100 m away. This hill is the second of a series referred to locally as Τοϋ Σπανοΰ χα Κεφάλια (literally, "the heads of the hairless one"), either because of their general lack of covering growth or, according to a man in the Pitsidia village, because the property was once owned by a man naturally without a beard (a σπανός). South of the southern " h e a d " and the modern well, at the point where the ground begins to level out and join the shore, was a spot called locally "Pelekia," or "Hewn Stones." Ac cording to some of the older villagers, on two occasions blocks were dug out from the sand here, once around 1921 to build a bridge on the Platys River just south of the resort town of Aghia Galene and north of Kokkinos Pyrgos. Earlier, although we do not know when, other blocks had been removed from ancient buildings in order to build part of the Preveli Monas-
Later Exploration I Excavation
13
tery, far beyond Aghia Galene, on the coast and not far from Spili (M. C. Shaw 1980-81: 7). One considerate old man, George Sphakakis, whose shore house we rented as a storeroom for the tools for many years (Pl. 2.28), 13 claimed that the blocks of the buildings at Kommos were such that they were joined to one another by dowels, "male and female fitting parts," as he described them. Later, when excavation did take place in the area, it became clear that "Pelekia" referred in particular to the blocks robbed out of the side walls of Temple C and also, perhaps, to those from Minoan Building J, further west along the slope down to the sea. If the latter is true, then this was not the first time that Minoan buildings served as quarries, for there is evidence to show that Room A l , built in the Hellenistic Period, was constructed at least partially of wall blocks formed by sectioning the much bigger stones of Building J or some other nearby Minoan structure. During our own work no trace was ever found of the blocks "joined by dowels," mentioned by George, nor was there evidence to show that any form of metal clamp was used in the construction. Although we originally planned to excavate in a number of places, especially on Vigles, the pressure of full-scale work to the north, and the necessity for us to expropriate even before trial excavations could take place (and pay for same), forced us to restrict our efforts to this single, crucial area. Thus the excavation described below enumerates and interprets work only at one point. Excavation of a Minoan cemetery, for instance, which traditionally lies outside the settlement and which would be particularly informative from the point of view of ceramic development and foreign interconnections, was not possible. Nor has it yet been possible to examine the Early Minoan settlement on Vigles and determine the extent to which it served as the local center before habitation on the shoreside hills, where we did excavate. Perhaps in the future this will be possible. The original aims of the excavation, the first large-scale excavation in Greece originating largely through Canadian institutions, were defined in our first application to the Canada Council (now the Social Sciences and Humanities Research Council of Canada) of 26 February 1974. The aims were as follows: (1) If Kommos had been an important town, then there would be a chance of recovering imported remains which would contribute to the refinement of Minoan chronology. If the town were well preserved, we might also discover a series of valuable superposed occupation levels. (2) We might recover shoreside port facilities such as warehouses and other special buildings. (3) If undisturbed by later occupation, the prehistoric contexts at Kommos might be discovered intact. Given the above and assuming we were dealing with a small town, we planned to sample it adequately in five campaigns of six to eight weeks yearly. As in any excavation, there were surprises. One was the unusual lack of massive burned
14
Kommos Site
destruction at the end of LM I, a time when many Cretan sites suffered conflagrations. An other was the unexpected scale of prehistoric constructions on the lower part of the hill, ex tending far to the south. The event least foreseen by the earlier survey work occurred in 1977 with the partial uncovering of the Greek Sanctuary, buried meters deep in the sand with no visible traces of architecture or sherds on the surface. Moreover, Greco-Roman remains in effect stop at the base of the hill and sherds of that date were simply not in the seaside scarp at the time of the original surveys. These discoveries prolonged the excavation beyond the original estimate. They also prompted us to continue the excavation of the monumental Mi noan ashlar buildings after a break for preparation of publications. When that decision to halt excavation, at least temporarily, was taken, we asked the American School of Classical Studies at Athens to reassign the excavation permit which we had used for a decade.
3. Synopsis of Excavation (1976-1990) 1976
(SEASON
1)
During the first season, trenches on the Hilltop (Pl. 2.20 [blessing of the site] and Pl. 2.21 [the first trench on the Hilltop]) and Central Hillside exposed sufficient remains of houses to en able us to see that we had discovered a significant Middle through Late Minoan town. (For the general site plan as of 1985, see Pl. 2.12. For plans of the Hilltop Houses, see PIs. 2.13, 2.14 [trench plan]; for the Central Hillside, see PIs. 2.15, 2.16 [trench plan].) LM II pottery, unusual outside Knossos, was found in some quantity. On the Hillside, the tilted con figuration of the bedrock down and away from the sea permitted the preservation of strati fied buildings with their associated deposits. A sounding in the deep sand along the south ern border of the property revealed roof tiles; clearly there was Greco-Roman construction on the south. (We did not know it at the time, but the sounding had come down upon the top of the tiles collapsed on the floor of Temple C.) Preparation: 26 May-30 June. Excavation: 1 July-25 August. (See Table 2.1 for staff and the years they participated in the project.) Table 2.1. List of staff during the years 1976-90, indicating the year, home institution, and role(s) on the Kommos excavation. Years at Kommos Name
Institution and Assignments
Alhbone, Richard
University of Cambridge, Pottery Assistant
Banou, Eleni
University of Pennsylvania, Pottery Analyst
76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 χ χ
χ
χ
χ
(continued)
Synopsis of Excavation
15
(Table 2.1 continued) Years at Kommos Name
Institution and Assignments
Begg, Ian
University of Toronto; Trenchmaster
Beladakis, George
Pitsidia; Foreman
Besi, Helen
Profiler
Betancourt, John
General Helper
Betancourt, Mary
General Helper
Betancourt, Michael
General Helper
Betancourt, Philip P.
76 77 78 79 80 81 82 83 84 85 86 87 88 89 90
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
X
X
X
χ χ
χ
x x x x
x
x
x
x
x
Temple University; Trenchmaster, Pottery Analyst, Assistant Director (1980-82)
X
X
X
X
X
X
X
Bianco, Giuliana
University of Toronto; Architect/Artist
X
X
Bikai, Patricia
Ceramic Expert
Blitzer, Harriet
Indiana University; Trenchmaster, Tool Analyst
Brundage, Nana (in Toronto)
University of Toronto; Chairman's Secretary
Burke, Edwin
Sheridan College; Photographer
Callaghan, Peter J.
University of Cambridge; Pottery Analyst
Campbell, Erin
University of Toronto; Trenchmaster, Cataloguer
Clarke, Joseph
Artist
Comstock, Betsy
University of Toronto; Cataloguer
Conner, Patricia
College Year in Athens; Profiler
Cotsis, Danae
College Year in Athens; Profiler
Cox, Bill
University of Toronto; Trenchmaster
Cronkite, Susan
University of British Columbia; Trenchmaster
Csapo, Eric
University of Toronto; Trenchmaster, Editorial Assistant, Epigraphist
Dabney, Mary K.
Columbia University; Cataloguer, Trenchmaster
Dabney, Taylor
Pratt Institute; Photographer
χ χ
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
(continued)
16 (Table 2.1
Kommos Site continued) Years at Kommos
Name
Institution and Assignments
DeMita, Frank
Haverford College; Profiler
DeVinney, Timothy
Albion; Photographer
Diamant, Steven
College Year in Athens; Technician
Downie, Susan
University of Toronto; Student Assistant
Duckworth, Elisabeth
University of Toronto; Cataloguer
Duclos, Rebecca (Kommos and Toronto)
University of Toronto; Cataloguer
Elias, Rick
Boston University; Surveyor
Fisher, Elizabeth
University of Minnesota; Trenchmaster
Geagan, Daniel
McMaster University; Epigraphist
Gifford, John A.
University of Miami; Geologist Profiler, Cataloguer
Gilmore, Valerie Grant, Lynn
London Institute of Archaeology; Conservator
Grove, Lori
Western Michigan University; Profiler
Hamann, Barbara
London Institute of Archaeology; Conservator
Harlan, Deborah K.
University of Pennsylvania, University of Minnesota; Cataloguer
Hayden, Barbara
University of Pennsylvania; Pottery Assistant
Hayes, John W.
Royal Ontario Museum; Pottery Analyst
Hedreen, Guy
ASCSA; Profiler, Trenchmaster
Hemans, Fritz
Boston University; Surveyor
Hennckson, Robert
University of Toronto, Trenchmaster
Hollyer, Jenny (in Toronto)
University of Toronto; Student Assistant
76 77 78 79 80 81 82 83 84 85 86 87 88 89 90
x
x
x
x
x
x
x
x
x
x
x
x
x
(continued)
Synopsis of Excavation
17
(Table 2.1 continued) Years at Kommos Name
Institution and Assignments
Hope Simpson, Richard
Queen's University; Surveyor
Jarkiewicz, Zbigniew
University of Toronto; Trenchmaster
Johnston, Alan
University College London; Pottery Analyst
Koehl, Robert B.
University of Pennsylvania; Pottery Analyst
76 77 78 79 80 81 82 83 84 85 86 87 88 89 90
Kousoulos, Athanasios
Conservator
Leesti, Elizabeth (in Toronto)
University of Toronto; Secretary
Lesgard, Marta
Queen's University; Conservator
Lewis, David
College Year in Athens; Profiler
Manderson, Rosemary
University of Toronto; Profiler
McEnroe, Cathy
Indiana University; Cataloguer
x
x
x
McEnroe, John
University of Toronto, Hamilton College; Trenchmaster, Surveyor
x
x
x
x
x
x
McGowan, Elizabeth
ASCSA; Trenchmaster
Monroe, Carole (in Toronto)
University of Toronto; Student Assistant
Nixon, Lucia
University of Toronto, University of Cambridge; Trenchmaster
x
x
x
x
x
x
Norris, Michael
University of Cambridge; Pottery Assistant
Orr, Ann
General Helper
Orr, Douglas
University of Toronto; Trenchmaster
Parsons, Michael
Queen's University; Surveyor
Patterson, Erin
University of Cambridge; Pottery Assistant
Pfaff, Julia
University of Toronto; Profiler
Phillips, Jacke (Kommos and Toronto)
University of Toronto; Cataloguer, Profiler
x
x
x
x
x
x
x
x
(continued)
18 (Table 2.1
Kommos Site continued) Years at Kommos
Name
Institution and Assignments
76 77 78 79 80 81 82 83 84 85 86 87 88 89 90
Reese, David S.
University of Cambridge; Faunal Analyst
Safran, Elizabeth
University of Toronto; Profiler
Schwab, Katherine A.
Institute of Fine Arts, N.Y.U.; Cataloguer
x x x
Sease, Catherine
Bryn Mawr; Conservator
χ
Shaw, Alexander C.
Queen's University; Photographer
Shaw, Joseph W.
University of Toronto; Director
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
Shaw, Maria C.
University of Toronto; Trenchmaster, Assistant Director (1983- )
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
Shaw, Robin A.
Queen's University; Photographer's Assistant
X
X
Shay, C. Thomas
University of Manitoba; Floral Analyst
Shay, Jennifer M.
University of Manitoba, Floral Analyst
Shubert, Steven
University of Toronto; Trenchmaster
Silverman, Eric
College Year in Athens; Profiler
Spaeth, Barbette
Johns Hopkins University; Trenchmaster
Stavrinos, Vicki
General Helper
Stewart, Sarah (in Toronto)
University of Toronto; Student Assistant
Tenody, Janet (Kommos and Toronto)
Queen's University; Cataloguer
Thurston, Lorna (in Toronto)
Secretary
Turner, David
Pottery Assistant
Vincent, Fran
General Helper
Vincent, Robert K.,•Jr.
Yale University; Photographer
x x x x x
χ
χ
χ
χ
χ
x x x x
X
X
X
X
X
X
X
X
X
X
X
X
X
χ
χ
χ
x x x
Warren, Elizabeth
General Helper
χ
Warren, Peter
University of Bristol; Trenchmaster
χ
Watrous, L. Vance
SUNY at Buffalo; Pottery Analyst, Assistant Director (1980-82)
x x x x
x
x
x
x
(continued)
Synopsis of Excavation (Table 2.1
19
continued) Years at Kommos
Name
Institution and Assignments
Whittaker, Helene (Kommos and Toronto)
University of Toronto; Cataloguer
Wright, James C.
Bryn Mawr; Cataloguer, Trenchmaster
Zernask, Linda
Trent University; Profiler
1977
(SEASON
76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 χ χ
χ
χ
χ
χ
χ
χ
χ
χ x
2)
A major effort began during the late spring to remove all the sand covering the property. A large front-lifter (not a bulldozer, which cannot maneuver easily) cleared away without in cident (few mines were found and those found were sufficiently rusted not to explode [Pl. 2.7]) some 30,000 m 3 of pure sand from above the earth level where the prehistoric re mains begin. To our surprise we discovered, in the southern part of the property, a series of Hellenistic buildings (i.e., those in Pl. 2.22) facing onto a court with an altar in the middle, clearly a sanctuary of some type. One of these buildings (B) was excavated completely. (For the general period plan of the Southern Area, see PIs. 2.17, 2.18 [trench plan], and 2.19 [trench plan]). Work on the Minoan houses continued to the north. Preparation and sand-clearing: 15 May-15 June. Excavation: 15 June-15 August. 1978
(SEASON
3)
We began to sense the general organization and size of the houses of the Minoan town and the relationship of one room to the other and of one house to the next. More evidence emerged to cast doubt upon a massive LM IB destruction by fire. In the sanctuary more buildings were cleared of earth and sand (Al, D, E, F) and a second altar (H) was discovered in the court. We continued to d u m p sand and earth onto the sterile cliffside slopes and, on the southern part of the property, upon the beach where the winter waves later would spread it out evenly along the shore line. During this season a formal foot survey of the Kom mos area was initiated. Preparation and survey: 29 May-13 June. Excavation: 26 June-25 August. 1979
(SEASON
4)
Major excavations on the Hilltop completed full-scale work there. We had obtained a sample of Late Minoan buildings and contexts on that part of the site (three houses and portions of others). Work on the Hillside in Middle Minoan levels (the major Late Minoan houses there were now complete) exposed rooms of a Middle Minoan house, one quite filled with
20
Kommos Site
pottery, suggesting a destruction by earthquake. A new purchase of land on the south permitted the exposure of the chief building of the Greek Sanctuary, a temple of a known Cretan type, as well as two altars (L, M) south of those discovered earlier. An unusual Geometric/Archaic shrine was discovered when an exploratory sounding was made below the temple floor. Removal of deep sand west of the temple revealed a Minoan building of monumental construction. During this season the formal foot survey of the Kommos area was completed. Preparation and survey: 22 May-12 July. Excavation: 24 June-24 August. 1980
(SEASON
5)
We worked selectively on the Hilltop in preparation for study and phased publication. Further stratified Middle Minoan levels were cleared on the Hillside and Early Minoan sherds appeared in a deep sounding there. In the sanctuary we found that Temple C rested on two earlier temples (A, B) and that the shrine found in 1979 was the central object of worship within Temple B. (For a general view of the site from the south, see Pl. 2.8.) A botanical study was undertaken in April by Jennifer M. Shay and C. Thomas Shay. Excavation: 23 June-27 August. 1981
(SEASON
6)
Sandclearing in the Southern Area was finally completed, with over 100,000 m 3 deposited on the shore. On the Hillside, erosion and later overbuilding made difficult our effort to expose an entire Middle Minoan building. Further work to the south exposed Altar U in front of Temple B. We also discovered that Minoan Building J was built next to another (T) with an unusually long facade of orthostates. The east-west road in front of Building T renewed speculation that the town of Kommos had indeed served as the epineion (harbor area of an inland center) of palatial Phaistos. Excavation: 20 June-14 August. 1982
(SEASON
7)
We carried out a major study of the artifacts and architecture already recovered. Of the two trenches excavated at the end of the season, one exposed the southern border of the Middle Minoan houses on the Hillside and the other revealed the eastern extension of the Minoan road as it emerged northeast of Temple C. The road implied the substantial nature of the Minoan remains still to be found east of the Greek temples. Major consolidation work was carried out in the area of the temples, in effect leaving them upon a massive consolidated platform overlooking the Minoan buildings at the lower level. Study season: 1 January-15 August (Harriet Blitzer, L. Vance Watrous); April-May (study of botanical remains by Jennifer M. and C. Thomas Shay); 15 July-15 September (remaining personnel). Excavation: 9-28 August.
Synopsis of Excavation 1983
(SEASON
21
8)
A season (20 June-16 August) during which actual excavation was limited to forty working days because of restrictions placed upon the permit by the Ephorate. Nevertheless, we made progress in determining part of the plan of Building T, despite the very deep, stratified layering above it. A series of trenches traced its northern wall at least 50 m to the east. Deep trenches excavated down to T's floor revealed empty rooms with some LM I pottery and, in the area southeast of the later Greek temples, a long colonnade belonging to a stoa was discovered. The stoa faced south upon a very large paved court. For the first time we could begin to trace over a large area at least four differing phases of use during LM I—III. Study season: Winter visits by Joseph W. Shaw. Excavation: 20 June-30 July, with study through 16 August. 1984
(SEASON
9)
A long L-shaped trench was laid in, with the longest leg being south of and parallel to the northern facade of Building T, the shortest leg extending north to cross over the east-west avenue north of T and continuing beyond in order to overlap with houses of the Minoan settlement. The trench was subdivided into contiguous sections, each with its own excavation crew. On the northeast we exposed a finely preserved section of facade and road, south of a well-appointed Late Minoan house (X) with valuable stratified levels. To the southeast was discovered the northeastern corner of Building T. Within that building we explored, in the longer portion of the trench, parts of T's rooms and halls. Superposed upon a portion of T was discovered an enormous east-west wall of LM III date, belonging to a building dubbed P. Study Season: May (Harriet Blitzer, L. Vance Watrous, Jennifer M. Shay, and C. Thomas Shay); 16 June-26 August (remaining personnel, with major emphasis on the photography and conservation of the Middle Minoan pottery). Excavation: 16 June-13 August. 1985
(SEASON
10)
Exploration to the east and south, below over 10 m of sand accumulation and often through 3 m of stratified Greek and Minoan remains, exposed the massive northern ashlar wall of Building P as it disappears eastward into the scarp bordering the property. To the west portions of four more walls of P, parallel to that on the north, were found to have constituted the side walls of galleries completely open on the west to the earlier LM I court. Five clear periods of LMI—III were now separable in this area. Set above the ends of two of the galleries of P, and extending almost forty m toward the sea, was discovered an unusually long Archaic Greek building (Q), just south of the Greek Sanctuary. Building Q was probably connected with storage and commerce. At the end of the season the excavation permit was returned to the American School of
22
Kommos Site
Classical Studies for reassignment and the formal study-writing period for Volumes I to IV, already well under way for some of them, began. Study Season: 13 June-25 August. Excavation: 24 June-14 August. 1986
(SEASON
11)
Season 11 was essentially a study season devoted to the examination of excavated remains, with priority being given to the study of the Greek pottery (the study of the Minoan pottery having been largely completed during 1985). On the site we removed an unsightly section of the fourth-century Greek retaining wall, as well as the Greek/Minoan levels below it. We also initiated a major program of supporting scarps by building walls next to them or by covering scarps with a thick layer of a cement/earth mixture. Modern retaining walls were masked by using the same effective method. Study Season: May (James Wright); 14 June-30 August (remaining personnel). Limited excavation: 8-21 August. 1987
(SEASON
12)
This study season was devoted to cataloguing, photography, and mending of Greek and Minoan pottery, with special attention to stone tools, plasters, figurines, and architectural fragments. Site work was restricted to the yearly cleanup followed by the completion of all major consolidation work of scarps and modern as well as ancient walls. Study Season: 12 June-20 August. 1988
(SEASON
13)
Short visits for study were made by Harriet Blitzer, L. Vance Watrous, Maria C. Shaw, and Joseph W. Shaw during the summer months. The two volumes on Minoan pottery (II, III) were submitted to Princeton University Press, while preparation for Volumes I and IV continued in Toronto and elsewhere. 1989
(SEASON
14)
Work in Pitsidia concentrated on preparation for publication of the Greek levels, by Joseph W. Shaw, Maria C. Shaw, Peter J. Callaghan, and Alan Johnston as well as Eric Csapo and Patricia Bikai. Also, partial renovation of the storeroom and courtyard in Pitsidia was begun by Rebecca Duclos, in preparation for further excavation seasons. Final photography for Volume IV (The Greek Sanctuary) was carried out by Alexander C. Shaw and Robin A. Shaw. A long-term lease of the main storeroom was arranged. On the site a new area to be purchased was laid out, extending that bought by another 5,000 m 2 . Study season: 16 JuIy-I September.
Site Logistics 1990
(SEASON
23 15)
In Pitsidia, the renovation of the storerooms was completed (Rebecca Duclos). Study of the Greek material continued (Joseph W. Shaw, Maria C. Shaw, Guiliana Bianco, Peter J. CaIlaghan, and Alan Johnston). During the winter the final portion of the land on the south was acquired, so that during the summer the area was fenced, although not without difficulty, since portions of the land were claimed by others. At the same time a road, built by the community over the purchased land and leading down to the shore, was replaced outside the fenced area. A portion of the newly purchased land was cleared of sand down to the Greek levels, in preparation for 1991, when excavation would take place. In order to reveal the remainder of the Greek Sanctuary on the northeast, superficial trenches (72A series) cleared the extensions of Buildings V and F that had been partially exposed in earlier seasons. Study season: 14 June-31 July.
4. Site Logistics When excavation at Kommos was first being discussed, we planned to build an excavation house near the site in the southeastern corner of the area first expropriated and considered bringing in electricity from nearby Pitsidia. Water, we thought, could be obtained in a similar way or simply by digging a well and equipping it with a pump. The house was to be built, for the sake of security and comfort, around a court and the relevant plans were prepared by Giuliana Bianco and Angelos Koutroubakis. Various concerns, however, prompted us to abandon the idea in favor of establishing our base in Pitsidia, a few kilometers away. The major concern was that of security, for, given the types of vagrant campers, usually foreigners, along the shore throughout the year and without a cadre of permanent guards, neither records nor antiquities could be protected. There was also the expense of building such a structure when, during the first year, the site had not been properly evaluated archaeologically. Moreover, what would happen if we ever wanted to excavate near such a structure constructed, of necessity, over deep sand? (Later, we were to find Altars C and H as well as parts of monumental Minoan Building T below where the western wing of the house would have been built.) Consequently, we established our base in Pitsidia. Our storeroom/workroom began as two rooms of an unfinished house (Pl. 2.24; the owner had died while it was being constructed), which we eventually completed (PIs. 2.25, 2.26). The courtyard, a pleasant, airy place open on two sides, was used for washing the pottery, drying it on the cement floor, and then studying it and other remains on collapsible tables. Common meals (breakfast and dinner) were served in the court of a private house across the street, first at one house rented only for the summer and then (1981) in another which was rented for the year and became a kitchen with a storeroom and a single bedroom that eventually became another storeroom. During
24
Kommos Site
these summers the personnel lived in various rooms scattered throughout Pitsidia village, although the Spinthakis house on the village square (first with three rooms, then with five) was rented every year through 1994. An excavation day began, therefore, in Pitsidia, followed by an early drive after breakfast in the large International Harvester Travelall vehicle to near the site, for the excavation could be reached directly only on foot or by tractor because of the deep sand. Perhaps a road could have been built down to the site, but the townspeople generally did not want to divert tourists away from Matala where many Pitsidians worked at or owned restaurants and gift shops. Usually we would park, therefore, at the church of Ayios Pandeleimon on Vigles and walk in the dawn north, down the path on the steep hillside, until we reached the shore. Passing the only house built near the shore, rented by us on a long-term contract for the duration of the excavation as a place to store our excavation tools (Pl. 2.28), we proceeded to the site. There we would join the workmen, who would bring down from the small guardhouse in the northeastern part of the site (constructed in the spring of 1977 at the request of the Antiquities Service) the various photographic and architectural equipment we stored there during the excavation season. The site was permanently fenced after the season of 1982 and the perimeter fence was extended as new land purchases were made. Because of the distance between workrooms/storeroom and site that resulted, there tended to be both "site" and "workroom" staff. Such artificial separations were partially avoided when the latter group would come in the afternoon, to inspect the trenches and swim, and on Saturdays when we did not excavate but, rather, held informal seminars or had discussions in Pitsidia (Pl. 2.26) or simply caught up with study or various recording and organizational chores, such as cataloguing. There were, also, trips to archaeological sites and other excursions on weekends.
5. Site Recording On the site, trenches during the first season were normally 5 m 2 , although after that point the dimensions of new trenches would depend upon our estimate of the form of the underlying architecture and other stratigraphy and the amount of time available to complete the trench during a given season. Each trench received a number corresponding to the notebook being used for it (e.g., Trench 22A, in Notebook 22, the "A" designating the first trench in the series). Trenches excavated nearby and recorded in the same notebook would recieve an Arabic numeral (e.g., 22Al); if some distance away, a new alphabetic numeration would be given (e.g., 22B). Plastic pails were used for pottery and other small objects recovered from a particular context. Scarp tags and pail and object labels were usually plastic garden labels or wooden tags that would be written on with black markers. At the end of the season trenchmasters were required to write complete, synthetic, illustrated evaluations of both the process and results of the excavation of their respective trenches. Similarly, the various per-
Recording and Cataloguing
25
sonnel in Pitsidia would write summary reports, especially the ceramic experts, who would attempt to evaluate the style and dating of specific, unusual pottery deposits or groups. Topographical control of the site was established before excavation began, when the site survey and grid were established. As we excavated, Giuliana Bianco then gradually transferred the gridpoints from the stakes driven into the sand to permanent iron markers set into cement on or into ancient walls. A permanent record of the original appearance of the area was guaranteed by numerous photographs as well as by the topographical plans, but also by means of a survey carried out, in the spring of 1976, by Dr. and Mrs. J. Wilson Myers, sponsored by the Julian Whittlesey Foundation, using a camera suspended from a balloon (see PIs. 2.9-2.11 and Whittlesey et al. 1977). Site photography was normally carried out by the excavation director, using two Nikon F2s, one for color and the other for black and white photography. He also used, as back-up cameras, a Leica (M2) and a I1Ii" Mamiya, which would usually give the best resolution for large site photographs. Site photographs (contacts) would normally be glued into notebooks shortly after they were taken. Especially useful was a Polaroid camera, which produced black and white photographs on the spot; these could be labeled and glued in the same day.
6. Recording and Cataloguing in Pitsidia Upon their return from the field, the weary trenchmasters registered their pails of pottery and other finds on separate forms. Later, they would receive a special analysis of each pail by one of the ceramicists, including a list of catalogued pottery. The cataloguing took place in Pitsidia, although a number of objects (usually, architectural fragments) still in situ were catalogued on the site. The system was designed so that objects could be catalogued according to material (we originally did not expect Iron Age remains); there was also a "Miscellany" category. A sample entry, to explain the system briefly, might be C 45 K 76A 10A/3:35, to be read as follows: C (clay, including terra cotta) object number 45, found (or catalogued) at Kommos in 1976 in area A, within Trench 1OA (the first trench of the series in Notebook 10), in Level 3 and recovered in pail (we used plastic pails, color-coded for each trenchmaster) number 35 (in that trench or, as later preferred, in that notebook). The "A" sign for area, after the date, was devised at a time when we thought that we would be excavating on Vigles ("Area B") as well. Normally there would be a chief cataloguer and assistant cataloguer during the excavation season. From 1983 on, we made use of a computer in Toronto, and by 1985 our entire catalogue list of fifteen thousand items had been entered, so that printouts were available in Crete, sorted by means of dBASE III according to type of material or object, trench, level, etc. The official excavation photographer, using sets of 35-mm cameras, was responsible for all object photographs (except during 1976 when the director did both field and object work). Simple facilities for developing and contact printing were provided in one of the storerooms. Sherd materials and other antiquities of minor interest were stored over the winter in the
26
Kommos Site
Pitsidia storeroom. Objects of any commercial value were sent shortly after discovery for safekeeping to the Herakleion Museum. All reports and notebooks, as well as the interviews with the townspeople, were copied. Beginning in 1977, multiple microfiche copies of all re ports and notebooks were made for distribution to the University of Toronto Archives (not the Department of Fine Art, where most of the original records were kept) and the American School of Classical Studies at Athens. Copies were also loaned for research purposes to the various contributors to the Kommos volumes. A modest library of crucial monographs, off prints, and xeroxes of articles was maintained in Pitsidia and augmented as we discovered new contexts on the site.
A p p e n d i x 2.1
Diary Entry Recounting the Exploration of the Kommos Area by Sir Arthur Evans The following (see also Section 1 above) is an excerpt from one of the diaries of Sir Arthur Evans, the one labeled "Evans, Knossos, 1924, 1925" in the library of the Ashmolean Mu seum at Oxford. The text below is part of the description of his long trip, reported in A. Evans 1928, which began from Knossos on Sunday, 22 June 1924.1 am grateful to the authori ties at the Ashmolean Museum for the courtesies extended to me when I visited the museum in search of the record in April 1977. Much of Evans's writing, in very faint pencil script, often illegible, is very difficult to read; thus there are gaps in the transcription below, where the brackets contain additions and/or clarifications. I would also like to thank the late Dame Joan Evans and Professor and Mrs. Peter Warren, as well as my wife, Maria, for their help in the "decipherment" of the difficult text, of which a sample page is illustrated on Pl. 2.1. Pitsidia. Camp in olive grove beyond Πηγάδια [the village spring]. Sea by "Μόλο" [?] about 35 mins. away. Very sandy. Ponente [west wind] blowing high sea, but cliffs to left effectual break against Νότιος [south wind]. In cove[?] well and stone drinking troughs for cattle. Somewhat brackish water as on edge of beach. To N[orth]—a knoll beyond which little stream of Pitsidia reaches sea. Now only run ning winter[?] for month but was probably more of a river in Minoan times and must provide good wells still. This knoll forms the boundary of a valley now buried in sand but formerly evidently central quarter of a large Minoan town. Pottery and other remains extend on both sides of it and up the flanks of W[est] headland to S[outh]. On that side, especially on upper terrace[?] abundance of EM and MM I sherds. On the North hill mainly LM I and some MM III (in vineyard above said to be tombs). On topmost spur of S[outh] headland, above little church of St. Pandeleemon a strong
which was clearly a square EM ossuary 4.50 Meters each way and walls
Notes
27
about 0.80cms. Just b e l o w it a p a r t of a r o w of larger blocks b e l o n g i n g to a primitive tholos. R u n n i n g a l o n g N j o r t h ] flank of plateau t o w a r d s this s o u t h h e a d l a n d very e v i d e n t r e m a i n s of a m o u n d a n d s u p p o r t i n g blocks of old r o a d w h i c h could b e traced s o m e w a y in Pitsidia direction. The e n d of t h e great South Road! a n d to great M i n o a n h a v e n o n t h e Libyan Sea! The w a t e r rapidly d e e p e n s so t h a t it m a y still h a v e b e e n a c o n v e n i e n t p o r t w h e n t h e l a n d w a s h i g h e r (?also m o u t h of Pitsidia brook.) O n S[outh] akropolis n o t e d w i t h M a n o l a k i a great a b u n d a n c e of EM s h e r d s — s o m e of EM I Pyrgos d a t e . A galopetra f o u n d o n Njorth] side h a s i m p r e s s i o n , a n amyg daloid of M M IHB t y p e w i t h t w o sprays w i t h i n curving tines[?] of r o u g h fabric. O n N [ o r t h ] hill M. n o t e d from M M Ia to LM. O n slope of H . P . o n e g e o m . piece a n d out b e l o w N [ o r t h ] slope t h e s t o n e s n e a r D form s e g m e n t of circle. N e e d for excav. T h e w i n d G h a r b i s El G h a r b "Libicon". Reefs of talc[?] rock. Rockly islet [ = Papadoplaka] a n d Paximadi b e y o n d it like g u a r d i a n sphinx. Walked to Svakoryako across very s a n d y valley b e y o n d t h a t of Pitsidia s t r e a m . This w a s a t u r g i d w a t e r c o u r s e w i t h w a t e r still a mile from sea. It n o w cuts b e l o w t h r o u g h d e e p s a n d layers. All this valley by sea Π α χ ε ι ά άμμος. S p h a k o r y a k o o n h e i g h t w[est] of this c o m m a n d i n g r o a d to n o r t h a n d Dibaki. V2 h [ o u r ] back to Pitsidia. Re m a i n s begin a little below P. as I f o u n d frag, of a LM I larnax w i t h Σ ί μ α [sign?]. At Τελωνειον r o w s of pithoi a c c o r d i n g ] to native p r o p r i e t o r ] . We saw a b u n d a n t frags. LM I. H e called it the Τ[ελωνεϊον] s o m e still in place a n d walls in every direc tion. Later M i n o a n center.
Notes 1. The toponym "Kommos," which conforms to the local pronunciation, as first pointed out to us by Stylianos Alexiou, is of masculine rather than neuter gender (the latter, Evans's use). The etymology of the name is uncertain, but it is cer tainly connected with the modern Greek verb κόπτω (cut). The derivative Κομμός may refer to a number of geological features. One is the tow ering, vertical cliff to the south. Another possiblity are the waves striking against the cliff or even upon the beach. Thirdly, as a local tradi tion has it, the name may refer to the small pieces of eel grass, or kommos, that tend to pile up on the beach after a storm, especially in the corner to the southwest where beach and high cliffs meet.
2. Chadwick 1976: 53. Later, after the destruc tion of Knossos, much smaller administrative units may have been established. For a recent review of the possibilities see J. W. Shaw and M. C. Shaw, in press. 3. One recollection of Evans's visit, not neces sarily accurate, is that he assembled a good deal of material for excavation but then, shortly after a visit by Federico Halbherr, he was taken ill and the English contingent returned to Knossos. There is no mention of this incident in Evans's diary. Halbherr, some sixty-seven years old at the time, was to die six years later in Rome. 4. The only possible remains of such a built way in this area were inspected by myself and L. Vance Watrous during the summer of 1975,
28 when we hiked from Amnisos to the Mesara along the route taken by Evans, in the same valley and to the east of the route at present taken by the modern asphalt highway. Although any other significant remains of the Minoan road had been bulldozed away during the recent road-clearing operations, along the approach into Ayios Thomas we found remains (also illustrated by Evans) of cyclopean walls supporting and bordering a narrow way clearly made for beasts of burden. Nearby, unnoticed by Evans, was a Minoan dwelling, to judge from the sherds recovered not far from the "road." Following Evans's logic, the remains should be attributed to a Minoan "station" guarding the road. In my view, however, the presence of Minoan remains along a natural route for a roadway simply indicates that the Minoans did live here and that they used a nearby road. This route, moreover, was not necessarily the only approach into the Mesara from the north. Interconnecting roads between towns must have existed then as they do now and until clear archaeological evidence, usually available only after careful excavation, shows that a given pathway bordered or supported by cyclopean masonry was actually constructed during the Minoan period, its date should remain unspecified. 5. Dr. Cyrus Gordon was considering excavation after the Second World War, but the Kommos area was so heavily mined that the project was abandoned. Upon the defeat of the Axis powers and their withdrawal from Crete in 1944, much of their explosive material was left behind. As a result, there were many civilian accidents, most occurring when the explosives (hand grenades, artillery shells, mines, etc.) were tampered with in order to remove the explosive charges. Such charges were then readily sold for various purposes, especially for fishing. At Kommos, at least two people were killed while removing the charge, before the job of minesweeping by the military agencies began. Few partially complete mines (e.g., that in Pl. 2.7) were discovered during the course of digging (some discoveries are not welcome, even to archaeologists!), but pieces of them, rusted and harmless, could often be seen in various spots and can still be seen along the shoreline to the north of the site. 6. Indeed I doubt, on the basis of personal inspection of shoreside constructions throughout
Kommos Site the Aegean, that special harbor works such as elaborate moles and docks were built out into the water in the Aegean area before the Greek Archaic period (J. W. Shaw 1990). 7. Others who have commented on the site, generally echoing Evans's interpretation, were Glotz 1923: 402-403; Guarducci 1935: 239; Lorimer 1950: 59, n. 5 and p. 93-94; Matt et al. 1958: 107; Cipriani 1961-62: 133-34; Willetts 1962: 33, 134-35; Spanakis 1965: 261, 293; and Branigan 1970a: 192-93, 202, 207; 1970b: 7, 132, 171. Only after some time has passed will general and/or selective evaluations by others of our own work emerge, but see Spanakis 1983: 331-34; Platon 1982, 1: 290, 408, 410; 2: 88; and Treuil et al. 1989: 293, 543, 548. 8. During this same period, I was commissioned to write a description of ancient harbor works, published later as "Ancient Greek and Roman Harbourworks," in Bass 1972: 88-112. My first visit to Kommos came about partially as a result of the research necessary for that project. 9. Other early visits (October 1968, March and May 1970, and September 1971), when I was often accompanied by archaeologist Maria C. Shaw, included extensive walks in the countryside from Matala north to the Geropotamos River. These confirmed in my mind that Kommos was the only significant settlement of prehistoric times along this stretch of barren but beautiful coastline. 10. The land expropriated then, a rectangle oriented north-south, was originally subdivided, before 1900, into three parallel, contiguous properties oriented east-west, all of which extended at least one hundred meters further to the east. The northern and southern sections extended further north and south, respectively, than our present property lines. The northern section was originally owned by Polychronis Fasoulakis. The middle section belonged to Konstantinos Ioannou Fasoulakis, while that on the south belonged to Emmanouel Stylianou Kadianakis. The second landowner, Konstantinos Fasoulakis, also called "Chalkias" (ironworker or smith, as he actually was) to distinguish him from a relative of the same name, was apparently the man whose tale of lines of pithoi was recorded by Evans. According to some of his descendants in Pitsidia, the various ancient objects which he found in his fields (including a
Notes stone head[?]) were given to the Herakleion Archaeological Museum or to passers-by. The well on the property was excavated after 1957 by a Mr. Askoxylakis of the town of Vorroi; he and, later, a teacher in the Pitsidia school planted most of the almirikia (tamarisk) trees presently visible on the property. Aristotle Fasoulakis, born in 1905 and for many years our chief mason on the Kommos site, was the grandson of Ioannis Polychronis Fasoulakis. Aristotle told me that his grandfather found a many-handled, presumably Minoan pithos in the only gully north of the southern hill ("Vigles"), outside the expropriated area, where the present deep layer of sand has been eroded in places. The location of the pithos is presently unknown. For more information about the local inhabitants' involvement with the site, see M. C. Shaw 1980-81. 11. Other topographical plans, done at scales of 1:500 and 1:1000, were then prepared by an Athenian topographer, Professor John Bandekas. Land on the northern hill (9,254 m 2 or 9.254 stremmata) along the shore was thus finally acquired in May 1975, at our expense, for the Greek public domain for archaeological purposes. A second, later purchase, finally com-
29 pleted in 1987, consisted of 2,000 m 2 (2 stremmata) and was a direct extension of our southern property line. This expropriation made possible excavation south of Building Al of the Greek Sanctuary, a project we were able to begin in 1979 (rather than substantially later) through the generosity of the owner of the land, Mr. Emmanuel Daskalakis. The further, final purchase of 5,000 m 2 (5 stremmata) from the same landowner made possible further exploration of the southern and eastern extensions of Minoan Buildings T and P. This purchase, in two sections, was completed in 1989 and 1990. 12. From the point of view of strictly local terminology, however, it should be pointed out that, when we began excavation, the area of "Kommos" was confined in local parlance to the corner where the north-south shoreline meets the northern cliffs of Nisos. 13. George died recently, but before his death he specified that the archaeologists working at Kommos should be allowed to continue to use his shoreside house. To him, to his wife, Georgia, his daughters, Elsa and Pagona, and to his son, Nikos, we are much indebted for their consideration.
C H A P T E R
3
The Physical Geology of the Western Mesara and Kommos John A. Gifford, with David S. Reese
1. Introduction 2. The Kommos Survey Area 3. The Site of Kommos 4. Holocene Sea-level Changes 5. Inorganic Raw Materials at Kommos 6. Inorganic Processed Materials 7. Acknowledgments Appendix 3.1. Grain-size Analysis and Distributions Appendix 3.2. The Invertebrate Fossils (David S. Reese)
1. Introduction In fundamental ways that have affected human activities, the physical environment of the Kommos region has changed over the past four thousand years. Field observations supporting this assertion are abundant in the western Mesara, but ambiguous interpretation of the physical record can arise from the difficulties in establishing a chronological framework for reconstructed environments and in determining what combination of natural or human agencies may have affected them. All of the data in this chapter concern the geological aspects of the physical environment; specifically, they are the result of five seasons of field studies at the excavation site and in the region of the Kommos archaeological survey between 1976 and 1983. Topics that naturally presented themselves for investigation include, at the regional level, the climate and geology of the area around Kommos, its geomorphic elements, its recent sedimentary deposits, and the processes controlling their evolution. At Kommos itself, the nature of the sedimentary 30
Introduction
31
matrix burying the site was the major focus of research, though identification of the inorganic raw materials that were excavated was a useful ancillary task. These themes are sufficiently distinct to warrant separate sections in this chapter, which first will deal with the region and then the site. As soils are covered in a separate chapter by Michael Parsons (Chap. 6), their only consideration here is in the context of paleoenvironments and their identification at Kommos itself.
Climate Crete's position in the central Mediterranean endows it with a typical Csa climate, according to the Koeppen classification (summarized in Butzer 1971: 57): mild, moist winters (coldest months 10-12 0 C) and warm, dry summers (hottest months 26-28°C), with extremely rare snow or frost at sea level. Vegetation growth is possible all year, though lack of rainfall will limit it during the summer drought. The marked seasonality of precipitation is also characteristic, as is its extreme variability on an annual or decadal basis (documented in Semple 1931: 92). Table 3.1 excerpts the rainfall records of two stations near Kommos, Mires (for 1968-69) and Tymbaki (for 1966-70), in millimeters/month. In comparison with the rest of the island, these are average values; the range is from 390 mm of annual precipitation at Ano Arkhanes to 816 mm at Zaros. The fundamental relationship between rainfall and erosion is well known: sediment yields (as slope wash) reach a maximum at about 250 to 350 mm of mean annual precipitation. Yields fall off on both sides of this maximum, on the low side due to a deficiency of runoff and on the high side due to increased density of vegetation (Langbein and Schumm 1958: 1076). The western Mesara does not therefore suffer the theoretical maximum of erosion for semi-arid climates, other factors being equal.
Regional Geology of the Western Mesara While the western Mesara Plain has been mapped at a scale of 1:50,000 (Bonneau et al. 1984), no detailed study of the region's geology has been undertaken and none was attempted as
Table 3.1. Rainfall records of two stations nearest Kommos, Mires (for 1968-69) and Tymbaki (for 1966-70), in millimeters/month (Platakis 1971: 55). SEP
OCT
NOV
DEC
JAN
FEB
MAR
APR
MAY
JUN
Mires
0
50
115
155
75
25
50
25
13
0
Tymbaki
1
44
77
132
78
43
68
13
9
1
JUL
0 0
AUG
AVERAGE
0
508
6
472
32
Physical Geology
part of this research project. The following comments concerning the pre-Neogene rocks are only based on casual field observations and interpretation of the 1:50,000 geological map of Greece, Tymbakion sheet (Bonneau et al. 1984) in light of Bonneau et al.'s earlier summary (1977). Plate 3.1 is a generalized version of the southwest quadrant of the Tymbakion geological map, omitting topographic and minor structural details. In the northwest of the mapped region most of the Psiloritis Massif (the highest peak of which, Mt. Ida, is the island's maximum elevation) represents the base of the pre-Neogene stratigraphic section, the Para-Autochthonous (Plattenkalk) Series. Thrust over it by the Alpine Orogeny is the Allochthonous Series, a sequence of major and minor nappes, the rocks of which originally were deposited to the north of Crete. The series begins in the western Mesara with the Late Triassic-Middle Eocene Tripolitza nappe of shallow marine carbonate rocks. At its base are limestones and dolomites forming the southeastern foothills of the Psiloritis Massif. These are in turn overlain by black, microfossiliferous limestones and dolomites that constitute the lower peaks of the eastern Ida Massif (e.g., Alikadam, three km north of the village of Vorisia). Of even more limited extent are the youngest rocks of the Tripolitza nappe, a neritic limestone and a flysch of Paleogene age; they form the bedrock between Aghia Varvara and Chryso and also extend west of the latter village. The stratigraphic sequence continues with the next major rock sequence of the Allochthonous Series, the Pindos-Ethia nappe. Dark grey, well-bedded limestones of this zone form the cores of the Kedros mountain range, west of the massif, and the Asterousia (Kophinas) Mountains, which form the south-central coast of the island. Limestones of the Pindos-Ethia nappe were deposited in late Paleocene (Senonian) to early Eocene times, ca. 75 to 55 million years ago. The rocks were thrust southward during the Alpine Orogeny, commonly into basins filled with detrital (siliciclastic) marine sediments that would later become flysch rocks capping each tectonic sequence. An attempt to define the stratigraphy of these largely undifferentiated flysch units was made by Hussin and Tee (1982). On the Tymbakion geological map four younger nappe sequences are represented by a very discontinuous mosaic of overthrust rock units only a few square kilometers in area, which are commonly grouped together as the Serpentinite-Amphibolite Association. They are (from oldest to youngest): (1) the Arvi nappe, consisting of pillow lava basalts mixed with marly, reddish limestones; (2) the Vatos schists, of metamorphosed pelites and greywackes; (3) the Asterousia nappe, including marbles, amphibolites, and orthogneiss; and (4) the Ophiolitic Complex, mafic rocks such as pyroxenites and microgabbros altered to some extent by serpentinization. The latter two nappes are of limited distribution in the Asterousia Mountains but occupy a relatively large area south of the villages of Lochria, Kamares, and Vorisia, along the northwest edge of the Mesara Plain. All these overthrust nappes were emplaced southward during the Alpine Orogeny into sedimentary basins that would become flysch units capping each tectonic sequence, exactly as is the case for the underlying PindosEthia nappe.
Introduction
33
By the beginning of the Miocene Epoch ca. 24 million years ago, regional tectonism declined and the primary factor controlling deposition became the base level of erosion, or the level of the Tethys Sea, the precursor of the Mediterranean. Sedimentary rocks deposited on Crete during all subsequent geological time may be extensively faulted and folded, but they have not been pushed horizontally for scores of kilometers as were the pre-Neogene nappe sequences. In what is today south-central Crete the Mediterranean alternately flooded and receded from the structural trough—the present Mesara Plain—separating the Psiloritis Massif from the east-west trending Asterousia Mountains to the south. An extreme marine regression occurred during the Messinian salinity crisis (Hsu et al. 1978). Sea-level oscillations of hundreds of meters were common, and from the alternating transgression and regression events of three to five million years ago were formed the fresh, brackish-water, and marine sedimentary rocks that dominate the western Mesara bedrock geology. Neogene limestones, marls, and marlstones occupy a large area northeast of Mires, through which the main road from Herakleion winds down to the alluvial Mesara Plain. These rocks probably underlie most of the plain; they rise some 30 to 40 m above its western end as faultbounded hills around which the Geropotamos River has incised its channel. Meulenkamp et al. (1977) have formally mapped the Neogene-age rocks in the Aghia Varvara-Ambelouzos area just east of Mires. These units are correlated, in Table 3.2, with the rock units in the area of Kommos 10 km to the west-southwest. Henceforth I will refer to the Tortonian-age rocks of the southwestern Mesara (including the Kommos-Matala area) as the Ambelouzos Formation and Messinian-age rocks as the Varvara Formation, though they may only represent the gypsum member of that unit. While this correlation appears probable, more extensive field and laboratory study of the rocks would be necessary before accepting it as certain. Meulenkamp et al.'s (1977) reconstruction of the central Mesara's geological history during the Late Miocene to Early Pliocene Epochs (ca. five million years ago) reveals a complex, small-scale sequence of sea-level fluctuations combined with differential subsidence and uplift, producing episodic erosion and a "patchwork" of sedimentary rock bodies. Major eastwest trending faults had, by that time, opened the trough that was to become the Mesara. In the Mediterranean region, the Mio-Pliocene boundary corresponds with a general transgression and predominance of marine deposition. The Neogene rocks of Crete generally provide excellent locales and substrates for settlement, as may be seen by noting the distribution of modern towns and villages on the 1:200,000 geological map (Creutzburg et al. 1977). In Minoan times also, areas of Miocene limestone bedrock particularly were favored, presumably because they are well-drained, possessed unlimited quantities of tabular building stone, and often afford commanding views over the surrounding lowlands. The Kommos survey notes many sites in the region (e.g., sites 64, 20, 36, 49, and 55), in addition to Kommos itself, that are located on the Varvara Formation.
34
Physical Geology
Table 3.2. N e o g e n e rocks of the Aghia V a r v a r a - A m b e l o u z o s area ( M e u l e n k a m p et al. 1977: 145-46, table 4) a n d provisional correlations with the 1:50,000 Tymbakion geological m a p , a simplified version of w h i c h a p p e a r s as Plate 3 . 1 . KOURTES FORMATION
Age: ?Zanclean (early Pliocene, ca. 5 m.y.a.) Description: white marly limestones and marls Correlative with: "Lower Pliocene" unit (PIi) VARVARA FORMATION
Age: Messinian (Upper Miocene, ca. 7-8 m.y.a.) Description: Ploutis (upper) Member: sands, clays, and siltstones with large authigenic selenite bodies ("gypsum reefs") Correlative with: "Upper Miocene-Messinian" unit (M 4 m/g) Description: Gypsum (lower) Member: bioclastic limestones and laminated homogeneous marl sequences with occasional evaporites (selenitic gypsum) Correlative with: "Upper Miocene-Messinian" units (M 4 m and M4k) AMBELOUZOS FORMATION
Age: Tortonian (basal Upper Miocene, ca. 8-9 m.y.a.) Description: irregular alterations of conglomerates, sands, and clays; fluvio-lacustrine, brackish, and shallow marine environments Correlative with: "Upper Miocene-Tortonian" unit (M3c)
When necessary, Minoan builders also took advantage of other less common rock types according to their physical properties. For example, the evaporites deposited within the Varvara marine marls during the Messinian salinity crisis commonly are of relatively pure, crystalline gypsum inclusions ("balatino gypsum"). Such deposits exist within a kilometer of Aghia Triada and were quarried as building stone for that site and for the nearby palace at Phaistos (Pernier and Banti 1951: 55). This and other central Cretan sources have been shown to have supplied gypsum not only for Minoan buildings on the island, but also for Bronze Age structures at Akrotiri on Thera and at Mycenae on the mainland (Gale et al. 1988). Certain of the Ambelouzos deposits include marine and brackish-water clay strata; such occurrences of secondary marine clays were exploited as sources of pottery clay from Minoan times up to the present day in the Ierapetra region (Gifford and Myer 1984) and probably also in the western Mesara. During much of the Pliocene Epoch (5-1.7 million years ago), Crete continued to experience local uplift and subsidence, which, in combination with the oscillating level of the Mediterranean, allowed the deposition of marine sediments that now lie several hundred meters above sea level. These marine marlstones, named the Kourtes Formation by Meulenkamp et al. (1977), are particularly extensive east of Mires. Three million years ago, at the beginning of the Late Pliocene, general tectonic uplift over Crete produced a relatively uniform sedimentary record across the island, and upper Pliocene white marls are common across the
Introduction
35
island, especially in the Herakleion Nomos. The site of Knossos is built on these deposits (Roberts 1979). Subsequent regional faulting and folding during the Late Pliocene and Pleistocene epochs have further deformed these latest rock units and, in particular, have depressed a block of the Varvara Formation below its correlative strata to the north and south of the western Mesara. Phaistos and Aghia Triada are situated on one hill of this particular fault block—the palace actually is built on the crest of a major fault scarp—and similar bedrock hills may be seen across the Mesara Plain when looking north and south from Phaistos. Plate 3.1 illustrates how the Geropotamos River has eroded its channel through this fault block just south of the village of Vori. In the geological past it is possible that the course of the river was to the south of this block, past the location of the modern village of Hagios Ioannis. Overlying the relatively soft Pliocene limestones are the most recent extensive sedimentary rocks in south-central Crete: terrigenous conglomerates, sands, and clays deposited at the Pliocene-Pleistocene boundary some 1.7 million years ago. Such rocks extend from Mires northwestward to the villages of Lagolion and Kalochoraphitis, about 220 m above present sea level. On Crete, the last 1.7 million years of geological time, the Pleistocene and Holocene epochs of the Quaternary Period, generally are recorded as unconsolidated or poorly cemented sediments deposited in a wide range of environments, from mountain slopes to alluvial plains to beaches. During the Pleistocene Epoch, which is defined (on paleoclimatological grounds) to have ended some ten thousand years ago, upland Crete was subject to coldclimate weathering phenomena, and evidence of local glacial activity—moraines and block streams—has been described from the peak of Mt. Ida (Fabre and Marie 1982). More intense weathering of the lower slopes of the Ida Massif to the north of the Mesara Plain produced widespread deposits of colluvial slope breccias (coarse, angular rock fragments in a matrix of redeposited terra rossa-type soil sediment). They are mapped on the Tymbakion geological map as "Undivided Quaternary" and constitute the subsurface deposits of the Mesara Plain through which the Geropotamos River has incised its present channel. Post-Pleistocene alluvium, colluvium, and the soils developed thereon form the actual surface of the plain. While the geological age and modification of those sediments is insignificant as compared with the underlying several thousand meters of rock, nevertheless the changes that have occurred are substantial on a human scale.
Geomorphology of the Western Mesara At the macroenvironmental scale, the western Mesara includes (1) the intersection of the Geropotamos River flood plain with the Mesara Gulf coast and (2) the western foothills of the Asterousia Mountains. The mountain foothills are here composed of the previously de-
36
Physical Geology
scribed Varvara and Ambelouzos formations, which overlap the latest pre-Neogene tectonic unit (Vatos schists and gneisses) that crops out only a few kilometers east of Pitsidia. Both on the map and on the ground, the region's geology is reflected in the topography of the western Mesara Plain, with "islands" of Neogene bedrock rising up to 80 m above the surrounding Geropotamos River plain. As mentioned previously, the most distinctive of these is the ridge on which Phaistos was situated. South of the village of Metochi, the alluvial plain disappears and the landscape is an expression of the Neogene bedrock erosion surface. Primary geomorphological elements in this area include dissected upland plateaus, rectilinear slopes, and small stream valley bottoms. But the Mesara Plain as a geomorphological system constitutes a drainage basin that through recent geological time has been maintained by the river system occupying it, and thus the plain is a key to understanding all other land forms of the region. The present Geropotamos River is extremely underfit for the size of the Mesara Plain, as is true of many other rivers in arid and semi-arid climatic zones. Though no discharge figures were available, streamflow at the road crossing just north of Phaistos, as observed during June 1977, was estimated to average only a few cubic meters per second. (Some of the river water, however, is diverted for irrigation upstream of this point.) The Geropotamos channel north of Phaistos is 2 to 3 m wide and has incised itself less than 1 m deep into a silty floodplain alluvium of very recent age (less than a century). Downstream towards the Mediterranean shore, the lower reaches of the Geropotamos are incised more deeply (2 to 3 m) into cobble- to boulder-sized alluvium that represents reworking of Quaternary slope breccias. The town of Tymbaki and the village of Kokkinos Pyrgos ("red tower," referring to erosional remnants of the breccias), a few kilometers to its northwest, are built on the Pleistocene breccia, which here forms a good example of a pediment sloping gently up to the southern foothills of the Psiloritis Massif. At the mouth of the Geropotamos, its alluvium forms a cobble beach that has in places been cemented into a discontinuous terrace 1 m above present sea level. This very limited deposit is identified on the Tymbakion geological map as a "Holocene beach deposit." A similar deposit has been observed by the writer on the coast just east of Ierapetra; it indicates a relative sea-level change of unknown date. The presence of this continuous beach ridge, and the fact that the flood plain's elevation rises from ca. 2 m at the beach up to ca. 20 m at the foot of the bedrock ridge where Aghia Triada lies, constitute strong evidence that the shoreline never lay in the immediate vicinity of Aghia Triada during the late Holocene and, therefore, it could never have been a Minoan port (cf. Bintliff 1977a: 619). Plate 3.5, a portion of a LANDSAT photograph of south-central Crete, shows the very abrupt change in coastal morphology that occurs at the Matala headland from a low-angle unconsolidated shore zone to a fault-controlled rocky shore. Being open to deep water on the west and experiencing dominant winds from the northwest, west, and southwest, the Mesara shoreline is dynamic, continually readjusting itself to be in equilibrium with wave energy. No longshore bars are apparent from air photographs or from observation of break-
Kommos Survey Area
37
ing wave patterns; it appears that the near shore slopes off rather quickly into several meters of water. Recent bathymetric data are not available for Matala Bay just seaward of the Mesara Plain, and one must turn to the Royal Navy hydrographic charts of the mid-nineteenth century for some idea as to the bottom topography (Great Britain, Admiralty 1862). The five-fathom line parallels the sandy shoreline of the coastal plain at a distance of 400 to 450 m offshore. Southward past Kalamaki and Kommos the contour line becomes more irregular but maintains that approximate offset. The rock reef off Kommos (see below, "Papadoplaka") is noted on the 1862 chart as an area of less than 2 fathoms' depth with rocks awash. At the base of the cliff forming the north side of the Matala headland, water depths range from 1.5 to 4 fathoms, with rocks awash; there is a notation of "landing" there, presumably for a small boat. The near-shore zone southward from the Matala headland to Cape Lithinos is predictably much steeper, with depths greater than 10 fathoms to be found only 200 m from the shore. The bottom of Matala Bay itself rises evenly from 6 fathoms at its mouth up to 1 fathom within 50 m of the small pocket beach at its head (Blackman 1973: fig. 1). Alluvium from the Geropotamos is the source of most beach sediment to the north of its mouth and, thanks to the predominant southerly longshore transport, supplies all the coastal zone sediment from its mouth southward to Matala. Thus a continuous sand beach curves 5 km from the Geropotamos southward to the steep rocky shoreline a few hundred meters south of the Kommos site. Some transport of sand in the near-shore zone around the Matala headland does occur, as the bottom of Matala Bay and off Matala is covered by the same Geropotamos sand. The finer size fractions of the river's alluvium are deposited in deeper, lower-energy bottom environments further offshore. North of the Matala Headland three bedrock ridges extend westward toward the shore of the Mesara Gulf. All are composed of uplifted and eroded strata of the Varvara Formation. However, none of the three either affects the present shoreline morphology or forms a barrier to longshore sediment transport; they were eroded back by wave action from the geological past up to the present. Their topography is subdued except on free-rock slopes, and they act as divides (the natural topographic lines of separation) for the three small drainage basins that will be discussed in Section 2. Kommos is situated on the southernmost of the three ridges. Some 350 m west of Kommos an isolated bedrock outcrop rises up from the flat, sandcovered bottom in 7 to 8 m of water, forming a rock reef called "Papadoplaka." It is the only near-shore rock reef along the 5-km beach, and its evolution will be discussed in the section dealing with the Kommos site itself (Section 3).
2. The Kommos Survey Area The survey area constitutes some 20 km 2 of low hills on the southwest edge of the Mesara Plain, along the shore of the Libyan Sea (Pl. 3.2). The narrow, sandy shoreline of the Mesara Plain extends southward to about the middle of the survey area, where it abuts
38
Physical Geology
the Matala headland, a rock cliff coastline that continues along and beyond the southern bor der of the area.
Geology Bedrock in the survey area consists mainly of the Varvara and Ambelouzos formations, which form the extreme western foothills of the Asterousia (Kophinas) Mountains. The older and stratigraphically lower of the two units—the Ambelouzos—predominates south of a major east-northeast-west-southwest normal fault passing through Pitsidia. It consists of shallow marine and brackish-water marlstones, fine-grained white limestones, and iron-rich clayey siltstones. North of Pitsidia, the down-faulted block exhibits the marine marls and limestones of the Varvara Formation (Pl. 3.1). The limestone types include muddy biopelmicrites and biomicrosparites (using the classification terminology of Folk 1974: 164). Marine fossils (primar ily molluscs and echinoids) are common in those units (Appendix 3.2) as are infilled biogenic structures such as burrows. Calcareous sandstones containing poorly defined lenses of wellrounded pebble conglomerates also are common in the sequence as are well-indurated marlstones with diagenetic inclusions of authigenic sulfides. Small-scale penecontemporaneous deformation (recumbent folding) of marl beds within the limestone strata is common in the Varvara Formation, reflecting tectonic instability at the time of deposition in shallow Messinian-age basins. Because the rock type known as marl plays an important role in the archaeological inter pretation of the region's geology, some discussion of the term and its precise relationship with the term "clay" is necessary. In geological usage clay refers both to a size range of sedi ment grains (with diameters smaller than 4 μιη) and to several families of minerals (mostly hydrous aluminium silicates) that are the most common constituents of the clay size range. In nature, clay minerals are so commonly found mixed with particles of the non-clay min erals calcite and quartz—in both the clay size range and the coarse silt size range—that the resulting sediment has the common generic name marl. Thus an unconsolidated marl may be defined as a mud-carbonate mixture with a minimum of about 35% each of the end mem bers. The lithified equivalents may range from argillaceous limestones to pure calcareous limestones and these form the end members of the general rock type identified as marl (or "marlstones") on geological maps. In addition to the minerals of the clay and silt size fractions, marls commonly contain sand-size grains of other minerals, rocks, and shells of microfauna (microscopic organisms such as ostracods and foraminifera) that inhabited the water column above the site of marl deposition. Specific identification of those remains, through the techniques of stratigraphic micropaleontology, is commonly employed to estimate the geological age of the associated deposits (e.g., Meulenkamp 1969).
Kommos Survey Area
39
Marl units may show little cementation, especially if they have not been exposed to weathering. Such friable marls exist under the Kommos site. Grain-size analyses of typical friable marl samples are presented in Section 3 on site geology. These marls contain a large terrigenous component of clay minerals and would be termed "muddy biopelmicrites" in Folk's (1974: 168) carbonate rock classification. Two marl strata composed largely of clay minerals were discovered in outcrops of the Ambelouzos Formation. One stratum, near Kommos survey site 38 (Ayios Stephanos spring) forms the eroding foot of the hillock on which the site is located. X-ray diffraction analysis of a sample shows it to be a mixture of clay minerals of the smectite group (2:1 lattice structure) and the kaolinite group (1:1 lattice structure). A second, relatively pure clay stratum was discovered only 200 m south of the Kommos site, at the foot of Vigles Hill, and will be described in Section 5, in the part dealing with pottery clay. Other minor rock types present in the survey area include a calcarenitic reef limestone facies of the Varvara Formation (M 4 m on the Tymbakion geological map) capping the Matala headland, most of which is composed of Ambelouzos limestones. This remnant of the Varvara is very well cemented relative to the underlying Ambelouzos rocks and, in weathering less rapidly, it tends to form scarps, slope breaks, and knick points in drainage channels. The calcarenite caps most of the high ground south of the Pitsidia ridge, particularly southeast of Matala, where it crops out as small cliffs at elevations from 100 to 140 m (Pl. 3.3) and accounts for the extensive upland plateaus in the area. Along the eastern boundary of the survey area a third formation exists below surficial sediments: a flysch consisting of poorly cemented calcareous sandstones and conglomerates. Because these rocks are prone to weathering, the topography and soils east of the survey area are quite different. More exotic rock types than the marls, limestones, and sandstones mentioned above exist at two locations within the flysch (Pl. 3.1). In the middle reach of the northern tributary of the Matala stream, near site 18, a rounded bedrock dome (superficially resembling an inselberg) has been exposed by erosion of the overlying Ambelouzos marls. The dome is so steep and prone to mass wasting that it exhibits a bedrock surface completely devoid of sediment. The dome is a fragment of the Pindos-Ethia nappe mentioned previously. While that tectonic unit is composed primarily of siliciclastic (non-carbonate) marine sediments, it commonly contains large exotic rock fragments such as exemplified by this particular exposure, a phyllite that was detached from its original location and deposited as an intact "fragment" (weighing several million tons) within the Eocene-age marine sediments. Such displaced blocks are termed olistoliths (from olistoma plus lithos). Similar blocks were mapped in the Ayiofarango Gorge, 10 km southeast of Pitsidia (Doe and Holmes 1977). Site 18 of the Kommos survey is located adjacent to the phyllite olistolith, which also contains a subsidiary deposit of low-grade marble. The Roman builders of Metellum (Matala) made
40
Physical Geology
use of marble from this readily accessible outcrop for columns and other architectural elements. South of the Pitsidia Ridge, site 34 is similarly located on a less well-defined hillock of phyllite. Because the slope angle is much less steep than that at site 18, a soil has developed on this bedrock, as noted by Parsons (Chap. 6, Section 1, under "Details of Soil Mapping Units, B8, Lithic Xerorthents"). However, only two bedrock types—Neogene calcareous marlstones and limestones— dominate the survey area, and so little can be said about preferential settlement sites as a function of rock type. Outcrops of limestone tend to weather less rapidly than the marls and to form resistent shelves. Throughout the area tabular slabs of this rock, easily pried out along bedding planes, are a common building material both in the historic and prehistoric past. The more fossiliferous limestone facies of the Varvara Formation can be found in beds up to 1 m thick on top of the Matala headland, and this rock was used by both Minoan (at Kommos) and Classical (at site 5) builders for local ashlar building stone. A further consideration of this rock is given in Section 3. In terms of the survey area's geological structure, the general trend is of southeastwarddipping monoclines bounded by small fault scarps. This is true of many ridges (e.g., Asphendilias and Pitsidia) and other topographic elements such as the Matala headland. But large variations in strike and dip of bedding planes may occur within short distances and they have affected local drainage. The Pitsidia stream channel, in particular, has been constrained in its middle reach by several small bedrock knobs reflecting local deformation, possibly syndepositional folding. As is true of the whole island, the western Mesara exhibits a large number of geologically recent fault systems. The south edge of Pitsidia village is built against an eroded fault scarp that exposes marine and brackish-water sandstones and calcareous mudstones of the Ambelouzos Formation. Excellent specimens of fossilized echinoids have been collected from the fault scarps immediately south and southwest of the village (Marcopoulos-Diacantoni 1973). It is not possible to estimate from the degree of scarp erosion when the most recent movement occurred along this section of the fault. From Pitsidia, the fault continues southwestward, forming the steep west slope of the Pitsidia ridge and the northern coastal cliff of the Matala headland. A sub-parallel fault delimits the southern coast of the headland and so accounts for the structure and formation of Matala Bay itself. Besides those major faults, numerous secondary faults were noted in the survey region, trending at right angles (i.e., northwest-southeast) to the primary regional fault orientation. High-angle, nearly vertical normal faults are exposed in the Matala headland, which in addition to extensive fracturing is tilted westward (contrary to the regional trend) in a monocline that is well exposed in both the north and south cliffs of Matala Bay. No traces of Quaternary marine abrasion platforms or terraces were observed in the survey area, though they are common in other parts of Crete (e.g., Angelier and Gigout 1974).
Kommos Survey Area
41
Drainage Basins and Ground Water Within the survey area three small, east-west trending drainage basins are delineated by Neogene bedrock ridges: the Matala basin in the south, the Pitsidia basin in the center, and the Kalamaki basin in the north (Pl. 3.2). Only the Pitsidia basin lies entirely within the survey area. In addition there is a coastal area of approximately 3 km 2 —including Kommos, Vigles Hill, and the Matala headland—that has no organized drainage system. The eponymous streams occupying their respective basins are small and seasonal; they all drain westward into the Mesara Gulf. The Kalamaki and Matala streams, with headwaters several kilometers east of the survey area, may flow as intermittent streams in winter for periods of at least one month's duration, but the Pitsidia must be an ephemeral stream that flows only in direct and brief response to rainfall falling on its minute drainage basin (2.7 km 2 ). Each of the three streams drains a different kind of surficial rock type. Half of the Kalamaki stream drainage is within the Varvara Formation, north and west of Pitsidia; east of the Pitsidia fault the Kalamaki stream drains both the Ambelouzos and the Flysch formations as well as a fragment of metamorphic gneiss belonging to the Asterousia nappe. The Pitsidia stream wholly drains rocks of the Varvara Formation, but more than half of its basin is mantled by an aeolian sand blanket that inhibits surface runoff of precipitation. Some two to three times larger than the others, the Matala drainage basin also extends considerably farther east into the flysch of the Pindos-Ethia metamorphic zone. Even at the height of summer, moisture from subsurface flow is found only a few centimeters deep in the middle reaches of its channel alluvium. Also, alluvium of the Matala stream contains a wider range of rock types than does that of the Pitsidia or Kalamaki. Despite this, cobbles from its channel appear not to have been a major source of raw material for stone tools at Kommos (see Chap. 8). The lower valleys of all three streams in the survey area are partly to completely covered by a sand layer that represents recent aeolian deposition from the coast. Consideration of this sand layer is postponed until later in this section; here it is only noted that the unconsolidated sandy nature of the alluvial plain cover allows all three stream channels to shift freely within the confines of their lower valleys in response to peak discharge events. A 1944 air photograph of the Kalamaki valley alluvial plain shows that the stream channel ran along the northern foot of the Asphendilias Ridge before reaching the shore; this is approximately 200 m south of its 1979 position. In following the main channels of the Kalamaki and Pitsidia streams upvalley from their alluvial plains, one can see that all their tributaries are first order, meaning that they themselves lack smaller tributaries (as well as alluvial plains). Thus the Kalamaki and Pitsidia streams are second order, a reflection of their small drainage basin areas. However, the larger Matala stream is third order because its southern (second-order) tributary possesses one tiny first-order tributary of its own.
42
Physical Geology
All three streams exemplify the "small watershed," a drainage basin so small that its sensitivity to high intensity, short duration rainfalls and to harmful land-use practices is not buffered by any channel storage characteristics. Response time between the inception of hillslope erosion and major deposition on the alluvial plains of small watersheds may be measured in months or even weeks rather than the years or decades of higher-order streams and rivers. The present channels of all three streams are incised from 0.5 to 3.0 m into their alluvial deposits. Thus they all exhibit an erosional regime. In the Kalamaki stream valley this erosion phase began some time between two thousand years ago and the present, since archaeological remains dated to Hellenistic/Roman time are exposed in the channel bank at site 56 (see below). Whereas the Kalamaki stream is more or less unconstrained by bedrock topography at its mouth and the Pitsidia stream is only partially constrained by a low coastal ridge of bedrock (at the south end of which lies Kommos), the Matala stream funnels all of its drainage into the narrow, steep-sided alluvial plain forming the landward portion of Matala Bay. During rare flood events, shoreline structures of modern Matala and ancient Metellum are (and were) endangered, particularly toward the north side of the bay. It is notable that the Roman foundations exposed at site 72 are buried under a meter of sediment. After exceptional flash flooding in the Matala Valley, the resulting changes in the topography must resemble those described for Kato Zakros's alluvial plain after a flash flood witnessed by the archaeologist D. G. Hogarth in 1901: "[It] . . . swept away the whole plain on May 15th, and in two hours changed the face of the landscape, leaving stones and naked rock where fields, vineyards and groves had been, and carrying to the sea 4,000 trees" (Hogarth 1900-01: 123). Less catastrophic depositional events would occur in the other two stream valleys due to their smaller size as well as unconstrained discharge. Ground water is not plentiful in the Kommos region. The water table intersects the surface in the form of springs at only three localities within the survey area. All three have small flow rates; the one at Ayios Stephanos was measured by Parsons as 1.3 litres/minute in early June 1979. Pitsidia's spring is associated with the major fault scarp on which the village is sited. Its flow rate was measured as 5.7 litres/minute. For the entire western Mesara, there are no springs identified in the catalogue of Kourmoulis (1979) with flow rates of 600 litres/minute, the minimum for that inventory. In the recent past, as agriculture has expanded in the stream valleys, land owners have turned to hand-dug wells that tap perennial subsurface ground-water flow concentrated in the subsurface channel fills near the shoreline. All three alluvial plains contain such wells, which are 2 to 3 m in diameter and from 2 to 6 m deep, depending on their distance from the coast. Such a well was dug through sand and alluvium of the Pitsidia valley, 550 m from the shore and adjacent to the stream channel. It was lined with 2-m-diameter concrete reinforcements and, according to a local laborer who was involved in its excavation, penetrates 3 m of
Kommos Survey Area
43
sand and 2 m of soft rock (presumably marls of the Varvara Formation) before reaching the water table. Ground-water flow through strata of the Neogene sedimentary rocks in the survey area is most noticeable along the present coastline, in particular where bedrock ridges approach to within a few tens of meters of the Mediterranean. It is possible to dig only a meter or so down through the back-beach sand at the base of these shoreline cliffs in order to encounter interstitial fresh water. Beachrock formation apparently is related to these subsurface seeps, as will be considered in the Kommos site geology section (Section 3).
Slopes and Colluvium Slopes in the survey area and the sediments that have been removed from them are described in this section. The study of hillslope forms and processes of erosion contributes to an understanding of the recent geological history of a region. In the case of an archaeological project this study will indicate the "survival rate" of sites of different periods and illustrate the evolution of a particular site's topography. Evidence is abundant that a major slopewash (and therefore alluviation) event began in the survey area during the second millennium B.C. Plate 3.3 is a topographic profile of a typical slope on marl bedrock of the Ambelouzos Formation along a measured transect from the shore of Matala Bay southeastward to the topographic high called Matalokephala. Angles of average slope segments (as measured in the field) are also shown. The concavity of the slopes in these relatively weak, homogeneous rocks is apparent and is characteristic of slopes on fine-grained sedimentary rocks in a Mediterranean climate (Young 1972: 232-33). The upper slope segment of 14° is nearly bare of loose sediment and detritus; it equals a denudation slope. At the slope base there is a continuous accumulation of detritus that indicates a net gain of colluvium in this zone; it is thus termed the accumulation slope. Intermediate between these two slope zones is the transportation slope, where eroding sediment is constantly in downward motion (Young 1972: 23). A very sharp break in slope (from 24° to 2°) marks the transition to a basal concavity (here composed of sand) that characterizes slopes in semi-arid environments (Carson and Kirkby 1972: 340). Accumulation slopes have only moderately thick soil covers compared with alluvial bottom land, but some farmers may gamble on using them for brief periods to increase agricultural yields. Unfortunately such slopes are far too steep for cultivation without terracing. An attempt by Pitsidia farmers in 1979 to grow wheat on an accumulation slope segment is revealing of geomorphological processes; it demonstrates the rapidity with which slope stability may be deranged. In the vicinity of site 38 Parsons has mapped the soils as Dystric Xerochrepts (Pl. 6.1). Between this site and the Ayios Stephanos chapel, the author observed
44
Physical Geology
in 1979 a wheat field planted on a 28° slope, with gullying already beginning to erode the soil surface. In 1980 the field had been abandoned and reverted back to annual grasses. Most of the gentle slope surfaces in the survey area are covered with a mantle, or regolith, of weathered bedrock ranging in thickness from several centimeters to several meters. On low-angle slopes its downslope movement is slow and soils have sufficient time to develop within the regolith. On slopes of moderate angle in the Kommos survey area, the regolith is less stable. The weathering processes by which it evolves from the underlying bedrock, and its general erodability, are a function of (1) intensity and duration of precipitation, (2) slope angle and length, (3) vegetation type, (4) organic mat cover, (5) soil texture, and (6) parent material (Butzer 1974: 59). The general aspect of the regolith in the survey area is similar to that described for Melos by D. A. Davidson and Tasker (1982: 86). Parsons has identified the soil types on many of those slopes as Lithic Xerorthents or Lithic Calciorthids (Chap. 6). Weathering of bedrock is commonly discussed in terms of the processes of decomposition, or chemical weathering, and those of disintegration, which is mechanical in nature. Both processes constantly operate to break down rock into weathering products of smaller particles, different minerals, and soluble products. For the highly calcareous and relatively soft limestones and marlstones of the Kommos region, decomposition is the dominant process. Since many bedrock slopes there are moderately steep with little vegetation cover, decomposition is free to act on continuously exposed fresh surfaces of soluble calcareous bedrock. This produces finer-grained sediments composed of more stable clay minerals. As the weathering products aie subjected to water erosion, the resulting mass of partially decomposed rock weathering products and unweathered rock fragments does not have the opportunity to evolve into a mature soil until it accumulates at the base of a slope. While in (slow) motion downslope, this regolith also constitutes colluvium (from colluvies, collection of washings or dregs), on the surface of which various immature soil types may develop. Movement of colluvium as surface wash reaches maximum rates in semi-arid climates such as that of the western Mesara, thanks to the lack of a continuous vegetation cover on many slopes, the thinness or absence of a soil organic (A) horizon, and the short-duration, high-intensity nature of much of the precipitation (Young 1972: 67). Several processes can interact to magnify slope erosion: (1) Ploughing and over-grazing destroy the surficial organic mat and thereby decrease infiltration capacity, which increases the volume/velocity of runoff in heavy rainfall events; (2) as slopes lose the vegetation cover, water erosion by rainsplash, sheet, and rill erosion is accelerated; and (3) poor cultivation practices (see Chap. 6) promote sheet erosion on low-angle slopes and rill erosion on highangle slopes. In semi-arid climatic regions, cultivation of slopes steeper than about 10° without using terraces will contribute to all three of the above processes. The equilibrium of slopes and
Kommos Survey Area
45
vegetation cover is very delicate in such environments and slight disturbance of any one factor (rainfall, agricultural practices, grazing) can trigger a magnified erosional response (Carson and Kirkby 1972: 349). The phenomenon of accelerated soil erosion usually involves a disruption of local environments " . . . that may show little output beyond small watersheds" (Butzer 1974: 63). The implications of such downslope movement must be considered as well when evaluating the distribution of sites in the survey area. Two distinct kinds of hillslope erosion may occur depending on whether or not an accumulation slope segment exists. If it does, then the slope will erode by a gradual lessening of inclination. If sediment is removed from the base of the slope (as is the case, for instance, along the shore cliffs of the survey area), then the slope will retreat back at a constant inclination (Schumm 1966: 101). In the latter case there is no hope of finding redeposited sherds or other indications of an archaeological site's nature or extent. Aside from coastal cliff slopes and vertical scarps of eroding stream channels, the majority of slopes in the survey area appear to be eroding by decline rather than parallel retreat. But hillslope angle cannot have lessened very much, since two thousand to three thousand-yearold structures may still be found on some slopes. The site of Kommos itself represents a special case, as do also the Minoan ruins at site 20 near Kalamaki; both have been buried for at least the last two thousand years by windblown sand and thus protected from surface processes. Colluvial deposits are dynamic and the dating of material included in them places limits on their inception. The degree of maturity of any soils developed at their surface is a less precise, but still useful, indicator of their age. In the three stream valleys of the Kommos survey area, colluvial deposits at slope bases interpenetrate alluvial terraces and their boundaries are vague. A further complication in defining colluvial deposits stems from the extensive deposit of windblown sand that blankets all three valley bottoms and covers the lower portions of many slopes. In order to date the inception and duration of colluviation, for archaeological purposes, geologists depend heavily on discovering datable cultural material (commonly potsherds) within the deposit. Even though such a date can only represent a terminus post quern for the inception of colluviation, it is far more accurate than any geological dating technique and considerably less expensive. Commonly the extensive vertical sections of colluvium needed for discovering datable material are confined to the middle reaches of stream channels, upstream from alluvial plain deposits. Several such sections were observed along the Kalamaki and Matala stream channels and one was studied in the Pitsidia stream. A few meters downstream from site 56, at the inland end of the Kalamaki alluvial plain, marl bedrock is visible at the base of the channel scarp and the complete sediment section was sampled and recorded (Pl. 3.6, Table 3.3; cf. also Chap. 7, Pl. 7.23). Above the decomposed bedrock of Unit 1, Unit 2 is a light greyish brown, gravelly mud derived from erosion
Physical Geology
46 Table 3.3. G e n e r a l s t r a t i g r a p h y at K o m m o s survey site 56 (Kalamaki s t r e a m b a n k ) .
Unit 4 Unit 3 Unit 2 Unit 1 Bedrock (marl)
Description
Thickness
sandy loam gravelly, muddy sand colluvium gravelly mud colluvium decomposed bedrock
10-30 cm 100-150 cm 20-100 cm 30-50 cm
of regolith and typifying much of the colluvium in that valley. This exposure of Unit 2 was found to contain small, abraded potsherds from an LM I spiral cup and an LM coarse ware vessel of unknown type. The Hellenistic/Roman floor and wall foundations of site 56 shown on Plate 3.6 were cut into this same unit. It, therefore, was deposited sometime between the Late Minoan and the Roman periods. The overlying sand of Unit 3 is a layer of colluvium derived from erosion of aeolian sand from the bedrock slope 100 m to the east. Unit 4 (not marked on Pl. 3.6) is a recently developed soil at the modern surface. As part of the laboratory study of geological samples from the Kommos survey region, detailed grain-size analysis of 52 sediment samples was performed at the Archaeometry Lab oratory of the University of Minnesota, Duluth. (Details of the analytical procedures are de scribed in Gifford et al. 1982.) Plate 3.4 summarizes the grain-size analyses of three samples from survey sites 56 and 20 on the Kalamaki alluvial plain as cumulative weight percentage curves. In this type of graph, the cumulative weight of each sample's 32 size fractions, expressed as a percentage of its total weight, is plotted against size fractions on the graph's X-axis expressed in phi (φ) units (see Appendix 3.1). Thus the steeper the segment of the cumulative curve, the greater the quan tity of that size fraction in the sample. Well-sorted sediments are those with all grains con centrated in just a few size fractions and their cumulative curves thus are nearly vertical. Such sediments reflect an effective natural sorting or winnowing mechanism, such as wind or wave action, which has segregated those few grain fractions from a larger range available in the sediment's source area. In contrast, sediment samples exhibiting cumulative grain-size curves that approach a diagonal line across the graph (lower left to upper right) possess a much wider possible range of grain-size fractions; these are characterized as exhibiting poor sorting and representing deposition in natural environments of low energy, incapable of winnowing out the finer clay and silt-size fractions. With these points in mind, it is evident from Plate 3.4 that the Unit 3 colluvium from site 56 is relatively well-sorted, with most of its grains having diameters in the two to three φ size fraction (i.e., 0.25-0.125 mm). The similarity of this sediment to a sand sample from site 20 is discussed below in the section on the recent sand cover.
Kommos Survey Area
47
In the Pitsidia stream valley, because of its smaller size and fewer scarp exposures, there is less evidence concerning recent colluviation. One colluvial scarp exposed along the Pitsidia stream valley near site 23 showed near its base a thin stratum of subangular bedrock pebbles and cobbles, within which were several badly worn but probably Minoan sherds. Sherds of Hellenistic/Roman date sometimes were found in isolated colluvial deposits along the slope bases of the Matala valley. At site 131 in the Matala stream a 2.8 m-high scarp of colluvium yielded a large fragment of a Hellenistic/Roman bowl from the middle of the deposit. Thus, in the Kalamaki drainage basin, there is evidence for an episode of colluvial deposition beginning in Late Minoan time or soon thereafter. Some evidence exists in the Matala valley for colluvial deposition during or after the Hellenistic/Roman period. The major colluvial episode in the Kalamaki valley, and thus the major period of hillslope erosion, must have ended during Hellenistic/Roman times, since a structure of this date was built into colluvium at site 56 and, after abandonment, this structure was covered by a sand deposit quite distinct from colluvium. Sherds of Hellenistic/Roman date occasionally were found in isolated colluvial deposits along slope bases of the Matala valley. All the colluvial deposits mentioned so far are of middle to late Holocene age, less than five thousand years old. Generally in the western Mesara region one must go north of Tymbaki, around Kokkinos Pyrgos, to see good late Pleistocene surface exposures such as were described above. Another locality where a comparable Pleistocene colluvium exits is 7 km to the southeast of Kommos, in the Ayiofarango Gorge near the chapel of Aghia Kyriaki. It is approximately 200 m north (upstream) of the Locality E7 described by Doe and Holmes (1977: 22). In the Kommos survey area, only two certain examples of Pleistocene-age colluvium were found, both immediately south of the Kommos site. On the present shoreline 100 m south of the Ayios Pandeleimon chapel, a steep-walled gulley in the coastal cliff is eroding back into sediments filling a narrow, fault-controlled bedrock valley that we informally named Vigles Valley (it is just to the south of that hill). The valley fill is composed of angular to subround pebble- to cobble-sized limestone fragments set in a fine-grained matrix of red sandy mud. It is a colluvial scree deposit and must represent the erosional products weathered from the interior slopes during the last glacial event experienced in the Mediterranean region. Its exposure here at the present coastline suggests that a large part of the deposit has been eroded away by postglacial eustatic sea-level rise. Very probably the small valley is filled with this Pleistocene-age colluvium, but the recent sand cover hides any surficial evidence of its presence. The second exposure was discovered in a small commercial sand pit in Vigles Valley, just off the dirt track leading down to the shoreline at Kommos. There the 10- to 15-m-thick sand cover has been excavated by machinery down to the contact with the underlying sediment, which is a deep red-brown, sandy loam containing limestone pebbles and cobbles. At two locations in this sand pit, illustrated in Plate 3.7, we discovered evidence of a well-developed
48
Physical Geology
soil on the underlying Pleistocene colluvium. Sherds discovered 10 cm deep in the soil horizon (site 133) were dated provisionally to Final Neolithic or EM I. Parsons has identified the soil formed on the colluvium as a Palexeralf (Chap. 6, Section 1, under "Paleosols")·
Alluvial Deposits Within a drainage basin, colluvium is constantly being produced at rock outcrops by weathering processes and transported down slope by gravity. Early in its inevitable migration out of a drainage basin and into the sea, most colluvium will be transported for some distance by the stream flow of the main channel and its tributaries. Because this more energetic process tends to remove the finer grain-size fractions, colluvial sediments carried or deposited by stream channel flow are altered enough to be recognized under the separate sediment category of alluvium. Alluvium in the Kommos survey area is derived from three sources: (1) weathering products of bedrock disintegration on slopes (slopewash), (2) reworked aeolian sand that was deposited in the lower reaches of all three drainage basins during the past several thousand years, and (3) remobilized alluvium from upstream deposits. The Kalamaki and Matala streams show, in upstream channel bank cuts, a basic sequence of alluvium overlain by colluvium that has been carried down from valley slopes and not yet entrained in channel flow. Downstream near the shore only alluvium is to be found, although in the Kalamaki and Pitsidia valleys it is nearly all covered by reworked deposits of the aeolian sand blanket. The channels cut into those two alluvial plains are nearly verticalsided and deep relative to their widths, reflecting the generally high percentage of finegrained sediment found in the alluvium (Schumm 1960). Many vertical, unvegetated channel scarps occur in the middle reaches of the Matala stream valley, indicating active erosion of its alluvial deposits. Deposits of stream channel floors are termed channel lag deposits. These are tabular bodies of sand, gravel, and coarser sediment, over which finer sediment is carried as the stream load during normal stream discharge. The coarsest sediments, cobbles and boulders, may be transported downstream only during exceptional floods with long recurrence intervals (Allen 1965: 129). The alluvium forming the modern channel lag deposits of all three streams ranges in size from sand up to cobbles and small boulders. Along the Kalamaki and Matala streams a clear correlation exists between the location of coarse lag deposits and channel-bank exposures of coarse alluvial fill, showing that present winter stream discharges are eroding material only from the pre-existing alluvial deposits, rather than transporting coarse alluvium from upstream sources. Under abnormal conditions of exceptionally high rainfall in the Kalamaki and Matala drainage basins, significant volumes of coarse sediment would be carried down
Kommos Survey Area
49
their valleys and introduced on to their alluvial plains in a manner comparable to the flood event of 15 May 1901 at Kato Zakro mentioned above (Hogarth 1900-01:123). Thus few or no traces of archaeological features would be expected on such terrain. Differences among the alluvial deposits of the three stream valleys are due primarily to the different rock types they drain as well as the range of drainage basin sizes. Small boulders often were observed in the base of deep channel scarps in the Matala stream channel, while they were infrequent in the Kalamaki channel scarps and absent in the Pitsidia channel. A direct relationship exists between the size of a stream (measured by average channel cross section or by area of drainage basin) and the maximum sediment load it can carry in flood events; it is thus to be expected that the coarsest alluvium is to be found in the Matala Valley. Each valley has its own alluvial units and they were not correlative or mappable from one to the next because of very limited exposures in the Pitsidia and Kalamaki valleys. In general, however, it is not feasible to map alluvial units from one drainage basin to even an adjacent one, as Pope and van Andel (1984) noted for the Argolid Survey. Only an areal marker horizon such as an ash fall or a soil horizon can be used with security for inter-basin correlations of Holocene alluvial units. The following discussion considers the alluvial deposits of the Kommos survey area as two groups: those accumulating within the confines of a narrow stream valley, where they may form terraces, and those that accumulate on the plain at the valley terminus where the stream reaches the Mediterranean. All three stream valleys exhibit relatively narrow, V-shaped cross sections, confining deposition to narrow valley bottoms less than 100 m wide. Colluvium has accumulated along most of the valley margins, where it has not been eroded by stream activity. Vertical accretion of alluvium predominates in all three drainages, as they do not have sufficient room in their valleys to migrate laterally. Alluvial deposits within each valley have themselves been partially eroded by recent channel incision, forming single pairs of depositional terraces (Ritter 1978: 269). Although it was not possible to date absolutely the time of incision of each valley's terraces, there is circumstantial evidence that in all three valleys they were formed in post-Late Minoan time. The clearest evidence was seen in the Matala Valley. Plate 3.8 shows a 50-m stretch of the Matala channel near survey site 130 (Moutsounia) where a terrace scarp formed by recent stream channel incision has exposed the alluvial deposit that comprises the terrace. Half a dozen large and unworn Middle Minoan pithos fragments, apparently from the same vessel, were observed to be intermingled with the alluvial cobbles (PI. 3.9). No sherds later than Minoan were ever noted in the coarse basal alluvium of the Kalamaki or Matala stream channels as well, which suggests that the regimes of all three streams were different during or soon after Minoan times, then carrying a greater maximum sediment load. This, in turn, might be linked to greater slope erosion.
50
Physical Geology Table 3.4. Characteristics of alluvial plains in the survey area.
Stream Kalamaki Pitsidia Matala
Alluvial Plain Area 0.56 km 2 0.37 km 2 0.64 km 2
Percentage of Drainage Basin Area 6.67 14.07f 4.3% *
Sediment Cover aeolian sand aeolian sand aeolian sand and alluvium
*Based on estimated Matala drainage basin area of 15 km 2
In contrast to the confined nature of deposition in the upstream valley channels, the alluvial plains of all three streams in the Kommos survey region are relatively open (Pl. 3.2). Some of their physical characteristics are given in Table 3.4. Most of the observations concerning Holocene alluvial plain deposition are derived from a very informative exposure in a commercial sand pit in the Kalamaki Valley, in the vicinity of survey site 20. Minoan walls were exposed here under 10 to 12 m of sand at the bottom of the main pit. The walls were constructed directly on a well-indurated, gravelly mud that contains several unworn Minoan sherds, including an MM IA handmade conical cup fragment. The grain-size distribution of this sediment is illustrated in Plate 3.4 ("K79S/20-FILL"). In contrast to the other two samples shown on this graph, the mud is extremely poorly sorted, with its component sediment grains spread throughout all four general size categories (gravel, sand, silt, and clay). In comparison with other analyzed sediment samples from the area, it most resembles artificial fill deposits from Kommos (see below, Section 3, "Geoarchaeology of Kommos"). In Mediterranean countries, the differentiation of alluvium on the basis of whether it was deposited during or subsequent to the last glacial period is of fundamental interest since the work of Claudio Vita-Finzi (1969, 1971) regarding sedimentation in Mediterranean valleys and its relationship to human activities and climatic change. We attempted to locate alluvium in the survey area that might represent Vita-Finzi's "Older Fill," but the Kalamaki and Pitsidia streams proved too small to exhibit traces of such deposits, if they ever existed there at all (see below). Vita-Finzi (1971) analyzed the grain-size distributions of alluvium samples from the Geropotamos River at localities near Mires and Aghia Triada. He proposed that at these and other localities a "daughter" deposit of "Younger Fill" had been derived from a "parent" deposit of "Older Fill" through the depletion of silt- and clay-size particles by running water. While his explanation may be correct, it is not applicable to the Kommos survey area, where no comparable Pleistocene fills occur in the valleys of the three streams and all of the visible
Kommos Survey Area
51
alluvium represents reworked Holocene weathering products or aeolian sand. Also, significant variability in the grain-size distributions of facies within the alluvial deposits of one valley make the task of correlation between even adjacent valleys (e.g., Kalamaki to Pitsidia) impossible. Thus the identification of specific parent sources becomes more tenuous.
Paleosols During the 1979 field survey a partly eroded soil was found to underlie the sand-covered land surface to the north and south of Kommos. One exposure of this buried soil already had been noted by L. Vance Watrous at site 75, about 200 m north of the excavated Kommos hilltop, during an earlier phase of the archaeological survey. Michael Parsons and the author studied this exposure, as well as one 200 m to the northeast that recently had been uncovered by erosion of the Pitsidia stream channel. A third buried soil exposure at site 133 was found underlying the thick sand deposit immediately southeast of Vigles Hill; the soil there had been exposed in the bottom of one of the sand pits by the pit operators' heavy machinery. These examples of buried soils are discussed in Chapter 6 (Section 1), where Parsons has identified those at all three exposures as being most comparable to the Palexeralfs of the southeast Matala uplands. It is possible that a large portion of the sand-covered Pitsidia stream valley to the northeast, east, and southeast of Kommos is underlain by these soils. Seven sherds of Final Neolithic or Early Minoan date were found below the exposed paleosol surface in the Vigles sandpit (mentioned previously), and, at site 75, sherds contemporary with the Kommos Middle Minoan settlement are embedded in the buried soil there. No sherds were observed in the Pitsidia stream channel soil exposure. There is strong circumstantial evidence that the soil's A horizon was eroded away during or after Late Minoan time. Some evidence that these soils formed the land surface during Late Minoan time comes from the scattered patches of soil sediment excavated in past seasons at Kommos. Plate 3.15 shows the cumulative grain-size curves of several samples of the buried soil; they will be discussed in relation to the use of soil sediments as a source of building material at Minoan Kommos. Plate 3.21 illustrates the X-ray diffraction pattern of the clay size fraction of a sample from the middle of the site 75 exposure.
The Recent Sand Cover Mantling the alluvial plains of all three valleys and the coastal ridges and extending more than a kilometer inland is a deposit of aeolian sand. Plate 3.2 shows the extent of the surficial
52
Physical Geology
sand deposit where its thickness is greater than 10 cm, covering about twenty percent of the survey area. The sand reaches thicknesses of over 14 m in deep pockets on the lee side of topographic barriers to the dominant northwest winds. This surficial sand deposit is not specifically mapped or identified on the 1:50,000 geological map of the Tymbakion region, but is readily apparent to visitors to Kommos who follow the Pitsidia stream valley from the village down to the site. Parsons notes in Chapter 6, Section 1, under "Details of Soil Mapping Units," that the soils termed Lithic and Typic Xeropsamments, which have developed (slightly) on the sand cover, make up about fourteen percent of the survey area (see Table 6.1). The sand's significance for understanding recent human activities in the area is demonstrated by geological survey and analysis. Commercial sand excavation at three localities in the survey area exposed these deep sand sequences. The pit on the Kalamaki alluvial plain was mentioned above. Minoan sherds in the sand cover overlying the walls at site 20 are sandblasted, reflecting the importance of aeolian processes along the southwestern Mesara coastal zone. The grain-size distribution of this sand is shown in Plate 3.4 ("K79S/20-SAND"). The similarity in grain-size distribution to the colluvial sand sample from site 56, approximately 500 m to the southeast, is clear and is supported by mineralogical comparisons of the two samples' modal sand-size fractions. Both samples, therefore, represent a post-Minoan sand cover, although the site 56 sample has been altered slightly—some finer grain sizes have been introduced—by its redeposition as colluvium. A substantial hiatus during the deposition of this sand unit on the Kalamaki alluvial plain is represented by a relatively hard stratum containing increased numbers of pebbles and Helicinid snail shell fragments. The stratum extends around the scarp sections of the Kalamaki sand pit at approximately 4 m below the present ground surface (6 to 8 m above the alluvium and Minoan walls) and marks a land surface buried by renewed sand deposition. Plate 3.10 shows another one of the commercial pits, between Pitsidia and Matala, as it appeared in 1979. The exposed sequence reveals high-angled rapid slipface deposition in its lower part, separated from overlying lower-angled sand strata by a discontinuity—a buried land surface. At the third commercial sandpit—southeast of Vigles Hill—a clear stratigraphic discontinuity also was noted at about 3 m below the present ground surface. Therefore, throughout the Kommos survey area it is assumed that the buried surface represented by the discontinuity records a time when sand accumulation, which had proceeded steadily and rapidly, ceased long enough for incipient cementation of the land surface. This subsequently was buried by renewed sand deposition. Although the sand blanket is concentrated on the alluvial plains of the three streams, to a lesser extent it also has covered the hill slopes surrounding them. Much of the sand there, however, has been eroded and redeposited in the valley bottoms and around the edges of the alluvial plains as a distinct colluvial unit. Plate 3.4 illustrates this fact: Unit 3 covering site 56 is nearly identical in its grain-size distribution to the primary sand deposit covering Mi-
Kommos Survey Area
53
noan walls at site 20 to the west. The author assumes, however, that the sand covering site 56 dates from the later period of extensive sand deposition and that the similarity in the two samples' grain-size curves reflects derivation from the same source. There is no doubt that the sediment source is the beaches of the adjacent Mesara Gulf shore. Sand continues to be blown on to the coastal strip adjoining the modern beaches, and small deposits of texturally identical sand were sampled from depressions at the crest of Pitsidia ridge between survey sites 16 and 74, some 70 m above and 1,500 m east of the present shoreline. While the source of sand deposition is not in question, the timing of this unusual geological event is uncertain. Much of the evidence comes from Kommos and, therefore, will be considered in Section 3, but some indications were noted at other locations visited in conjunction with the site survey. First among these is site 20 itself, where stone wall foundations accidentally were uncovered under 12 to 14 m of sand in the main Kalamaki sand pit. They were constructed on a fill composed of sediments derived from weathering of the surrounding bedrock hills and from the decomposition of earlier structures at the site, as described previously. A sherd from the fill at the base of the foundations was identified as MM IA. Both the fill and the wall foundations built on it were buried by two major, distinct episodes of sand deposition that must have begun after the mid-second millennium B.C. It proved worthwhile to examine carefully every square meter of accessible scarps in these commercial sandpits. Through the 1979 survey season, a total of five wind-eroded Minoan sherds and one probable Minoan stone tool had been located in the sand layer immediately above the lowest indurated ground surface horizon exposed in the Vigles and Matala sand pits. These artifacts suggest that, at least by Late Minoan time, a substantial fraction of the sand blanket already had been deposited southeast of Kommos. The sand accumulating in such depressions downwind of topographic barriers remained undisturbed in an environment stable enough to minimize post-depositional erosion and, therefore, the depressions preserve a relatively complete record of the depositional stratigraphy. The agricultural potential of all three alluvial plains, which prior to the event may have possessed a more fertile soil cover, would have been degraded by sand deposition as discussed by Parsons (Chap. 6, Section 3, "Retrodiction of Carrying Capacity"). In addition, because of the high infiltration capacity of medium-sized, well-sorted sand, surface runoff from rainstorms would have disappeared immediately. Surficial infiltration capacity of sands can exceed 50 mm of rainfall per hour (Young 1972) and it is unlikely that rainfall events of this magnitude occur on Crete more than about once per decade. Also, since well-sorted sands hinder the upward migration of groundwater by capillary action, the local water table would not have risen appreciably, so available water for crops would have been diminished. Today the poorly developed soils on the sand cover are not very productive and, soon after their deposition, they would have been even less so.
54
Physical Geology
Summary: Late Quaternary Geological History of the Kommos Region The present landscape in the survey area has developed on a bedrock sequence of relatively soft limestones and marls of Neogene age. Other than the partially exposed Pleistocene colluvium noted south of Vigles Hill, there is no record of deposition in the area for most of the Quaternary Epoch. Only some three thousand to four thousand years ago can we identify the beginning of slope erosion and accumulation of the weathering products as colluvium and alluvium in stream valley channels and on their plains. The three streams of the survey area have been eroding their valley deposits since some unknown time in the recent past, perhaps only the last few centuries. Two basic geological processes have operated during the past few thousand years in the Kommos area: erosion of slope sediments and deposition of cojluvium and aeolian sand. The combined effect of these processes has been to decrease the amount of agricultural land available in the region surrounding Kommos. While it is not possible to estimate the amount of slopewash that has occurred in the region there has been a clear net loss of sediment cover; sediment production on slopes has been outpaced by the rapid weathering of Neogene bedrock units in the Kommos region. Basal alluvial deposits containing Minoan sherds are exposed in modern channels of the Matala and Kalamaki streams. The stratigraphically higher sediment deposited in those valleys upstream from their small alluvial plains is slightly or completely un-reworked colluvium. This major shift from an alluvial to a colluvial regime seems to have taken place at the end of the Minoan period. The Pitsidia stream may be too small to show this sequence (or it is deeply buried), but the E7 locality at Aghia Kyriaki in the Ayiofarango Gorge (Doe and Holmes 1977: 22) shows a decrease in the bedload capacity of that stream at about the same time. An LM IIIA-B cup base found by the writer in the "Fines" unit at that locality in 1979 places an earliest possible date on the stratum and suggests that it was deposited between 1,300 and 30 B.C., despite Bintliff's assertion (1977a: 613) that the unit is not alluvium, but rather the "anomalous" result of localized human activities. As mentioned above, in 1969 Claudio Vita-Finzi presented archaeological and radiocarbon evidence for a two-phase Quaternary deposition scheme across the Mediterranean Basin: an "Older Fill" deposited about 40,000 to 8,000 B.C. and a "Younger Fill" dating from ca. A.D. 400 to 1800. The Younger Fill is eroded from, and occupies channels cut into, the Older, in many instances. Their contrasting morphologies and sedimentary structures suggested to VitaFinzi that the Older Fill was deposited by intermittent flow or even mass wasting (mudflow), while the Younger represents a lower energy depositional environment and seasonal rather than intermittent stream flow. He postulated a climatic origin for the Younger Fill, based on its widespread and apparently synchronous deposition across the Mediterranean basin, asserting that an anthropogenic origin through increased slopewash was inadequate (Vita-
Kommos Survey Area
55
Finzi 1971: 188). Defining the temporal and spatial distribution of this post-Pleistocene alluviation in the Mediterranean has been the object of several geoarchaeological projects in recent years, with the ultimate goal being identification of the phenomena as responses to either synchronic, basin-wide climatic change or to diachronic, local human disturbance of the landscape through farming and herding, or some combination of the two. The slope erosion that would have supplied the material for the Older Fill deposits was, in the Kommos region, at its height only 3,000 to 3,500 years ago. It had begun at an earlier, unknown time within the late Holocene rather than in the Late Quaternary Epoch. In addition, there is no evidence in the three stream valleys for an alluvial unit less than two thousand years old correlative with Vita-Finzi's Younger Fill. Other geoarchaeological studies of soil and slope erosion in the Mediterranean basin (Butzer 1974, 1980; D. A. Davidson 1980; D. A. Davidson and Tasker 1982; Doe and Holmes 1977; Lippman-Provansal 1985; Pope and Van Andel 1984; Rust 1978; Van Andel et al. 1986; Wagstaff 1981; Zuidam 1975) all indicate that when detailed studies of small catchments are undertaken, the resulting syntheses of geomorphological history reflect interdependent local factors more than pan-Mediterranean, synchronic climatic change. In fact this was recognized in non-archaeological studies of changes in catchment erosion (e.g., Douglas 1967) some time ago. It appears that the two fills described by Vita-Finzi were over-generalized phenomena seen in the limited data base then available and that accelerated soil erosion in the Mediterranean basin during the late Holocene can be attributed more to local human impact on the landscape than to climatic variation (Butzer 1980: 138). In the Kommos region, as elsewhere in the Mediterranean, there is circumstantial evidence for a relationship between intensive land use and the onset of slope erosion. By Late Minoan time the cultivation of slopes around the Kalamaki, Kommos-Vigles, and Arolithia site concentrations would have breached whatever natural vegetation cover existed, fostering sheetwash and sediment accumulations as both colluvium and reworked alluvium in valley channels. All three of the small stream valleys in the Kommos region would have received large quantities of slopewash, but it is best recorded as alluvium at survey site 56 and along the middle reach of the Matala stream's north branch. The geoarchaeological results of the Argolid exploration Project (Van Andel et al. 1986) suggest a relationship between Early Helladic cultivation of certain valleys in the southern Argolid and the accumulation of the local Pikrodafni Alluvium unit. In the Kommos region the apparent low density of settlement in Early Minoan time did not initiate slope erosion to such an extent that it is recorded in the preserved colluvial and alluvial units. Regarding the rate at which substantial alterations in landscape characteristics may manifest themselves, archaeologists who believe the natural response time might be too long to interact with human activities should consider the work of Leopold (1976), who shows that changes indeed can be measured on the scale of decades, even in areas where human influ-
56
Physical Geology
ence is minimal. He presents data on long-term monitoring (7-15 years) of arroyo erosion trends in New Mexico (channel cross sections, gully headcut retreat, and sheet erosion). All three data types show that the rate of erosion and deposition had slowed during the period 1968-77, relative to the previous seven years. Also, the half-lives of recent gully systems in Colorado as calculated by Graf (1977) are some two orders of magnitude shorter than the Middle to Late Minoan periods, suggesting that whatever disequilibrium we may observe today in the drainages of the Kommos region is a result of more recent disruptions and not those of the Bronze Age. Roberts's (1979) study of geomorphology and environmental change in the Kairatos River valley at Knossos provides a geographically closer example of local factors affecting slope erosion. The rock formations exposed in that valley channel are nearly equivalent, if not correlative, with the Pliocene marls of the eastern Mesara (they also are known locally as kouskouras). The most recent alluvial unit in the Kairatos is dated by Roberts as contemporary with Vita-Finzi's Younger Fill. But instead of accepting his climatic causal explanation, Roberts proposes that its deposition may have been controlled by the increase of unconstrained spring flow into the valley after the late fourth-century destruction of Roman aqueducts. With the rebuilding of an aqueduct system in the seventeenth century (to supply Ottoman Candia), Roberts proposes that the unconstrained flow ceased, as did slopewash and local alluviation. However, to argue for the priority of human influences over climatic ones in effecting changes in the erosional/depositional regime is not to insist that all such changes are negative. From an agricultural survey of Epirus conducted in 1967, J. Hutchinson (1969: 89) concluded that the naturally high rate of erosion in that region had not been affected by man's agricultural activities and that the effect of erosion on the agricultural potential of the region has been substantially favorable; little good land has been lost and a large and productive area of new and highly fertile land had been gained. For the Kommos survey area, one might imagine that the buried stone check dams and terraces observed on the Kalamaki stream tributary to the east of site 56 (Sendones-Langos area; see Chap. 7, catalogue of sites) were constructed in response to worsening soil erosion at the end of the second millennium. Visible remains of terraces were noted by Parsons in other parts of the survey area, but they are probably of Hellenistic/Roman date and are located near known coeval sites. Minoan terracing, like other Minoan structures, may exist at the base of the Pitsidia valley slopes, but later colluvium would cover them and only by chance might local erosion or human activity bring them to light. The shoreline of the southwestern Mesara Plain has supplied windblown sand to the inland coastal zone from the second millennium B.C., in one major and several minor episodes of accumulation. The exact timing of the sand deposition may be related to recent sea-level oscillations and local tectonism, as discussed in the synthesis of Section 4.
Site of Kommos
57
3. The Site of Kommos From the previous consideration of the regional or mesoscale environment of Kommos (Butzer 1980: 38), we now focus on the site microenvironment. This interpretation is based on sediment grain-size analyses of site strata and multivariate statistical analysis of the resulting data.
Geology and Geomorphology As mentioned before, the middle of the Pitsidia valley contains a number of small bedrock knobs and ridges that rise above the general level of its floor. Along the valley's seaward edge there is a larger ridge oriented north-south that blocks the center third of the alluvial plain, forcing the Pitsidia stream around its northern edge. The "Hilltop Houses" of Kommos were built on the southern crest of this ridge and down its southern slopes to the level ground separating the ridge from Vigles Hill, which rises some 40 m above the highest point of Kommos. It was on this level surface that the public structures of Kommos were built. The strike of the ridge's bedrock is north-south in the vicinity of Kommos, turning sharply to east-west at its northern tip overlooking the Pitsidia stream; dip slopes of its strata range from southeast, 5° to 15° beneath the Kommos structures, to north, 2° to 3° at the northern end. In the 15-m-high cliff backing the present sand beach is exposed a sequence of white (2.5Y 8/2) fissile marl strata alternating with greyish yellow (5Y 8/4) sandy fossiliferous mudstones and limestones that show penecontemporaneous deformation. To the southeast of the Kommos excavation, Vigles Hill exhibits on its bare western slope a southeast-dipping sequence of marine marls, mudstones, and sandstones, as does Asphendilias Ridge to the north of Kommos. At survey site 75 just north of Kommos, the bedrock underlying a paleosol there is a distinctive pisolitic limestone that is not present in the Kommos cliff section. These rocks all appear to correlate with the Varvara Formation defined by Meulenkamp et al. 1977 (cf. Table 3.2). Plate 3.12 (top) conveys an idea of the site's general morphology prior to excavations beginning in 1976, as viewed from the southeast at an angle of 15° above the horizon. A clear transition is visible along the seaward slope from the free bedrock slope in the northern half of the area (detailed stratigraphy is not displayed on this figure) to the low-angle sand cover in the southern half. In the lower portion of the same figure, an attempt is made to portray the bedrock surface as if all the overlying sand and unconsolidated sediments and cultural strata were absent, in other words, something like the appearance of Kommos prior to habitation. It is a less realistic rendition than the upper illustration because it is based on only about 250 data points, locations non-randomly distributed across the site area where bedrock was recorded at the bottom of excavation trenches and test pits. Thus the near-vertical dis-
58
Physical Geology
continuities in the central section of the figure are not real but do indicate steep transitional slopes between the sub-horizontal surfaces. Plate 3.12 indicates two things. First, there is a distinct intermediate bedrock surface transitional between the hilltop and the flat surface to the south. This step, also visible in lowaltitude balloon photographs of the site, underlies the structures of the central hillslope. It was cleared for a distance of 15 m westward in Trench 2Al (J. W. Shaw 1977b: 228). While there is no clear indication on the coastal cliff, there may exist a small fault scarp between the central hillslope and hilltop areas. Second, the southern edge of the Central Hillside area also represents a well-defined slope break between southeast-dipping bedrock strata encountered in excavation squares and the sub-horizontal bedrock surface, essentially a monocline, in the southern third of the site where the major Minoan structures are located. Although bedrock continues to rise inland (eastward), it does so much less rapidly than the present sand-covered surface. There is some indication from the subsurface contours that bedrock curves around to the northeast and decreases in elevation. Samples of the limestone from the shore cliff at Kommos and from survey site 70 on Vigles Hill were prepared and studied in thin section. They were identified as biomicrosparites, in Folk's (1974: 164) classification of limestone rocks. Under the microscope, partly dissolved "ghosts" of foraminifera tests were observed, as were unaltered sponge spicules of chalcedonic silica and secondary growths of an amorphous, opaque mineral, probably pyrite. Microfossils such as foraminifera, sponge spicules, and radiolaria are common constituents of those marine sedimentary rocks, which also contain localized concentrations of macrofossils such as oysters and echinoids (Appendix 3.2). While it is quite clear that the Middle Minoan buildings on the Hilltop and Central Hillside at Kommos usually were founded directly on bedrock (e.g., south wall of Room 14c, Cliffside House), it also happens that the dip angle of bedrock strata below the hilltop is less steep than was the Minoan ground surface. Therefore the several rock units exposed in the coastal cliff would have been exposed on the southeastern (landward) slope as the Minoan builders set wall foundations, leveled floors, dug drainage channels, and otherwise modified the original pre-habitation surface of the hill. In numerous excavation trenches different bedrock types were exposed, in areas ranging from a few square centimeters to several square meters. Below the Hilltop marls and friable fossiliferous limestones are the prevalent rock type, but on the southwest crest of the Hilltop a resistant fossiliferous limestone forms the bedrock surface and partially accounts for the existence of the Hilltop. This bedrock type correlates well with the calcarenitic facies of the Varvara noted on the crest of the Matala headland. The upper surface of the bedrock exposed below Trench 12A2 shows a weathering crust (duricrust) comparable to Pleistocene weathering horizons developed on limestones in other parts of the Mediterranean. In the vicinity of Trenches 2A/9A the bedrock is a much softer mudstone. Because it weathers more rapidly than the overlying calcarenitic limestone, large blocks of the latter rock are continuously be-
Site of Kommos
59
ing undercut and broken away. Most of the blocks on the backbeach at the foot of the cliff were detached from the calcarenitic limestone stratum. Distinctions among these rock types are a part of the Cretan dialect. The generic name given by local inhabitants to the rock identified here as Neogene calcareous marl is kous kouras; it was identified as the sterile bedrock underlying many structures at Kommos (e.g., J. W. Shaw 1977b: 223; 1981a: 216; 1982a: 172). Wagstaff (1972) also notes that marls are termed kouskouras in the Myrtos region, as does Roberts (1979: 232) for the Pliocene-age bedrock around Knossos. Warren (1969a: 256) defined kouskouras as "any soft, white Tertiary lime stone." The harder fossiliferous limestones are termed poros and are distinguished on the basis of their suitability for building stone. Asprochoma appears to be a fine-grained, less in durated marl that was used especially for sealing roofs. Deposits of it were noted in the exca vations (e.g., Space 8 of Trench 4A; J. W. Shaw 1977b: 220). Lepidha is a name given to a greenish grey or blue-grey clay commonly associated with kouskouras; it was used as a packing/leveling material in the structures of Kommos (e.g., J. W. Shaw 1982a: 173, 183). A. Evans (1928: 328) describes this material and its appearance in the roof terrace cement of Knossos and states that it "crops out on opposite hill-sides East and South of the Palace," presumably in the Pliocene marl stratigraphy noted by Roberts (1979: fig. 3). A material of the same name has been associated with weathered schist (J. W. Shaw 1973: 317), but, if that is the case, the name is used for a very broad range of dissimilar geologic raw materials. Certainly the results described in the following paragraph point to a sedimentary origin for the Kommos lepidha. Cumulative weight percent curves of three kouskouras bedrock samples from excavation trenches 9A/5:36T, 20B/4:92, and 35A2/9:120 are shown in Plate 3.13. In addition, a bedrock sample not specifically identified as kouskouras, from an exposure 15 m northwest of Build ing J, is included in the plate ("80A/BR NW J"). The four are almost identical in their grainsize distributions, being very poorly sorted mud or sandy mud, with a median grain size of 6.8 to 7.2 φ and having 34.8 to 40.9% clay. Microscopic examination of the sand fractions shows they are are foraminifera-rich (Globigerina and Orbulina spp.) calcareous muds with almost no detrital minerals. For comparison, a sample of material identified as lepidha ("35A2/8:119") is shown in the same figure. Other than a slight depletion in silt (4-8 φ), it is very similar in its grain-size distribution to the kouskouras samples and is identical in terms of its microfossils as well. In comparing the minerals present in the clay size fractions of kouskouras and lepidha samples, little difference is evident. All contain quartz and calcite as well as showing strong XRDA peaks indicative of the clay minerals kaolinite and smectite. Lepidha may be an in-situ weath ering product of the marl bedrock, but there is no sign in the lepidha samples analyzed of alteration of the two clay minerals, as would be reflected in a decreased intensity of their X-ray diffraction peaks. Over the decades prior to the present archaeological excavations at Kommos, there had
60
Physical Geology
been little sediment erosion or deposition in the excavation area. The free rock face below the Hilltop structures always has been eroding slowly as a result of direct attack by major storm waves because there is no accumulation slope present between the active beach zone and the base of the cliff. Equilibrium existed between the influx of windblown sand on the central and southern portions of the site and its erosion through winter storms. The high permeability of this sand cover fostered infiltration of rainwater and attenuated erosion over the central hillslope. Local informants recall that after major storms enough sand was temporarily swept from the southern portion of the excavation area to expose some architectural features of the site. Wind erosion is occurring along the crest of the shoreline cliff, thereby maintaining Minoan foundations at or very near the ground surface and thus accounting for Evans's discovery of them. Second World War air photographs of the Kommos vicinity show a natural drainage channel leading from the base of Vigles Hill at its closest approach to the site westward across the sand-covered ground surface, depositing a small alluvial fan in the vicinity of the (buried) Minoan seaside buildings at the southern end of the site. This deposit subsequently was covered by the influx of windblown sand. Today, slopewash from the west slopes of Vigles is channeled within a 2-m-deep drainage gulley eroded where the hillslope meets the coastal sand deposit. This is some 150 m south of its position during the mid-1940s. As the major agent of erosion at Kommos, wave attack was concentrated on the shoreline cliff during Minoan and post-Minoan times, as indicated by truncated foundations of major structures in Trenches 1B3 and 12A2. This erosion continues during the present winter storm regime, as the only detached bedrock blocks to be found along the cliff base are from marl and soft limestone near the base of the section. In a comparable geomorphological setting, Wagstaff (1972: 280) noted the loss of a strip of uncertain width of habitation at the Fournou Korifi settlement, which occupies a ridge of Pliocene marls on the south coast of East Crete. During the several centuries' occupation of the Hilltop Houses at Kommos there would have been incessant erosion all along the cliff top, and in Trench 13A a downslope retaining wall was built on earlier strata to stabilize a building endangered by this natural process (J. W. Shaw 1978a: 115). It was noted that no walls were found in any of the 2A series of excavation trenches (J. W. Shaw 1977b: 223), presumably owing to loss by erosion. Wave erosion of the soft Neogene bedrock ridge has shaped the form of the shore of the Kommos ridge northward to the mouth of the Pitsidia stream. The bed of this small ephemeral stream is incised into bedrock for the last 200 m of its course across the modern back beach zone; it, therefore, never could have served as a natural shelter for ships as A. Evans (1928: 90) suggested. An aspect of the site's geomorphology that Evans more perceptively noted is the "halfsubmerged reef," as he termed it (A. Evans 1928: 88), lying some 300 m due west of the Central Hillside area of Kommos. This rock reef, known as Papadoplaka (the "Priest's Slab"), is an outcrop of Neogene bedrock rising abruptly from a level, sand-covered bottom 7 to 8 m
Site of Kommos
61
deep. The morphology of the outcrop is that of an eroded, possibly fault-bounded fragment of undeformed marl and limestone strata dipping southeast at 5° to 10°, similar to the bedrock underlying the excavation. As a result of this, the eastern dip slope of Papadoplaka facing Kommos is quite gentle, while the western edge of the reef drops abruptly into deeper water of the Libyan Sea. There is a slight trace of a horizontal planation surface on the dip slope at 2 m below sea level. Plate 3.11 shows the relationship between the site and the submerged eastern edge of Papadoplaka (most of the outcrop is to the left of the photograph). A topographic sketch of the outcrop is given in Mourtzas (1988: fig. 3a). Less than ten percent of the total area of this outcrop presently rises above sea level, as estimated from the area of outcrop visible above the sand bottom. Because the shoreward side of this rock reef is the dip slope, a lowering of only 2 m of sea level would expose an area of rock extending toward Kommos some two to three times larger than at present, or about twenty to thirty percent of the entire outcrop. Water depths between this enlarged offshore reef and Kommos would be ca. 5 m, and, at least, a submarine ridge would exist between it and the mainland, as is the case today at the site of Traphos (ancient Lasais) some 10 km to the southeast of Kommos on the south coast. This will be considered below in terms of the sea-level history of the area (Section 4). While the bottom between Kommos and Papadoplaka is sandy, it is a very thin covering of mobile sand over a level bedrock surface. In Plate 3.11 one can see, midway between the modern shoreline and the submerged eastern edge of Papadoplaka, a patch of this bedrock surface temporarily exposed and covered with marine vegetation (darker tone), quite distinct from the otherwise light sandy bottom. One may observe underwater small ridges ten to 20 cm high protruding from the sand cover, marking more resistant bedrock strata. Pebbles, cobbles, and small boulders have accumulated in the lee of these ridges. This is the area where in 1983 a Minoan stone anchor was discovered (see synthesis of Section 4). Kommos lies at the southern end of the Mesara coastline's sand beach (Pl. 3.5), but there is no evidence of nearshore sand accumulation in the form of longshore bars off the site. Sediment transport occurs in the foreshore zone driven primarily by a strong southerly vector of longshore current (R. A. Davis 1978a: 267), but in the nearshore zone the drop-off into water deeper than three m prevents the formation of longshore bars. Bottom transport of sediment in the nearshore zone west of Kommos (3-8 m deep) evidently occurs as a thin blanket of sand constantly moving southward over a nearly level bedrock surface. Plate 3.11 also clearly shows the beach and nearshore morphology of the Kommos site. The present beach is narrowest off the Hilltop area (ca. 15 m wide), where it is backed by the Neogene bedrock cliff, allowing only a slightly developed backshore berm. This is in contrast to the beach profile just to the south, from the Minoan seaside structures southward. There the pre-excavation beach width is estimated to be ca. 30 m, and there is a wide backshore sloping inland to the sand-covered ground surface. The beach at Kommos responds rapidly to slight changes in the wind and wave regime
62
Physical Geology
(commonly from west to northwest or north and back to west), building out or eroding back by several meters over periods of just a few days. A steep foreshore zone exists during the constructional phases, fostered by strong and constant northwest winds (the meltemi); subse quent erosion of the constructional beach profile, when the winds shift to westerly, produces ephemeral sand scarps up to half a meter high in the swash zone. Plate 3.14 illustrates the grain-size distribution of a representative beach sand sample ("MODERN BEACH") from the swash zone west of the Central Hillside area at Kommos. It is a well-sorted sediment composed of 99.8% sand-sized mineral and rock grains having a mean size of 1.41 φ (0.375 mm). The most common (modal) size fraction, 0 to 0.5 φ, shows metamorphic rock fragments and several populations of quartz grains (from very wellrounded and polished to angular), both derived from the Geropotamos River, and a few fos sil peneroplid foraminifera tests derived from the erosion of Neogene bedrock. Although no sediments coarser than granule size now constitute the beach in front of Kommos, numerous isolated rock slabs, partially or completely buried in back beach sedi ments, represent either fragments of Neogene bedrock detached by storm waves from the foot of the coastal cliff or tabular blocks of beachrock similarly detached in the nearshore zone and carried by storm waves 10 to 20 m inshore. Some of the bedrock blocks are visible in Plate 3.11 at the foot of the Hilltop scarp. The sand beach at Kommos is so responsive to relatively slight changes in wind and wave directions because it rests on a 5 to 15 m wide deposit of beachrock underlying the present swash zone. The beachrock is clearly visible in Plate 3.11 as a dark band (due to a coating of marine vegetation) sloping at a 2° to 3° angle from the swash zone into 1.5 to 2 m of water. There the deposit ends in a vertical face that drops into 3 m of water. The deposit's offshore edge has been undercut and collapsed along most of its length. Its thickness ranges from 0.5 m near the north end of Kommos to 1.5 m just off the Central Hillside area. From the balloon photograph (Pl. 3.11) it is clear that the deposit is narrowest off the Hilltop area (10-12 m wide; off the southern half of the site it widens to 40 m). Boekschoten (1963: 243) describes this particular occurrence (at "Pitzidia") in his survey of beachrock around the island. He was mistaken, however, in his estimate of this deposit's width and also in his assertion that beachrock deposits are absent directly in front of coastal bedrock outcrops.
Site Ground Water Conditions and Beachrock Formation The water well at the south end of the modern retaining wall along the shoreward edge of the Kommos site, some 80 m from the present shoreline, gives the most direct evidence of ground-water level. In June 1977 the water surface in the well was 1.7 m above sea level, with 5 m of fresh water to its bottom. Using the Ghyben-Herzberg equation (Freeze and Cherry 1979: 376) it can be calculated that approximately 70 m of fresh water existed below sea level in the vicinity of the well at that time.
Site of Kommos
63
The free-standing water surface in the well at Kommos marks the top of the zone of phreatic water, the water table, below which the bedrock is saturated with water. In the Kommos area the aquifer, or body of rock and sediment that serves as a reservoir and conduit of groundwater, is unconfined and variable in its elevation, depending on the seasonal influx of rainwater. Because the fresh water of the aquifer is slightly less dense than the salt water of the Mediterranean, it "floats" on the latter and extends several tens of meters deeper below sea level than the water table rises above sea level. Moreover, the less dense ground water is constantly moving from inland areas of greater hydrostatic potential to the sea. Extensive mixing of the two water types only occurs just at the junction of the water table with the open sea surface and the air, which at Kommos is within the sand of the beach backshore zone. The rise in the general level of the water table at Kommos since Minoan times is shown by several instances of architectural features and cultural strata extending below the summer water table in the southern part of the site (J. W. Shaw 1981a: fig. 6; 1982a: 183). Today, Building J's eroded western foundations extend to the level of fresh-water seeps from the marl bedrock surface underlying the foundation blocks, at an average elevation of 2.6 m above sea level. There is indirect evidence in the form of a raising of the floor level in the western part of Building T by 22 cm (to 3.2 m above present sea level) in LM I to LM IIIA1 Phase 2 time (J. W. Shaw 1984a: 272). North of this area, on the paved Minoan road running alongside Buildings T and J, ground water was noted to seep into the lower part of the bedrock drainage channel along the north side of the roadway (J. W. Shaw 1982a: 178). It is unlikely that this 30-cm-deep channel would have been cut originally below the contemporaneous water table, and 30 cm would be a minimum estimate of the net rise in the water table at Kommos since LM I—III time. This rise, which doubtless also occurred beneath the rest of the site, must be linked with a rise in sea level since the Late Bronze Age. The interaction of this process with changes in the water table and the land surface will be considered in Section 4. Present-day formation of beachrock was discovered below the modern beach sands in a series of shallow auger holes in the backbeach zone 30 m south of the main sand dump. A transition zone of friable, partially cemented grains was encountered in many auger holes; this graded within a few centimeters to a fully cemented rock that the auger could not penetrate. In contrast to the contemporaneous beachrock formation occurring beneath the backbeach zone at Kommos, there exists in the surf zone a deposit of beachrock that is clearly relict, having been cemented below the foreshore of a beach extending significantly farther to the west than today's. The exposure of this beachrock deposit indicates that the beach at Kommos has retreated since its cementation, and retreat requires a relative sea-level rise (some combination of a eustatic rise of the sea surface and a tectonic lowering of the local land surface). Petrographic study of thin sections of samples from the submerged beachrock deposit off
64
Physical Geology
Kommos revealed that three distinctive carbonate cement types are present; from the surface to the base of the deposit they are: (1) a homogeneous type composed of well-sorted and rounded terrigenous sand grains with a single ΙΟ-μιη-thick rind of isopachous micrite cement (the "hard micrite cement" of Alexandersson 1972: 212) with empty intergranular pores; (2) beachrock of fine-sand-sized, well-rounded terrigenous sand grains cemented by a fourlayer cement of alternating micrite and sparry, low-magnesium calcite crystals; and (3) a beachrock type composed of 90% biogenic grains (foraminifera, micromolluscs, and cor alline algae) and 10% subround terrigenous grains, all indistinctly cemented by a pasty micritic cement. Beachrock types 2 and 3 record different beach environments in terms of cement types and grain types, respectively. Beier (1985) has demonstrated for Bahamian beachrock that ce ments of multiple morphologies reflect diagenesis in environments ranging from marine phreatic (intertidal zone) to vadose (near-surface soil zone) depending on the position of sea level. The multiple morphology of Kommos Type 2 beachrock probably also reflects an os cillation of sea level, as discussed below (Section 4). Although an extensive search was made for datable potsherds in the beachrock west of Kommos, none were found. Radiocarbon dating of beachrock cement is possible in theory, but the practical problems of isolating a pure sample and of multiple cementation events makes this an equivocal technique. The total absence of any sherds throughout this large beachrock deposit, in the immediate vicinity of a major settlement, suggests that the deposit either predates or postdates its time of occupation. However, in the latter case, one might expect to find some re-worked sherds included.
Geoarchaeology of Kommos With the establishment of the Kommos settlement in the early second millennium B.C., the area ceased to be a natural system in dynamic equilibrium with the physical environment and became an artificial environment subject more to human than natural alteration. Succes sive building phases, the construction of retaining terraces, and the leveling/infilling of small areas, all would have modified patterns of local drainage, erosion, slopewash, and sediment accumulation, which would produce deposits on a spatial scale comparable to those of a pri mary archaeological nature (floors and fills). The Minoan builders of Kommos structures used the entire range of locally available inor ganic raw materials: (1) two primary types of local limestone bedrock for wall and stone floor construction, plus occasional use of more exotic rock types such as red schist (J. W. Shaw 1978a: 127);
65
Site of Kommos
(2) soil sediments as mudbrick and packing for stone walls; (3) clay-rich marl bedrock as a floor surfacing and as a major component of roof surfacing; (4) sand (possibly) intentionally mixed with soil sediment as a sub-surface packing material below stone floors; (5) natural beachrock boulders and slabs in wall construction; and (6) beach pebbles mixed with lime to make the underpavement known as chalikasvestos, which is characteristic of Late Minoan courtyard surfaces. From this, three sub-topics present themselves for consideration at Kommos: the prehabitation surficial deposits, the evolution of the artificial fill mixtures that dominate site sedimentation throughout its existence, and the nature and timing of the natural sediment influx recorded in the site's stratigraphy. Because paleosols, more specifically buried soils, were discovered both north and south of Kommos (Chap. 6, Section 1, under "Paleosols"), it was considered likely that a soil horizon had developed at the site itself prior to construction disturbances in the second millennium B.C. A search was made, therefore, through geological samples and excavators' notebooks for indications of sediments that may have once been part of a soil profile. In the sense used by Butzer (1971: 170)—"older soil materials that have been eroded and redeposited by streams, gravity or wind action"—soil sediment represents a useful geoarchaeological category. Table 3.5 lists those examples that were noted in the stratigraphy of Kommos. Grain-size distributions of three of these samples have been plotted in Plate 3.15, along with paleosol samples from site 75 (middle of B horizon) and the Pitsidia stream soil exposure 300 m northeast of site 75, at a depth of 1 m below the present surface ("79S/PITSOILlM"). Sherds from the site 75 paleosol exposure were tentatively classed as Middle Minoan (Chap. 7). The sample from the Hilltop Trench 13A ("77A/13A/3:43") is described by the excavator as ground-up bedrock with cultural material, predominantly Middle Minoan. Sample 79A/28B/3:14 is a reddish sediment, possibly mudbrick, in situ on top of Wall 58. Its date is MM HB. While the percentages of medium to fine silt fractions vary among these samples, Table 3.5. Excavated soil sediments derived from a Kommos paleosol. Year
Trench
Level
Pail
K77A K77A K77A K78A K79A K80A K80A
HA 13A 13A 22A2 28B 34A2 36B[N]
3 3 3 3 3 4
8 27 43 102 14 34
Character
possible paleosol (79NBp99; XRD-silt) possible paleosol (79NBpIOO; XRD-sample) possible paleosol (79NBp. 99) "Top Wall 58" (mudbrick or soil sediment) "red soil" (paleosol?; 91) "below MM II stone floor" (93; XRD-sample)
66
Physical Geology
the similarity of their sand and coarse silt fractions both in terms of weight percent and min eralogy (from X-ray analysis), make it very likely that all five represent the same paleosol, developed on the Kommos ridge and eastward into the Pitsidia alluvial plain prior to exten sive sand deposition there. Most traces of the original paleosol at Kommos would have been erased through artificial mixing of various sediments during the settlement's habitation (see below), but small fragments of the soil sediment survived in the matrix of more homogenized fills (cf. Pl. 3.15 "22A1/3:8FILL"), to be recognized by their more reddish color and finer texture. Whereas most of the sand at Kommos was deposited during and after the latest habitation phases, excavations in several trenches show that sand was present in Late Minoan and even Middle Minoan levels (J. W. Shaw 1978a: 148; 1982a: 169, n. 14), often mixed with finer slopewash to form a sandy colluvium. Plate 3.16 illustrates three such samples from the site and one sample of an equivalent sediment from survey site 56 (described above). Middle Minoan Barbotine Ware excavated from Pail 38, Level 3 of Trench H A (J. W. Shaw 1979a: 148), just above bedrock, allows the inference that this poorly sorted, slightly gravelly sand was being deposited in the southern part of Kommos during the early second millen nium B.C. It is almost pure sand (88.4%) that could only have been derived from the contem poraneous beach zone by wind deposition. The 1 to 1.5 φ fraction shows well-rounded and polished igneous rock fragments derived from the beach as its main component, plus subround, non-polished igneous rock fragments, fossiliferous marl bedrock chips and grains, and foraminifera (Orbulina and Globigerina spp.) tests. With only a few percent of gravel-, silt-, and clay-sized particles in the sediment, its sand fraction must have been deposited ei ther very rapidly (so as to greatly dilute the influx of these different-sized particles) or during a time when little slopewash was being eroded from Vigles Hill and deposited in the south ern part of Kommos. Sample "K80A/36B/4:32" of Plate 3.16 represents a nearly pure sand (96%) from a small hole or depression (not an animal burrow) in the northeastern corner of Trench 36B; the sherds from this pail were found to be of LM III date. Again, the mineralogy indicates this is windborne beach sand, only slightly mixed with finer-grained sediment to form a colluvium. The "K77A/SCARP LOWER" sample of Plate 3.16 is a sand sample from a 1977 geological section in the southern part of the site between Trenches 1OA and H A . The sample repre sents the beginning of a major event of sand deposition preceding the most recent one, which I believe dates from soon after the site's abandonment in the first century after Christ. As is true of the other two samples, rapid sand influx or the lack of different-sized sediment components have produced a sediment composed of ninety percent sand. The preceding two samples may be compared in this figure with "K79S/56-UNIT 3" (which also was shown in Pl. 3.4); this illustrates the similarity of sediment mixtures that may arise from one geological process (colluviation) acting on the same sediment types (aeolian sand and weathered marl bedrock) in different parts of the survey area.
Site of Kommos
67
While the samples illustrated in the previous three figures have been identified on the basis of their similarities in mineralogy and grain-size distributions, by far the largest number of sediments analyzed from Kommos show no modal size fraction (i.e., they are extremely poorly sorted) and belong to the general category of "fills." Fills also constitute (by volume) one of the most common sediment types at Kommos; almost 3 m of it cover certain parts of the Minoan habitation levels (e.g., J. W. Shaw 1977b: 238; 1984a: 269). Both the archaeological and the geological characteristics of these sediments identify them as very heterogeneous mixtures of materials reworked from two main sources. Plate 3.17 shows the grain-size dis tributions of five representative examples of this category. It is clear from the shapes of these five cumulative grain-size curves that no modal size fractions are present; the curves approximate a diagonal line connecting the lower left with the upper right corner of the graph, showing a monotonic accumulation of size categories. These sediments are termed gravelly or sandy mud and are very poorly to extremely poorly sorted. Their median size ranges from 4.2 to 6.0 φ, and their sand fractions all contain min eral grains and rock fragments derived from the buried soil cover, the local marl bedrock, and occasional microartifacts such as microsherds, charcoal fragments, and plaster frag ments. Two of the samples contain large numbers of phytoliths in their coarse-silt-size frac tions, indicative of admixtures of plant material. The coarse-silt fractions also contain the subround, quartz grains that are common in the soil sediments. These sediments were described by the excavators as "soft, sandy m u d , " "earth fill," "sand bedding" [under floor slabs], "soft brown earth," and "compacted earth and sand with granules of soft limestone;" they range in age from EM I—II through Classical. They represent recycled building material, which itself was derived from the soil cover around the site, and, to a lesser extent, from the marl bedrock. A comparable class of archaeological sed iments was identified and characterized in the strata of Troy (Gifford et al. 1982: 99). Although their macroscopic appearance in the field can vary in both color and texture, sediment grain-size analysis is a reliable means of identifying them. At Kommos, where both the bedrock and compacted fill deposits in the southern excavation trenches could look iden tical to the naked eye, this is the only means of differentiation. Fifty-two sediment samples were analyzed from Kommos and the survey area. While it was subjectively possible to recognize "types" of sediments and group them together on the basis of presumed depositional processes (as illustrated by PIs. 3.13, 3.14, 3.15, and 3.16), a more quantitative method was needed to formalize the analysis. With the availability of com mercial multivariate statistical analysis programs for microcomputers, it has become possible over the last few years to analyze large data sets such as this one, which originally consisted of a matrix of 52 sediment samples by 32 grain-size classes. Recent work on other geoarchaeological sediment samples (Vitali et al. 1992) indicated that the original 32 phi grain size categories into which each Kommos sample was divided repre sent too fine a division. Therefore it was decided to limit the statistical analysis to a subset of
68
Physical Geology
size classes, specifically 7 classes formed by the following combinations: (1.5 + 2 φ), (2.5 + 3 φ), (3.5 + 4 φ), (4.5 + 5 φ), (5.5 + 6 φ), (6.5 + 7 φ), and (7.5 + 8 φ). These 7 combined classes, representing the fine sand to silt size categories of each sample, were then recalculated to 100% and subject to correspondence analysis (Cibois 1987). The resulting scatterplot of the 52 samples and 7 grain-size variables in the first two facto rial axes is shown in Plate 3.18. A parabolic distribution of the 7 size variables ("phi2""phi8") about the first axis is evident, which reflects an underlying data structure related to grain size. Of more archaeological interest is the distribution of the 52 samples, which also is non-random. Because correspondence analysis allows the display of all samples and all vari ables on a single plot, examination of the scatterplot gives some insights into the processes of sedimentation, both natural and cultural, at Kommos. Three sample concentrations are identified in Plate 3.18 as "Bedrock," "Soil Sediment," and "Eolian Sand" because those are the geological sources of the samples in each respective field. In reality a continuum of sediment types exists for the Kommos samples; samples ap pearing at intermediate locations are mixtures of these three end members, which are repre sented by three of the figures of cumulative grain-size curves described above (PIs. 3.13, 3.14, and 3.15). For example, a sequence at the bottom of Trench 35A2 contributes much information to the question of early deposition at Kommos (see J. W. Shaw 1981a: 216 for description). The low est two levels, 8 and 9, are marl bedrock (Pl. 3.13), perhaps somewhat weathered in the case of Level 8, which was identified as lepidha. Above this marl, Level 7 was described as a typi cal fill sediment (Pl. 3.17) dating to the Early Minoan period. Overlying this fill is a stratum of reddish brown sand also of Early Minoan date. In Plate 3.18 the Level 8 and 9 samples appear in the bedrock field, while the Level 7 fill sample lies "outside" it, suggesting it is mixed with soil sediment. This would represent some natural mixing through local erosion, as there is no evidence of permanent habitation of the site during Early Minoan time. The overlying reddish brown sand (not analyzed) sug gests that a sediment source, presumably the late third millennium beach, was introducing sand into Kommos before there were any structures there. As another example, a comparison of the samples illustrated in Plate 3.16 with the "Eolian Sand" field of Plate 3.18 shows them to lie around the periphery of the field, reflecting their origin through mixing of sand with finer sediments from other sources, which ultimately would be the local soil cover and the eroded bedrock. The more basic conclusion to be drawn from examination of Plate 3.18 is that many of the samples characterized as fill represent mixtures of bedrock and soil sediments. That bedrock predominates could be the result of several factors, including the continued addition of " n e w " eroded bedrock sediment during centuries of digging on the site, whereas the original volume of soil sediment was a "non-renewable" raw material source (at least on this time scale).
69
Site of Kommos
From Plate 3.18 there is some indication that specific artificial mixtures were made for spe cial construction purposes (e.g., the mud plaster found in Level 3 of Trench 9A appears to be primarily composed of fine-grained bedrock, and the Middle Minoan floor bedding of Level 4, Trench 36B, is the very heterogenous mixture desirable for such an application). It is notable that the most anomalous sediment sample plotted in Plate 3.18 (i.e., that far thest from all other samples on the plot) is also the one that might be argued to reflect an undisturbed mixture of two Late Minoan sediments: the "ballast" discovered in the Snake Tube (J. W. Shaw 1977b: 229). This sediment is a pale yellow (2.5Y 7/4), extremely poorly sorted, slightly gravelly sandy mud, containing 49% sand-sized sediment grains, mostly in the 0.5-1.5 φ range. The 0.5-1 φ fraction shows subround to well-rounded igneous and metamorphic rock fragments, angular limestone fragments containing foraminifera and sponge spicules, and well-rounded and polished quartz grains; the 1-1.5 φ fraction is identi cal to modern beach sand in its mineralogy. A second and much smaller mode occurs in the fine-silt size (7-8 φ), which corresponds to a soil sediment mode. X-ray diffraction of the coarse-silt size (4-4.5 φ) shows minerals identical to those of the soil sediments. This is an anomalous, artificial mixture that never was subject to geological processes that would have sorted its grains into a more "natural" distribution. Plate 3.18 shows three trimodal sediment mixtures from the southern portion of the site. In all three cases the same sediment type is represented, a slightly gravelly, muddy sand with a median size of 1.9 φ and 63 to 68% sand. The "KOMMSCARP LEVEL 2" sample is geological. In the southern portion of the site area there appears to be a layer of sand mixed with muddy sediment derived almost certainly as slopewash from Vigles Hill; this was rather abruptly succeeded by a nearly pure sand deposit described above (PL 3.16,
"K77A/SCARP
LOWER"). The sediment sample from Pail 25, Level 4 of Trench 36B, represents a Middle Minoan floor bedding, and that from Pail 8, Level 3 of Trench HA, represents a stratum con taining a post-Classical sherd. The last-named sample, with an elevation of 6.3 to 6.7 m, is probably related to the ground surface of the Greek Sanctuary complex to the west (as de scribed in J. W. Shaw 1980a: 218ff.). It was also noted during the excavations that artifical floor packings often were found to be "a mixture of earth and beach sand, a type of packing used even today to prevent settling . . . " (J. W. Shaw 1977b: 213). The largest contributor by volume to the sediment cover at Kommos is windblown sand. Nearly pure sands are represented in Plate 3.18 by samples such as "77A/SCARP UPPER SAND", "80A/BLDG P 75-80 CM," or "80A/37A/4:40 SAND," all three of which also appear in Plate 3.14. This sand originated in the shore environment, as there is no other geological source of such well-sorted sediments in the western Mesara. What has emerged from the excavations is abundant evidence for sand deposition of widely different magnitudes during and after the site's occupation. As mentioned above, the Kalamaki and Pitsidia alluvial valleys are covered, and the Matala stream valley partly filled (Pl. 3.2), by a sand blanket that is partially colluvial in im-
70
Physical Geology
mediate depositional origin but which ultimately represents an aeolian deposit. Some indication of the time of deposition was given by the Minoan walls buried at survey site 20 in the Kalamaki stream valley. But the detailed stratigraphy of Kommos allows more precise bracketing of local sand accumulations prior to the regional post-Roman sand deposition. After the founding of Kommos, the presence of sandy sediments in Middle Minoan levels indicates the sediment source was never far away, although there is no sign of a continuous stratum over the site (J. W. Shaw 1981a: 217). A small deposit of sandy colluvium was excavated in Trench HA, Level 3, Pail 38 (PIs. 3.16, 3.18), which is dated by its Barbotine Ware to the early second millennium B.C. (J. W. Shaw 1979a: 148). Other evidence for Middle Minoan sand deposition was noted in the Central Hillside area (J. W. Shaw 1982a: 169, n. 14). When we come to the Late Minoan levels at Kommos, evidence of sand is both direct, as in the ballast of the Snake Tube (Pl. 3.18) and the LM III sandy colluvium in Trench 36B (Level 4, Pail 32; cf. PIs. 3.16, 3.18), and also indirect, as from the eleven stone tools excavated in LM IHA (or later) deposits on the Hilltop (1B3N, 1B2, 4A2, 12A, and 13A), the Central Hillside (9A), and the Southern Area (HA) that had received substantial wind abrasion after their use, as noted by Harriet Blitzer in 1979. It appears that in the Hilltop area sand has alternately covered and been eroded from the Minoan layers (J. W. Shaw 1977b: 206 n. 16). It is in the southern portion of the site, therfore, where the most complete depositional record exists. Sand also occurs in Geometric levels; sample K80A/37A/4:40 of Plates 3.14 and 3.18 represents a pure deposit of aeolian sand that occurs as a lens deposited between the twelfth and tenth centuries B . C , when the site was unoccupied. It appears that a bimodal sediment (mixture of sand and muddy colluvium) both underlies and seals this pure sand lens. With the decline in use of the Greek Sanctuary complex (Temple C) sand deposition became widespread over the southern portion of the site. The sand layers associated with the Upper Temple Dump slope upward towards the sea (J. W. Shaw 1981a: 228). This could be explained by a backbeach dune occupying the area between the temple and the shore at that time. The 2- to 3-m-thick sand deposit covering the Greek Sanctuary (J. W. Shaw 1979a: 162) prior to excavation accumulated there primarily after the last occupational phase of the Greek Sanctuary (first century after Christ). It overlies a mixture of windblown sand and colluvial mud and includes two sub-units: a lower sand (Pl. 3.16, "K77A/SCARP LOWER") and an upper deposit (Pl. 3.14, "K077A/SCARP UP"), the latter containing Second World War artifacts deposited within the past fifty years. Together they represent the accumulation of windblown sand on an erosional surface or redeposited sand that has lain over the site for about two thousand years. The distribution of sand in the Minoan, Archaic, and Classical/Hellenistic strata of Kommos is thus complex and discontinuous. Only in the case of pure sand sediments (Pl. 3.14) is it reasonably certain that we are dealing with primary aeolian deposits and not reworked colluvium. Although the number of sediment analyses is insufficient to detail the history of
Holocene Sea-level Changes
71
pre-Hellenistic sand deposition, clearly more sand-sized sediment appears in the earliest levels of Kommos and in its post-LM III levels than in other strata. These are periods of little activity at the site. One explanation for why this sand was deposited over the site, and probably across all the survey area, primarily in a single event dating to post-Hellenistic time, is offered by the history of sea-level change in the area during the late Holocene, as considered in the next section. But a related question exists: Why did sand deposition only begin in Early Minoan times (there is no sign of aeolian sand in the paleosols) and why is it so variable over the site? The following explanation is offered: sand influx over the area of Kommos was, during the second and first millennia B.C., relatively constant and likely somewhat greater than its present rate of deposition. (The reasoning behind a higher rate is linked to sea-level change, discussed in Section 4.) What we see in the stratigraphy of Kommos is not just a record of sand influx by wind deposition but the net result of that factor plus fluctuation in the human activities that tend to remove sand from the site. These would likely include some purposeful removal, but far more important would be the daily human activities of building, movement of people and domestic animals, clearing of ruderal vegetation, etc., which all would tend to foster erosion and mixing of sediment from the hilltop and slope, and deposition in the sea (regardless of its precise level). The present excavations have restored much of the original topography to Kommos, although without the solid terrace and structural foundations that would have further channeled and directed run-off from heavy rainstorms. Nevertheless, one can now observe how, after a sudden thundershower, significant quantities of sediment are carried down the relatively steep slope of Kommos and out to the beach. Several centuries of continuous occupation, including attempts to manage slopewash in the confines of the settlement, would have prevented much accumulation of pure aeolian sand.
4. Holocene Sea-level Changes During the first half of the Holocene Epoch, the coastal zone, a major geomorphic element of the present-day Kommos region, was absent. Although Crete was specifically excluded from consideration in a review of Late Quaternary coastlines in the Aegean (Van Andel and Shackleton 1982) owing to its tectonic instability, there is no doubt that the Mesara Gulf, along with those of Chania and Kissamou, were large expanses of sub-aerial continental shelf during the Pleniglacial (maximum ca. 19,000 years B.P.). We are concerned here with reconstructing the details of a very local, relative, sea-level rise curve for the southwestern Mesara coast over the last 4,000 years, for the position of sea level relative to the shoreline of Kommos had both geological and cultural effects. Because, over the past two decades, much work on Mediterranean Holocene sea-level changes has focused on Crete, we are more cognizant of its large neotectonic component and, correspondingly, less inclined to extrapolate results from each study to other parts of
72
Physical Geology
the island, let alone from the rest of the southern Aegean to Crete (Pirazzoli 1980). Evidence for major tectonic coastal change in historic times was, in fact, noted more than a century ago at the Graeco-Roman port of Phalasarna by Captain T. A. B. Spratt (1865) during his exploration of Crete. The more recent studies have gone beyond description to explanation and present rather detailed models of neotectonic activity. These will form the basis for the reconstruction of changes at Kommos.
Present Theories Several general categories of forces and processes contribute to determining the position of a shoreline at any coastal locale, and, as the the marine/terrestrial boundary varies geographically through time, some geological record of its position may be recorded in the sediments of the coastal zone. The study of those sedimentary records, usually through coring, offers the best chance to reconstruct coastal environments at a scale comparable to that of the archaeologist interested in regional subsistence, trade, and settlement patterns (Kraft et al. 1985). In the Kommos coastal region, there is an unfortunate absence of sedimentary sequences in the nearshore zone; the record apparently is limited to only a few decimeters of mobile sand overlying a submerged bedrock surface. This situation is common around much of the rocky Cretan coastline, and, therefore, many geological studies have had to rely on other indicators of past sea levels, notably the altitudes of structures in coastal archaeological sites (Hafemann 1965; Flemming et al. 1973) and the altitudes of radiocarbon-datable solution notches. Use of the latter, which are simply horizontal indentations or undercuttings in limestone bedrock caused by corrosion at the land/water interface, was pioneered on Crete by Laborel et al. (1978) who were among the first to collect fossil marine intertidal organisms from solution notches and date them by radiocarbon analysis. With absolute dates and altitudes above (or below) present sea level for a number of coastal localities, one may begin to see patterns in the vertical displacement of areas that reflect the fault blocks believed to form the island of Crete. Pirazzoli et al. 1982 (refining the data of Thommeret et al. 1981) propose the existence of a crustal block including all of west Crete and bounded on the east by an inferred fault trending south from Rethymnon. (This block does not include the western Mesara.) According to the hypothesized reconstruction of Pirazzoli and his colleagues, the crustal block experienced a series of ten rapid, small-scale tectonic subsidences, ranging from 10 to 25 cm of vertical displacement each time, between 4,000 and 1,700 radiocarbon years ago. Each of those inferred "sinkings" of the block (relatively to sea level) was separated by periods of tectonic quiescence lasting about 250 years. These ten events thus produced the well-known notch series visible along the coast of western Crete (illustrated, for example, in Thommeret et al. 1981, fig. 4).
Holocene Sea-level Changes
73
Then, approximately 1,530 ± 40 radiocarbon years before present (between A. D. 533 and 601 at a 66% confidence level, according to the correction tables of Stuiver and Pearson 1993), a larger block, extending farther to the east, was uplifted and tilted to the northeast in a single tectonic event. According to figure 1 of Pirazzoli et al. (1982) this second crustal block extends to the central Mesara, but its southern boundary is north of Kommos, roughly along the course of the Geropotamos River. Because of the northeastern tilting, maximum elevations of 8 m or more produced by this putative uplift event are centered at Elaphonisi in extreme southwest Crete and are manifest primarily along the west coast (as at Phalasarna; cf. Hadjidakis 1987). Their line of 1-m uplift passes from Rethymnon to north of Matala. However, Pirazzoli and colleagues note that tectonic movements "of an opposite direction" (i.e., subsidence) evidently have affected Matala. Their block boundary to the south and east is actually a "hinge line" of zero displacement and does not correspond necessarily to any of the mapped or inferred faults of the western Mesara. I doubt that the boundary is there, since differential uplift along an east-west fault in the Geropotamos Valley would have changed the drainage pattern to some degree, possibly visible in the aqueduct at Gortyn. The work of Pirazzoli et al. (1982) models the geological and radiometric dating evidence in western Crete but seems inapplicable to the immediate vicinity of Kommos. The soft Neogene marls and limestones of the southwestern Mesara coast are unsuitable for the development of marine solution notches, and, moreover, the differential weathering of these strata in the surf zone tends to mimic the appearance of solution notch sets. We are left to consider two other general sources of evidence before turning to Kommos itself: the possible correlation of major earthquakes with tectonic uplift and the archaeological evidence for sea-level change at Matala. In Pirazzoli (1986: table 1) appears a compilation of destructive earthquakes affecting ancient Crete between A.D. 344 and 553. One or more of those is believed to have produced the uplift identified in western Crete. Besides the well-known personal account of A. Evans (1928: 315ff.), there are several other references to changes of sea level in Crete as a result of earthquakes. The Graeco-Roman author Philostratus records (in Apollonius of Tyana, IV, 34,4) a "great earthquake" that struck Crete in A.D. 66, which caused the sea to withdraw seven stadia from the land (Spanakis 1983: 228). Davaras (1976: 82) also mentions catastrophic earthquakes ravaging Crete in A.D. 251 and 375. In his compendium of 764 earthquakes of Richter magnitude greater than 5.5 that struck the north-central Mediterranean between A.D. 1092 and 1976, Galanopoulos (1977) records sixteen within a 50-km radius of Kommos. Data concerning them are presented in Table 3.6. Numbers 1, 3, 5, and 14 occurred on land, to the northeast and east of Kommos; the rest were submarine earthquakes in the Mesara Gulf or the Libyan Sea. All of those would have been felt by residents of Pitsidia, and that of May 1959 must have caused some damage (its
74
Physical Geology Table 3.6. Earthquakes of magnitude greater than 5.5 that occurred within 50 km of Kommos during the period 1902-76 (data from Galanopoulos 1977: table 1). Ordinal depth of shock categories are S: shallow (0-65 km) and I: intermediate (65-300 km).
Date 1. 1 August 1923 2. 6 March 1930 3. 2 January 1937 4. 24 July 1948 5. 14 May 1959 6. 4 March 1963 7. 8 April 1964 8. 9 April 1965 9. 25 August 1965 10. 15 September 1968 11. 18 September 1968 12. 25 December 1968 13. 29 April 1972 14. 5 November 1972 15. 24 December 1973 16. 7 April 1974
Location 35 0 N, 25Έ 35 0 N, 24.5Έ 35°N, 25Έ 35.2°N, 24.4Έ 35.1 0 N, 24.9Έ 34.9°N, 25.2Έ 35.0 0 N, 24.3Έ 35. F N , 24.30E 34.7°N, 25.1Έ 34.7°N, 25.0Έ 34.7 0 N, 25.0Έ 35.O0N, 24.3Έ 34.8°N, 24.7Έ 35.0 0 N, 24.8Έ 34.8 0 N, 24.70E 34.7 0 N, 24.70E
Depth
Richter Magnitude
150 100 15 100 6 39 64 39 10 17 30 58 48 32 48 38
6.75 6 5.5 6.5 6.5 5.5 5.5 6 5.5 5.75 5.5 6 5.75 6 5.5 6
I I S I S S S S S S S S S S S S
km km km km km km km km km km km km km km km km
Mercalli intensity was IX), but none was catastrophic. The submarine earthquakes in the Mesara Gulf, especially number 2, would have generated small seismic sea waves along the Mesara coastline. Bousquet and Pechoux (1977) point out how the excavations of Pernier and Banti at Phaistos showed three major rebuilding phases associated with the First or Early Palace. The Italian excavators interpreted the destructions of the second (around 1750 B.C.) and the third (around 1650 B.C.) as resulting from catastrophic earthquakes. Bousquet and Pechoux also present a compilation of earthquakes affecting Greece, among which those felt on Crete date to 368, 267, and 265 B.C. and to A.D. 55, 66, 110, 365, and 448. Most are estimated to have been of Mercalli intensity VIII, although this would be based only on a contemporary de scription of the damage contained in the ancient sources. Given that in the eastern Mediterranean major intermediate-depth earthquakes occur with a predictable return interval averaging one per 150 years (Ambraseys 1978: 204), the strong but circumstantial evidence for major earthquake damage at Kommos is supportable. Space 25 in Trench 28B shows catastrophic collapse of an overlying upper story due to an earth quake (J. W. Shaw 1980a: 244). J. W. Shaw (1982a: 172) also sees the effect of a major earth quake in the upslope tilt of the southern wall of Space 44, west of Space 25, and believes the MM III phase was terminated at Kommos by a major earthquake (J. W. Shaw 1981a: 244).
Holocene Sea-level Changes
75
However, none of the definitive signs of archaeoseismic destruction, particularly articulated wall tumble from lateral ground motion (Karcz and Kafri 1978: 251) have been excavated. And while there are one major and several minor faults passing within a few kilometers of the site, none show any geomorphological sign of motion within the past 4,000 years. South of the Kommos-Matala area there are no published geological reports of sea-level indicators around Cape Lithinos that might bear on the question of recent tectonism in the area. East of there, Blackman and Branigan (1975: 25) could find no trace of submerged ancient ruins at Kaloi Limenes: "Indeed the indications here and at Lasaia . . . are of only very slight net change in sea level between the Graeco-Roman period and the present; this of course does not exclude the possibility of, say, earth movements of roughly equal magnitude up and down, or a rise in sea level plus coastal uplift in the intervening period."
The Evidence from Matala Tombs near sea level along the north side of Matala Bay have been broken into car-sized fragments by mass wasting and wave action. A distinct 10-cm-deep solution notch has formed (oblique to the bedding) in the largest tomb fragment, indicating that it and sea level have been in the same relative position for at least the past few centuries. Along the south shore of the embayment the maximum submergence of all tomb floors we investigated appears to be about 2 m, supporting A. Evans's (1928: 87) and Blackman's (1973) estimates of 1.8 to 2 m of relative sea-level change at Matala. I. F. Sanders (1982: 181f.) presents the archaeologically orthodox interpretation of the sealevel record at Matala: the probable Early Roman rock-cut tombs are submerged by ca. 2 m, but a slipway on the south side of the bay is at the correct functional height above present sea level. If the latter feature dates from the Classical or Hellenistic age, then two tectonic events are recorded; the shoreline around Matala fell to —2 m between ca. 300 B.C. and A.D. 100, the tombs were cut, and then sea level rose again to almost the same position it had held in the Classical/Hellenistic age. Alternatively, the land might have subsided by about the same amount it had risen, to produce the same relationships observed today. Blackman (1973) discusses his observations on the Matala shipshed-slipway complex, termed a neosoikos, which has a 12° slope to its floor, as measured by him in 1971. (It is now under the concrete floor of a restaurant and inaccessible to further study.) He addresses the two major questions of the shipshed's age and the position of contemporary sea level. The shipshed, he decides, is not datable on form alone and could be of any date between Classical and Roman times. He notes that the adjacent rock-cut chambers have entrances now submerged to —1.5 m and floors now at - 1 . 8 m. Blackman believes the chambers to date from the first to second century after Christ, indicating a relative rise in sea level of over 2 m since that period. He (Blackman 1973: 21) appreciates the problem of dating the neosoikos before a single 2-m land subsidence, because then
76
Physical Geology
its length would have been over 45 m and "It would seem to have been built when sea level was very similar to what it is now." If so, and if his dating of the tombs is correct, then there was an approximate 2-m elevation in land late in the first century B.C., and a subsequent 2-m subsidence some time after the second century after Christ. I have no other relevant observations on sea level at Matala and will assume Blackman's interpretation of the archaeological evidence is definitive. However, the submerged feature (the "Plaka") in the central nearshore zone of the bay is definitely artificial and not a natural beachrock deposit. A sample of the rock forming the Plaka was collected, thin sectioned, and compared with a sample of beachrock from the swash zone fronting the Kommos excavation. The Kommos beachrock exhibits multiple distinct carbonate cementation episodes (described above), which is not atypical for beachrock. The Plaka sample, on the other hand, shows (1) pockets of fine sand (both biogenic carbonate and terrigenous), sharply divided from a non-sand matrix; (2) extremely poor grain sorting; (3) no packing of grains, with many "floating" free in the cement matrix; (4) penetration of uncemented grain edges into cement voids; (5) a grumose, micritic cement texture; and (6) cement voids not associated with grain contacts. All these observations point to an artificial lime cement rather than natural beachrock and its most likely date would be Roman. That it is now submerged beneath half a meter of water indicates little about local sea level in the absence of information about the nature of the substrate (rock or sand) or about the function of the Plaka.
Synthesis: Sea-level Rise and Sediment Deposition at Kommos When discussing sea-level changes and paleogeographic reconstructions from an archaeological standpoint, it is important to contrast eustatic rise, which is a reflection of the melting of the polar ice sheets at the end of the last glaciation, with local relative rise (or fall), which only reflects the net contribution of many coastal processes, of different rates and magnitudes (including eustatic rise), to the position of sea level at a precisely-delimited stretch of coastline, such as the western Mesara. All those who have considered the question of sea-level rise on Crete agree that the archaeological evidence from Matala, interpreted in the most parsimonious manner, indicates a zero net local sea-level change there over the past two thousand years. Flemming (1978: fig. 9) shows computer plots of total coastal displacement and rate of displacement around Crete, assuming for Matala first a net change of zero and then one of —2 m. The plot based on zero net change shows more convincing general trends and fewer residuals, which, as Flemming (1978: 429-30) states, " . . . does not actually prove that the net displacement of Matala has been near zero for the last 2000 [years], but . . . indicates a high probability." A basic assumption underlying the following synthesis is that a local uplift of the shoreline approximating 2 m of vertical displacement affected the coastline at least from Matala to Kommos in the first century B.C., this uplift possibly being associated with local (i.e., south-
Holocene Sea-level Changes
71
central Crete) faulting or tectonism. Moreover, the Matala-Kommos area is assumed to have suffered no major tectonic coastal displacement during the preceeding two thosuand years and none subsequent to the event just hypothesized. Unfortunately, no further evidence for this can be offered here, and, if this assumption is rejected, then the synthesis must also be. A relative sea-level-rise curve for the Kommos region is presented here (Pl. 3.19), covering the period from mid-Holocene time (3,000 B.C.) to the present. No such curve exists for the western Mesara coast, although Pirazzoli et al. (1982) have proposed one for the western half of the island. As a well-known locus of tectonic activity during this time period, and lacking the alluvial embayments where complete sedimentary records might be cored and dated, Crete is not well suited for such an undertaking. Two options are available in this situation: to rely on sea-level data from other parts of the Aegean or to explain the known geological and archaeological facts in terms of a sea-level rise that fits them. Neither is an adequate approach, because in the former one is relying on data that are almost certainly incorrect for Kommos and in the latter approach the argument is very ad hoc and borders on circularity. Nevertheless, I believe it is of more interest to pursue the ad hoc approach, as it has the virtue of generating new perspectives on sea-level change for south-central Crete. The narrative history of late Holocene sea-level rise that follows is graphically summarized in Plate 3.20; it explains several fundamental geological and archaeological observations, for example, at Kommos: (1) the apparent absence of aeolian sand in the coastal region prior to the late third millennium B . C ; (2) the choice of that location for a major Minoan port facility in Middle Minoan time; (3) modifications to the floor levels of seaside buildings in Late Minoan time and postabandonment destruction of the western edge of Buildings J and K; (4) the continuous sand deposition throughout the second and first millennia B.C.; (5) a major event of aeolian sand deposition in post-Hellenistic time; and (6) the distribution and age of beachrock west of the site and at Matala: (1) the shipshed at present sea level and (2) the submerged tombs. A key geomorphologic element in this proposed reconstruction is the level bedrock platform forming the nearshore bottom from Kommos (Pl. 3.20b) northward to at least the Geropotamos River mouth. Without detailed bathymetric charts its exact morphology remains unknown, but, as mentioned previously, it was noted from Kommos offshore to the position of Papadoplaka, which marks its western edge. Beyond that rock reef, the bottom drops off immediately from 6 to 7 m depth to greater than 10 m. Irrespective of whether this nearshore platform is a classic wave-cut terrace or whether it is due to structural control (horizontal
78
Physical Geology
Neogene bedding), its sediment cover of pure sand is very thin and mobile, being carried southward by a southward coastal current. (Pl. 3.20a illustrates the contemporary coastal morphology of Kommos and "Papadoplaka.") With an average depth in the Kommos area of 6 to 7 m, and indications of similar depths along its length northward, this bedrock platform could not have functioned as a conduit of nearshore sediment transport during the last eustatic low stand of sea level. Throughout the late Pleistocene and early Holocene Epochs, alluvium from the Geropotamos River would have been deposited directly into deeper waters of the Mesara Gulf, and, as a result, beach development may have been minimal along the succession of now-drowned paleoshorelines of the early Holocene transgression. When the eustatic rise reached - 6 to —7 m, the entire bedrock platform would have been flooded almost immediately and a shallow nearshore zone created for transport of sediment as well as a relatively wide, sandy beach. Sand from this beach would have been available for wind transport and deposition inland from the shoreline of the time. A time for the initial flooding of the platform could be assumed, in the absence of other evidence, to be in the mid-third millennium B.C. (Pl. 3.20c). From that time on there was a variable quantity of beach sand available for inland transport by the prevailing west and northwest winds. Sand would have begun to cover the soils that had been developing on the Kalamaki, Pitsidia, and Matala alluvial plains and uplands since the end of the last ice age. Sand also began to accumulate at Kommos contemporary with the earliest traces of human activities there. With the continuing rise of sea level, the bedrock outcrop called "Papadoplaka" became an offshore islet and would have been joined to the coast by a sand tombolo deposited in the lee of the dominant wave fetch from the northwest, west, and southwest quadrants (Pl. 3.2Od). No similar rock reefs exist in the nearshore zone of the western Mesara, and it would have attracted anyone involved with watercraft as a natural shelter and anchorage. By LM I time (mid-second millennium B.C.) the water depth between Kommos and Papadoplaka still was only 3 to 4 m and the tombolo still in place, though perhaps more liable to erosion during major storms. As the local sea-level rise continued at a rapid rate from —5 to —2m during the second millennium, the tombolo "drowned" and was dissipated by wave energy. No trace remains and the only very indirect evidence for its postulated existence are Minoan stone anchors found midway between Kommos and Papadoplaka in 1983 and 1985. However, an anchor could be lost in several meters of water as well as through burial in a tombolo beach, so this is not given much weight. Besides the inexorable erosion of the natural harbor formed by the tombolo, inhabitants at Kommos were faced with rising dampness in their low-lying structures owing to the concomitant rise of the local fresh water table along with sea level. The floor level of Building J was raised a meter in LM IIIA2 (J. W. Shaw 1984a: 274, n. 41); being among the lowest buildings
Holocene Sea-level Changes
79
at the site, it would have been among the first to experience the rising dampness. Building T's floor also was raised (J. W. Shaw 1984a: 272), probably for the same reason. The upper levels of Building J indicate abandonment in LM IHB time, ca. 1300 B.C. Sea level then is postulated to have been between 2 and 3 m below its present level and the water between Kommos and Papadoplaka some 4 to 5 m deep, enough to have permanently eradicated the tombolo (Pl. 3.2Oe). As the beach zone migrated eastward along with sea level during the early first millennium B.C., larger quantities of sand began to be deposited over the southern part of Kommos, building up 0.1 to 0.4 m east of the abandonded Building J and showing characteristic aeolian cross-bed sets. With the re-occupation of the site the partially exposed walls of the large Minoan structures would have offered obvious foundations for the Protogeometric temple excavated there. In the Kommos-Matala area the relative rise of sea level is assumed to have continued rapidly up to about the first century B.C., when it was near its present position and when the neosoikos was constructed at Matala. By that time Kommos apparently had ceased to be a locus of settlement for maritime trade because its natural tombolo anchorage, the only one on the Mesara coast north of Matala, had vanished (Pl. 3.2Of). To have the local relative sea level as high as suggested by the Matala shipshed at such an early time (assumed to be the late first millennium B . C ) strongly implies that there is a tectonic component to the net change, but no attempt will be made here to define it further. An observable effect of the inferred high stand was the erosion by storm waves of the seaward portions of the low-lying structures at Kommos, notably Buildings J and N. It appears that even in the recent past major storms have affected those structures, and a run-up of large waves to two to three m above mean sea level and a horizontal distance of some 50 m over a low-angled bedrock foreshore is entirely feasible. Given a floor level at +2.7 m in Building J, it is almost impossible to imagine a first-millennium-B.c. sea level at Kommos higher than the present one, as Pirazzoli et al. (1982) postulate for the west end of Crete. In that case Building J would have been eroded away completely. Based on the Matala evidence, a rapid tectonic uplift of the coastal region by approximately 2 m vertically is hypothesized to have occurred sometime between 100 B.C. and A.D. 200. At Matala, with its steep shoreline, this allowed the cutting of tombs only in the newly exposed bedrock of the north cliff. However, from Kommos northward, where the nearshore zone is more nearly level, a wide expanse of sand would have been exposed to wind action in a very short time (Pl. 3.2Og). It is proposed that this became the new source of the unique aeolian sand blanket that was soon deposited throughout the western half of the Kommos survey area during the first few centuries of this era. Before it disappeared from the vicinity of Kommos, the wide, sand beach was the formation site of an extensive deposit of beachrock in the subsurface zone, the deposit that today lies submerged in one to two m of water off the site. From ca. A.D. 200 it is only necessary to allow a reasonable eustatic rise of some 0.1 m/cen-
80
Physical Geology
tury to bring sea level to its modern position, which happens to correspond closely to its position in early Roman time. The beach at Kommos would be progressively diminished in width, and the beachrock formed during the first few centuries after Christ would gradually be exposed in the nearshore zone. That deposit was found to contain no potsherds because no great use was being made of Kommos during the time of its formation. More sand continues to be deposited over the survey area but nothing like the volume deposited during the first few centuries after Christ. Sand continues to be deposited inland of the modern beach zone, but several instances were noted during the geological survey of muddy colluvium from the surrounding hillslopes actually covering pure sand deposits. The depositional event of ca. 1,800 years ago could not now be repeated because its unique source was submerged by the subsequent transgression. The hypothetical reconstruction of coastal evolution shown in Plate 3.20 has little to recommend it other than its explanation of many field observations, both archaeological and geological, while making minimal demands on the basic facts of late Holocene eustatic sea-level rise. The local curve of sea-level rise (Pl. 3.19) cannot in any sense be considered an attempt at constructing a "true" curve. In particular the assumptions of initial sea-level position at 4,500 B.p. and its very rapid rise during the second millennium B.C. are purely ad hoc; there is some circumstantial evidence for tectonic depression of the area during the second and first millennia, prior to the sudden uplift identified from the features at Matala. The discovery of in situ, datable intertidal fossil organisms attached to the base of Papadoplaka would be one test of the reconstruction presented here; another would be the direct (radiocarbon) or indirect (included sherds) dating of the beachrock deposit just west of Kommos. Yet another test might focus on changes in the lower channel of the Geropotamos River (assuming the uplift extended that far to the north), and the cemented beach deposits noted above (Section 1, under "Geomorphology of the Western Mesara") might reflect such uplift. However, in the absence of more field work, there is no simpler curve conceivable that explains as much.
5. Inorganic Raw Materials at Kommos In this section the mineralogical composition and derivation of several categories of inorganic raw materials found in the Kommos excavations are described.
Building Stone Two rock types are most common in the constructions at Kommos: a white, fine-grained limestone and a grey, fossiliferous limestone. The former represents the Ambelouzos Formation, and the latter represents the calcarenitic "poros" limestone (marine biosparite, M 4 m) of the Varvara Formation, located topographically at the top of the Matala headland and the Pitsidia Ridge overlying the Ambelouzos Formation (M3C).
Inorganic Raw Materials
81
Good examples of the use of the former rock type are offered by the large irregular slabs of white limestone as floor paving in Space N i l , Trench IB (J. W. Shaw 1977b: 212, Space 11) and the broken, canted limestone slabs, probably from an upper story, found in Space 25 (Trench 28B) above a Middle Minoan storeroom (J. W. Shaw 1980a: pi. 55a). Petrographic examination of a thin section of these slabs showed the fine sand- and silt-sized grains to consist of one-third siliceous sponge spicules, one-third angular terrigenous rock fragments, and one-third foraminifera tests, floating in a fine-grained (micritic) carbonate cement. It would have been a simple operation to pry slabs of this rock from the bedrock scarp below the Hilltop, and an unlimited supply was also available from Vigles Hill just a few hundred meters away. The use of "poros" limestone for the orthostat blocks of the Late Minoan public structures at Kommos (J. W. Shaw 1982a: 182) is understandable from the viewpoint of its physical properties; it is the only rock unit in the Kommos area that can be found in strata thicker than 0.3 to 0.4 m and the only one with reasonably homogeneous and isotropic composition over distances of more than a meter horizontally within any one stratum. The rock type represented by the "poros" limestone is a "sandy biosparite" in Folk's (1974:171) carbonate texture terminology: in thin section it shows a matrix predominantly composed of sparry calcite cement, with twenty to thirty percent fine quartz sand or coarse silt. Blocks such as those in the north wall of Room 16, Building T, could have come either from quarrying at the south end of the Pitsidia Ridge or from the top of the Matala headland, perhaps by a short over-water trip directly to the site. There is no evident use of Pleistocene eolianite for ashlar blocks at Kommos. This rock is known as ammoudhia and Bronze Age quarries of it have been noted at many coastal locations around the island (J. W. Shaw 1973; Soles 1983). The fossiliferous limestone used was perhaps a compromise between its less desirable characteristics of working and its availability in the immediate vicinity. A small percentage, perhaps one to two percent, of Middle Minoan and Late Minoan wall stones in structures of the Central Hillside were built of well-rounded, boulder-sized beachrock slabs. Although not sampled and analyzed petrographically, the macroscopic appearance of this beachrock differed from the deposit exposed in the nearshore zone, having wellrounded granule- to pebble-sized rock fragments as its primary grains. Beachrock slabs of this size are not common in the present shore zone, suggesting an earlier phase of beachrock development might underlie the visible deposit.
Pottery Clay Section 3 above (under "Geology and Geomorphology") includes a discussion of clay bedrock types in the vicinity of Kommos from the standpoint of their grain-size distributions. Both kouskouras and lepidha clay bedrock types in and around Kommos are secondary clays, redeposited as sediments in some location other than where they were formed, as is
82
Physical Geology
true of many of the Neogene clay deposits of Crete (e.g., the Ierapetra region; Gifford and Myer 1984: 119). Clay-rich (in the sense of clay-mineral rich) bedrock types were noted at four locations during the Kommos survey, but one of the most accessible to Kommos is to be found about 200 m south of the site, at the base of the western slope of Vigles Hill. Grain-size analysis was performed on a sample of this clay, which proved to be very similar—almost identical— in its size distribution to the sample of lepidha discussed in Section 3 under "Geology and Geomorphology" (see Pl. 3.18, "80S/CLAY NR GEORGE" and "80A/35A2/ 8:119LEPI"). In fact only one other sample had a higher percentage of clay-size mineral grains (Pl. 3.18, "77A/13A/3:29 CLAY"), but it was not clear if this small deposit of clay had been specifically collected for potting or just represents a disturbed portion of the local bedrock. An attempt was made to characterize and compare the two sediment samples having the highest percentage of clay-size mineral grains (Pl. 3.18) by X-ray diffraction analysis (XRDA). Both of the samples were friable and were disaggregated by standard techniques to isolate clay size fractions for XRDA. One-degree 29/minute patterns were run at 0.57inch of chart paper using Cu K„ radiation and a time constant of 0.5 seconds. Plate 3.21 shows the portion of the chart between 4° and 28° 2Θ for these two samples. The similarities are apparent in peak intensities of several clay mineral species, including kaolinite (12.9° 20), the illite group (9.25° 20), and the smectite group (6.4° 2Θ). The third sample plotted in Plate 3.21 is the XRDA pattern of the clay-size fraction from the B horizon of the survey site 75 paleosol, discussed above. Similarities in clay mineral species are also apparent, although the peak intensities are significantly greater in the soil-derived clay minerals.
6. Inorganic Processed Materials Several groups of inorganic material that have been processed to some degree in preparation for their particular application are considered here: lime and clay plasters (and their mix tures), pigments, and miscellaneous inorganic materials. Processing included both physical mixing of raw materials and their chemical modification by various pyrotechnological processes. For the locations of the materials consult Plates 2.13 (Hilltop = HT), 2.15 (Central Hillside = CH), and 2.17 (Southern Area = SA).
Lime Plasters The first group includes lime plasters, which are the most positively characterized of the sev eral to be described. Six samples, subjected to X-ray diffraction analysis (XRDA) of ca. 100 mg
Inorganic Processed Materials
83
of powder under the same laboratory conditions as noted above for the clay samples, are listed below, with excavators' descriptions. 1. 9A/36, CH, Space 7b, MM III. Plaster from mortar S 137.
4. 28B/9, CH, above Space 25, MM III. Plaster in pot C 1913.
2. 20B/24, SA, sounding east of Round Building
5. 9A, CH. Traces of plaster on LM I wall.
3. 19A/28, CH, sounding (area shown in J. W. Shaw 1979: pi. 56a), MM III. Floor plaster.
6. Nineteenth century after Christ (lime plaster [asvesti] from the small stone structure approximately 400 m east of Kommos).
All of the first five samples are composed of calcite (calcium carbonate) plus silica quartz in an approximate ratio of 2:1 to 3:1. No gypsum (calcium sulfate) or clay minerals (smectite, kaolinite, or illite) are present. Such a pure lime mortar could have been prepared by burning crushed shell or selected local bedrock (the biosparite); the quartz percentage present in all samples makes the latter source more likely. Lime plasters evidently were used in Middle Minoan time for both wall and floor coatings. The sample from the LM wall in Trench 9A may be a reused Middle Minoan building stone, with its original interior plastered surface turned to the outside when rebuilt into the later wall. Sample 6 was analyzed for comparison with the Kommos samples. It too shows calcite and quartz as the predominant minerals but includes a small admixture of gypsum (hydrated calcium sulfate). The overall XRDA pattern is quite distinctive from the Minoan samples. Some comparisons can be made between the Kommos Middle Minoan wall plasters and the wall plasters from Early Minoan structures at Fournou Korifi, studied by Cameron (1972: 312ff.). While the analytical techniques employed were fundamentally different, the Kommos samples' estimated 2:1 to 3:1 calcite-silica ratio is more similar to the Early Minoan Fournou Korifi samples than to the EM II sample from Knossos cited by Cameron (1:1) and certainly to the LM II sample from Knossos (50:1). This adds some support to the idea of regional traditions or practices in Minoan builders mentioned by Cameron. Regarding plaster analyses, the standard reference concerning Minoan plasters has been Heaton (1911). He notes a high percentage of silica (SiO2) and alumina (Al2O3) in many of his plaster samples, which to him suggested a mixture of two components, lime and zeolitic clay. Heaton's work is now outdated. All his analyses were done by wet-chemical techniques, which did not allow the identification of the actual mineral species present in a sample, but only its elemental composition (commonly calculated and expressed as the oxide equivalent). X-ray diffraction analysis was not applied to the identification of minerals until the 1930s. Also, the family of minerals known as zeolites are hydrous silicates of aluminum, with sodium and calcium as the important bases. However, they are not clay minerals in modern classifications schemes and could not have been mixed with lime to produce any kind of known plaster. With anything less than a massive percentage of zeolite, the proper-
84
Physical Geology
ties of the resulting mixture w o u l d be little different from those of p u r e lime plaster. Lastly, zeolite minerals are relatively rare.
Clay Plasters The second g r o u p comprises clay plasters; it is n o t as h o m o g e n e o u s as the lime plasters in t e r m s of the XRDA p a t t e r n s but h a s the c o m m o n characteristic of t h e total absence of calcite in all s a m p l e s , i.e., just the opposite of the lime plasters.
1. 29A1/68, SA, Temple B, eighth to sixth century B.C. The XRDA pattern shows very intense peaks that are in good agreement with the clay mineral kaolinite (ASTM standard pattern 14164), with only a trace of quartz present. The crystalline structure is so well defined in the sample that it possibly represents raw material, not (yet) processed or treated in any way. 2. 36B/1, SA, above Gallery 2 of Building P, Hellenistic. This pattern is comparable to the preceding one; the only certainly identifiable mineral is kaolinite. The pigment providing the red coloration (1OR 4/9 dry Munsell color) is not identifiable as any mineral on the XRDA pattern. This suggests that it may be a microcrystalline variety of hematite (Fe 2 O 3 , dehydrated iron oxide), which is a common component of red ochres. It is possible that very small ochre deposits, unmappable at regional map scales, occur in the metamorphic terrains of Crete. Alternatively, certain of the local red paleosols may have been levigated to concentrate the clay-sized hematite component for use as a pigment. 3. 44A/14, SA, exterior area south of temples, eighth century B.C. The XRDA pattern is identical to the previous samples in its primary component (kaolinite); it also shows a trace of hematite (ASTM 13-534). It is probably not cinnabar. 4. 42A/40, SA, exterior court area south of tem-
ples, eighth century B.C. This sample is the white portion (unpigmented) of a wall plaster fragment. It contains more quartz than the previous three samples but is otherwise identical. 5. 42A/40, SA, exterior court area south of temples, eighth century B.C. The red portion from the same fragment as the previous sample. Clay mineralogy is again kaolinite, but there are no traces of a pigment mineral on the pattern. 6. 43A/26, SA, exterior area west of temples, eighth century B.C. Again, a kaolinitic clay with minor amounts of quartz. No trace of a mineral pigment, though it may be the amorphous limonite (see below). 7. 43A/45, SA, exterior area west of temples, eighth century B.C. Well-crystallized kaolinite with strong quartz lines in the XRDA pattern. Either quartz-rich sediment has been added as a binder to this clay plaster or it may have been a component of the pigment added (ochre or terra rossa). 8. 43A/40, SA, exterior area west of temples, eighth century B . C The sample analyzed has a pinkish cast (7.5 YR 8/4); it is identical in its XRDA pattern to sample 1 above. 9. 44A/9, SA, exterior area south of temples, eighth/seventh century B . C Again the primary mineral is kaolinite and again poorly crystallized and with no trace of quartz or pigment minerals.
C l a y - l i m e Plasters Two s a m p l e s containing a mixture of calcite plus a m i n o r but significant p r o p o r t i o n of a n u n k n o w n clay mineral (possibly a n o t h e r species of kaolinite) are g r o u p e d h e r e .
Inorganic Processed Materials 1. 27A/50, HT, above Space 018, LM IIIA2/B. As was true of the pure clay plasters, no trace of the pigmenting agent is discernible on the XRDA pattern. 2. 44A/25, SA, exterior area south of temples, eighth century B.C. While exhibiting a pattern
85 identical to the preceding sample, this one has the macroscopic appearance of a very soft, ferruginous marlstone. Both samples could represent utilization of that unit of the local bedrock to produce a clay plaster.
Pigments and Miscellaneous Inorganic Materials Ten samples of pigments and exotic inorganic materials were analyzed by X-ray diffraction. 1. 9A2/65, CH, Space 13, LMIIIA2/B. This sample's Munsell color is 1OR 3/6 (dry) and it is in fact a mixture of hematite (ASTM 13-534) plus a small quantity of quartz. 2. 27A1/9, HT, above Space 018, LM IIIA2/B. An unusual mixture of minium (lead oxide, ASTM 8-9), cerrusite (lead carbonate, MI 1492), and anglesite (lead sulfate, ASTM 5 577). It is chemically possible that the cerrusite was heated to produce the orange-red pigment minium according to the following reaction: 400°C 3PbCO 3
> 6CO 2 + Pb 3 O 4
It should be noted that none of the red pigment samples described previously here show any trace of minium. 3. 34A/15, SA, south of Temple C, Hellenistic. Only strong calcite peaks and minor quartz peaks are present on the XRDA pattern. It may be just a lime plaster fragment colored by Iimonite (yellow ochre), the naturally occurring, amorphous hydrated species of iron oxide. 4. 35A2/95, CH, Space 35, MM III. The XRDA pattern shows a mixture of the chlorite mineral aphrosiderite (ASTM 12-243) plus an unidentified mineral of the talc family. 5. 40A/62, CH, Space 35, LMIHA. This material is a mixture of aphrosiderite plus the talc family mineral called steatite. It is likely that the metamorphic terrains of Crete contain small deposits of these two minerals, which would have been a source of the "soft stone" for small sculpted objects.
6. 44A/7, SA, exterior area south of temples, eighth century B.C. Strong calcite and quartz peaks are present, plus a number of unidentified peaks that may represent the pigment in this lime plaster. 7. 42A/25, SA, court east of temples, eighth/ seventh century B.C. A lump of specular hematite (Fe2O3) with a micaceous habit, probably destined to be ground into red pigment. 8. 43A/59, SA, exterior area west of temples, eighth century B.C. Gypsum (hydrated calcium sulfate, ASTM 6-46) is confirmed by the XRDA pattern. 9. 44A/40, SA, Building N, Space 12, LM III. The XRDA pattern conclusively matches that of the artificial pigment termed "Egyptian Blue" (ASTM 12-512), which is a calcium copper silicate. This is the same as a previously analyzed sample (K76A/2A2/4:39). Egyptian Blue has also been identified in pigment samples from the Unexplored Mansion at Knossos (Profi et al. 1976) and from the Akrotiri excavation on Thera (Profi et al. 1977). 10. 34A2/32, SA, south of Temple B (cf. J. W. Shaw 1981a: 235), seventh century B.C XRDA produced a totally amorphous pattern, although the sample appeared to be crystalline under a 1Ox hand lens. When examined under a petrographic microscope at 45Ox and crossed polarizers, the sample was found to consist of fragmental silica microfossil tests (sponge spicules and diatom frustules). Although this unusual sediment type has not been identified in the immediate vicinity of Kommos, Freudenthal (1969: 38) notes its occurrence in laminated marls of the Khairetiana Formation (up-
86 per? Miocene) in the vicinity of Khania and it has been reported from at least one other locale, near Heraklion (Frydas 1985). A reasonable ap-
Physical Geology plication imaginable for this material is the same as it was used in the historic past, as a metal polisher.
7. Acknowledgments I wish to thank Michael Parsons and Giuliana Bianco for assistance in the field survey at Kommos during various seasons and Christopher L. Hill for analytical asssistance at the Archaeometry Laboratory, University of Minnesota (Duluth).
A p p e n d i x 3.1
Grain-size Analysis and Distributions Sediment grain-size analyses were done using standard techniques of sieving of the sand fraction and pipet analysis of the silt-clay (mud) fraction (Folk 1974). The raw size fraction weights were reduced by a computer program that also calculates basic moment measures and plots a cumulative distribution graph of the weight percent of sediment in each size frac tion. This Fortran program (GRAIN.EXE) was modified and enhanced by G. Caballero (Uni versity of Minnesota Archaeometry Laboratory) from a version published by Slatt and Press (1976). Grain-size distributions of clastic sediments, both terrigenous and carbonate, are com monly assumed to approximate a Gaussian distribution of sizes. The basic statistical parame ters (moment measures) that describe a frequency distribution of sediment particles (or any other entity) are the distribution's average (mean or median), its sorting (dispersion of the distribution about the average), its symmetry (skewness), and its peakedness (kurtosis). While these summary statistics traditionally have been used to correlate sediment deposits and as a means to identifying their depositional environments, such an approach was not used in this study. Rather the information contained in the grain-size analyses was subjected to multivariate statistical analysis, following the suggestion of J. C. Davis (1970). Correspon dence analysis was applied as described above. On each of the final cumulative grain-size graphs, the abscissa represents sediment size in terms of the phi (φ) inverse logarithimic scale commonly used in sedimentology. Phi values of —1 φ to + 4 φ include the sand size range, +4 φ to +8 φ the silt size range, and + 8 φ to +14 φ the clay size range. The size distribution graph's ordinate is calibrated as the cumulative weight percent of the sample in each size fraction. In addition, X-ray diffraction analyses of clay (> 8 φ) and coarse silt (4-4.5 φ) size fractions were performed on 38 samples, again as an aid in characterization, correlation, and determi nation of depositional origin of the samples. Initial groupings of samples with similar coarse silt mineralogies were again accomplished through K-means cluster analysis of a matrix of 16
Appendix 3.2
87
variables (X-ray peak intensity heights of common mineral species) by 38 cases (sediment samples).
Appendix 3.2
The Invertebrate Fossils David S. Reese Kommos produced 470 marine invertebrates: 423 oyster (Gryphaea, Crassostrea, Ostrea) frag ments from 150 deposits (Pl. 5.4M), 32 scallops (Pecten, Chlamys; 16 deposits), 9 bivalve inter nal molds, 3 coral, 2 gastropods, and 1 echinoid fragment. These are listed by period on Ta ble 3.7. Most of the fossils are LM III (42%). Iron Age Kommos produced over 175 fossils, including at least 125 oysters. Fossil oysters, scallops, and echinoids are seen today in outcrops on the coastal cliffs at Kommos and at survey site 79 on Vigles Hill, both of the Varvara Formation. There are also outcrops of the Ambelouzos Formation to the south and southwest of the modern village of Pitsidia. Both these formations are to be dated to the Upper Miocene, seven to nine million years ago (Marcopoulos-Dicantoni 1973). Six of the oysters contain hematite: MM II—III (2), MM III (2), and LMI (2) and it is possible that they were collected to utilize this pigment. One oyster is water-worn and another worn. It is also possible that at least some of the fossils are simply eroded from the building stones and bedrock and are in the archaeological deposits by chance. Worth special mention are the large fossil collections and the unique fossils. The LMIA-III Table 3.7. Fossils from Kommos. Period
Oysters
MM I + MM I II MM II MM II—III MMIII MM III- LM I LMI LM Ι-Π LMII LM IHAl LM ΠΙΑ LM IIIA2 LM III LM IIIA2-B LM IHB
13(4) 6(2) 19(5) 107 (31) 41 (14) 41 (21) 7(4) 1 19(7) 8(6) 23 (15) 43 (16) 44 (16) 51(8)
( ) = number of deposits.
Others
Scallops 1
bivalve, small coral fragment
— — 20(5) 4(3) 1
3 bivalves, 3 small coral fragments 2 bivalves (2), small gastropod
— 1
echinoid fragment
— 1
— 1 2(2) 1
bivalve bivalve, Strombus internal cast, colonial coral bivalve
88
Physical Geology
area above Space N14, west of North House on the Hilltop, produced both a bivalve and a Strombus internal cast. A second Strombus (S 63, 29 mm long with a maximum width of 25.5 mm [Pl. 5.8]) comes from the mixed LM III and later deposit in the Southern Cliffside (4A2/33). The LM HIA Court 016 of the Oblique House on the Hilltop produced 2 oysters and 1 colonial coral. In the Central Hillside, the MM II—III fill in Room 16 produced 6 oyster fragments, 2 containing traces of hematite. The MM III Room 25 produced 11 oysters and vessels in that room yielded 4 more and 4 scallop fragments. The contemporary disturbed floor of Room 28 produced 4 oysters and 1 small bivalve. The MM III paving south of Room 28 produced 2 oysters and a coral fragment. Also in this area the MM IH-LMI Room 9 floor and rubble on it produced 11 oysters and 1 scallop and Room 52 produced 8 oysters and 2 scallops. In Space 7 the LM II dump produced 17 oysters and 2 scallops and the contemporary fill over Room 47 produced the only echinoid remain. The LM HIAl Room 7 yielded 6 oysters. Room 2 (LM IIIA2-B) produced a holed water-worn oyster (S 19, 38 by 35.5 mm with a 6.5-mm-diameter hole), possibly a pendant (see Mary K. Dabney, cat. no. 34, in Kommos 1.2). The LM IHB Room 5 produced 6 oysters. In the South, the LM IIIA2-B Building T and Road produced 16 oysters and the upper floor of Corridor 7 produced 4 oysters, 1 scallop, and 1 small bivalve. The LM IHB level above the upper floor of Building J yielded 19 oysters. As suggested from the Kommos material and the comparanda presented below, some of the Minoan fossils from Kommos are likely to be objects collected as curiosities (like the waterworn shells) or for ornamentation. For example, the Glossus from Malia, Malthi, and Khirokitia may have been collected because that shell resembles the human heart.
Comparative Data In the past, little attention has been paid to fossils from archaeological sites (Reese 1985b, 1992a: 775), although I have discussed elsewhere fossil (and recent) shark remains from sites in the Aegean and Cyprus (Reese 1984a). As N. J. Shackleton has noted for Myrtos (Shackleton 1972: 325): among the shells examined were a few fossils. At first these were assumed to be present by chance, since fossiliferous rock is close by the site. However, one cannot exclude the possibility that the occupants collected fossils as well as shells. Certainly a variety of different species in elegant condition were present. The fossils from Aegean and Cypriot Bronze Age sites are noted briefly here to better understand the Kommos collection. The MM II Lower West Court Sanctuary Complex at Phaistos yielded two fossil shells from the passage into Room LXII, apparently a hearth area
Appendix 3.2
89
(Gesell 1985: 127). The MM ΠΙΑ fill of Tomb XVIII in the Gypsades cemetery at Knossos pro duced a fossil fragment (Reese 1982a: 249). There is a fossil Glycymeris internal mold from the LM II Royal Road. The MM III-LM IA Poros tomb produced a fossil bivalve with some red paint on it and a small, worn fossil oyster (Reese 1992b). Juktas produced 8 bivalve internal molds, 2 oysters, 2 scallops, 1 gastropod, and 1 echinoid. Vathypetrou produced 1 oyster fragment. The Idean Cave yielded 1 scallop fragment. The Malia palace yielded a Glossus humanus (L.) (= Isocardia cor [L.]), a Heart cockle, suggested to be an amulet (Effenterre and Tzedakis 1977: 18:3, pi. 74d). Pyrgos produced 13 oysters, 3 bivalves, and 1 gastropod. What is probably a fossil Conus comes from Ierapetra in southern Crete (Buchholz and Ka ra georghis 1973: 49, obj. 456). Chania produced 15 oysters, 11 from pits. Fossil deer antlers are known from a Minoan shrine at Knossos (Bate 1918: 221; Jarman n.d.) and a sawn fossil bone is known from the Unexplored Mansion (Evely 1984: 245). Akrotiri on Thera produced at least 1 fossil scallop. On the Greek mainland, EH II Perachora produced 18 scallops from 5 deposits and 4 bi valves, including 1 Glycymeris. Mixed later levels produced 1 scallop, 1 Glycymeris, and 1 oys ter. Lerna yielded a Gryphaea (oyster) from a Middle Neolithic deposit, another oyster from an EH II level, and a scallop from an MH/LH I shaft grave. Nichoria produced over 300 ma rine invertebrate fossils, mainly oysters and scallops, which mistakenly were identified as food debris (Reese 1992a: 775; Sloan and Duncan 1978: 70, pi. 6-13:8, 13). One LH HIB oyster fragment has a hole and may have been an ornament (Reese 1992a: pi. 1-6). Malthi in Messenia produced 1 Glossus (called Cardium isocardium; Valmin 1938: 360, pi. XXVI 0 10; on dis play in the Kalamata Museum, no. 1211). Ayios Stephanos produced 15 oysters dating from MH II to LH IHB, 1 LH I scallop, and 5 other bivalve fragments of EH II-LH II date. In Turkey, Troy produced fossil mussels (Gejvall 1939: 4) and a fossil vertebra of a member of the Dolphin family, called "a very curious petrified bone" (Schliemann 1880: 323). There is a fossil gastropod from the Chalcolithic at Alishar Huyiik (Gries 1937: fig. 259:e635), and Early Bronze Age deposits produced 2 additional examples (Schmidt 1932: 72, fig. 86). At Gozlu KuIe, Tarsus, an EB II fossil(?) Acanthocardia was perforated at the umbo (Goldman 1956: 340, fig. 45:7). The 2000-to-1600-B.c. site of Polati southwest of Ankara produced a fossil oyster (Dilgimen 1951: 75). On Cyprus Aceramic Neolithic Rizokerpaso-Cape Andreas Kastros produced 4 Dentalium, 12 bivalves, 11 pieces of solitary coral (Cladocora debilis) with 6 from one pit, and 1 echinoid (Reese 1978: 87 and unpublished additions). Khirokitia produced a Glossus (Wilkins 1953: 439, pi. CIIL953), a Crassostrea gryphoides, a portion of a gastropod (631/c), a Strombus coronatus (161), a Glycymeris (733p), and a colonial coral (1010). There are also 2 Dentalium (Stanley Price 1976: 9; Reese 1978: 87). Middle Cypricot I Alambra produced 2 oysters. The Late Cypricot IA2-B2 (1600-1475 B.C.) Tomb 20 at Aghia lrini-Paleokastro in the northwest produced beads of Dentalium rectum (Durante 1972: 289-90, fig. 308). HaIa Sultan Tekke yielded various fossil
90
Physical Geology
bivalves: Crassostrea stentina, Spondylus gaederopus, Donax trunculus, Pitaria, Loupes (Lucina shell), Chama (Jewel box), Chlamys, and coral (Cladocora cespitosa) (Demetropoulos 1979: 136, 137, 140, 141). Maa-PalaeoL·stro produced 2 gastropods, one a Conus internal mold and the other a gastropod internal mold fragment that could be strung (Reese 1988a: 459-60, pi. A:ll-12). A fossil shell from Late Cypricot III Enkomi in the east was perforated to make a spindle whorl, pendant, or loomweight (Dikaios 1969: 92, 677, no. 1116).
C H A P T E R
4
The Modern Flora and Plant Remains from Bronze Age Deposits at Kommos C. Thomas Shay and Jennifer M. Shay, with the assistance of Katherine A. Trego and Janusz Zwiazek
1. General Introduction The Modern Flora, Vegetation, and Ethnobotany (J.M.S., C.T.S., K.A.F.) 2. Introduction 3. Methods 4. The Environments of the Kommos Area 5. The Flora of Kommos 6. Plant Communities of the Kommos Area 7. Plant Communities beyond the Survey Area 8. Human Ecology and Ethnobotany 9. Human Influences on the Landscape 10. Summary The Charcoal and Seeds from Bronze Age Deposits at Kommos (C.T.S., J.M.S., J. Z.) 11. Introduction 12. Methods 13. Inferring Past Plant Uses 14. Results 15. Environment and Ecology during the Minoan Period 16. Acknowledgments Appendix 4.1. Plant species list for Kommos area, Crete Appendix 4.2. Charred wood anatomical descriptions
91
92 Appendix Appendix Appendix Appendix Appendix
4.3. Distribution of pails (provenience units) by trench at Kommos 4.4. Distribution of pails by date at Kommos 4.5. Charcoal remains from Kommos 4.6. Charcoal remains from Bronze Age deposits at Kommos 4.7. Measurements of selected seeds from Kommos
They must have been flower lovers too, these Minoans, for the flowers that surrounded their cities, palaces and villages are repeated in their arts: frescoes, paintings and pottery show such flowers as . . . the sea daffodil with lilies, rushes and, most delicately drawn and painted; and as you visit the sites you will find the same flowers blooming still. (Huxley and Taylor 1977: 51)
1. General Introduction Plants were prominent in Minoan art but they also served Minoans in many practical ways. As part of the Kommos project, the botanical studies explore the kinds of plants available to the inhabitants of Kommos during its long occupation and the ways in which these plants were used. This ancient site also offers an excellent opportunity to study relationships between the inhabitants and the local environment. During Minoan times (MM I-LM III, ca. 2,000-1,200 B.C.), the agricultural settlement of Kommos became a trading seaport. Later, from the beginning of Proto-Geometric until Roman times (ca. 1,000 B . C . - A . D . 150), it was the focus of a sanctuary and temple complex. In addition to analyses of Kommos's rich archaeological record, its local environment and culture has been investigated through studies of the geology, soils, land use, ancient settlement patterns, and botany. This chapter is divided into two parts: an examination of the modern flora, vegetation, and ethnobotany of Kommos and the analysis of the charcoal and seeds from Bronze Age deposits there. The botany of Kommos is best appreciated within the context of Crete. Crete is the fifth largest island in the Mediterranean Sea, with an area of about 8,700 km 2 . Rugged mountains up to 2,450 m high form an arc along its east-west axis, and steep gorges, innumerable valleys, upland plains, and lowlands contribute to the diversified topography. The geology is complex: a mixture of limestone, dolomite, and metamorphic rocks which have given rise to a variety of soils. The flora of Crete, as elsewhere in the Mediterranean basin, originated from both temperate and tropical regions some eighty million years ago in the late Cretaceous period (Axelrod 1973, di Castri 1981). In more recent geological times, environmental conditions, particularly summer drought, fire, and grazing, have resulted in a flora composed of a large number of spring-flowering annuals, herbaceous perennials with underground storage organs, and drought-resistant shrubs. Crete has been isolated from the mainland for more than five million years, with the result
Introduction (Modern)
93
that ten percent of the flora is endemic or unique to the island (Barclay 1986). Unlike northern Europe, where glaciers eliminated many species, much of Crete escaped glaciation and retained its rich flora. Geologic history, diverse topography, soils, and climate have resulted in a native flora containing 1,600 species of flowering plants (Barclay 1986). Since the Neolithic, ca. 6,000 B.C. (Willetts 1977), in Crete and throughout the Mediterranean, climatic change, land clearance, fire, and overgrazing have allowed shrublands to expand (Le Houerou 1981, Naveh and Lieberman 1984). During this time Crete's flora has been enriched by numerous crops, ornamentals, and weeds introduced from elsewhere. The microcosm of Kommos typifies the rich Cretan flora and the factors that shaped it.
The Modern Flora, Vegetation, and Ethnobotany (J.M.S.,
C.T.S.,
K.A.F.)
2. Introduction This section includes the flora and vegetation; which plants are available for food, fuel, and other purposes; which species are being exploited; and how their exploitation influences the landscape. We are interested in how the inhabitants adjust to their natural environment and how these adjustments in turn affect their environment. Our research plan combines the approaches of landscape ecology and human adaptability. On the one hand, landscape ecology examines "the dynamic role of man in the landscape and . . . its ecological implications" (Naveh and Lieberman 1984: 9). On the other hand, human adaptability deals with the interactions of human populations with each other and their environments as they adjust to specific environmental situations (Moran 1979). Accordingly, we consider three components of the Kommos area: (1) environmental features including climate, geology, landforms, and soils; (2) the flora, plant communities, and plant resources; and (3) human influences as people adjust to the environment through land clearance, cultivation, plant gathering, animal grazing, and fire (Table 4.1). We use this study of contempo-
Table 4.1. Environment vs. human adjustments and influences on the landscape of southern Crete.
Environment
Resulting landscape
Climate Geology Landforms Soils
Flora Vegetation Plant resources
Human adjustments and influences Cultivation Land clearance Plant gathering Fire Grazing
94
Flora
rary ecology to help interpret the ancient environment, vegetation, and human influences during the occupation of Kommos.
3. Methods Field work was conducted from 25 March to 20 April and 12 to 31 July 1980, from 9 April to 10 May 1982, and from 9 to 18 May 1984. From our base in Pitsidia, we (J. M. Shay and C. T. Shay) made plant-collecting trips using the area's extensive system of trails (Pl. 4.10). We focused our efforts on the immediate vicinity of Kommos and the Pitsidia Valley but also included adjacent valleys and hills. Our aim was to collect plants and specimens of wood in a variety of habitats as a prelude to a vegetation survey and analysis of charcoal from the Kommos excavations. Plants were assigned preliminary identifications using Flora Europaea (Tutin et al. 1964-80) and prepared in two sets. One was sent to Dr. W. Greuter, Director of the Berlin Botanic Gardens, to be verified; the other was retained at the University of Manitoba. To explore the ethnobotany of Kommos, we first determined which plants in the vicinity are at present used by local people or could have been used in the past. During field work we observed some local plant gathering and interviewed several people regarding plant uses. Our visits, however, covered the April to July period, only part of the gathering cycle. We have thus supplemented our few observations and interviews with written sources (e.g., Turrill 1929; Niebuhr 1970; Huxley and Taylor 1977; Sfikas 1979). Plant communities were sampled from 8 to 18 April 1980 and 19 April to 11 May 1982. At each of 23 locations we made a list of species and estimated their abundance by the amount of ground they covered and the frequency of their occurrence. The percentage of bare ground and the amount of ground covered by each species of woody shrub was estimated in ten randomly located 2 x 2 m sampling plots using a slightly modified Braun-Blanquet (1932) scale of cover classes (i.e., class 1 = less than 5% ground cover, 2 = 6-25%, 3 = 26-50%, 4 = 51-75%, and 5 = 76-100%)). The presence of each herbaceous species in the plot was also noted. A 0.25 x 0.25 m quadrat was nested in the northwest corner of the larger plot, and the cover of herbs, mosses, and bare ground and litter was estimated using the same scale. Specimens were collected in cases of uncertain identity. At each location, three soil samples of about 200 g each were obtained and notes made on slope, aspect, soil depth, and evidence of grazing. All plants were classified according to their life form (Raunkiaer 1934), that is, according to the way in which they survive from one growing season to the next (e.g., annuals survive as seeds). Plant communities were also compared according to their species composition. The 23 sampling sites were analyzed with several ordination (classification) procedures to detect relationships among sites and among species. The ordination technique presented here is called reciprocal averaging (ORDIFLEX; Gauch 1977). It is a multivariate technique that for ease in interpretation summarizes the abundances of a number of species. Both frequency
Environments
95
(the percentage of quadrats in which a species occurred) and the percentage of plant cover were used. Plant cover classes were converted to percentage cover by using their midpoints, e.g., cover class 1 (less than 5%) = 2.5%, 2 (5-25%) = 15.5%, etc. Initially, all 23 sites were ordinated. Then outlying sites were omitted and the remainder rerun. Next, shrubs were ordinated without the herbs, because shrubs are the life form that characterizes most Kommos plant communities. Soil texture was measured using standard soil sieves and the pipette method of Smith and Atkinson (1975). The alkalinity (pH), conductivity, and concentration of calcium, available nitrogen, and phosphorus were analyzed at the Manitoba Soil Testing Laboratory, Winnipeg, using the procedures of McKeague (1978). To explore the links between soils and vegetation, soil texture and chemistry were related to site scores derived from the reciprocal averaging using product-moment correlations (Pearson's r; Sokal and Rohlf 1981).
4. The Environment of the Kommos Area Climate The climate is typically Mediterranean and semi-arid with mild wet winters and hot dry summers (Nahal 1981). At Gortyn, 20 km northeast of Kommos, January temperatures average 11.8°C and those in August, 27.4°C (Zohary and Orshan 1965). Annual rainfall is 484 mm, most of which falls from November to March, some in intense downpours. Heavy rains result in the loss of potential soil moisture through runoff. There is little or no rain during the dry season, from June to August.
Geology and Landforms The landscape around Kommos is diverse. Steep, craggy hills sparsely covered with spiny bushes and herbs flank the site on the south and form cliffs above the sandy beach. The rugged hills and cliffs contrast with the valley bottoms near the coast that are covered with sand. Inland, these sandy areas give way to small cultivated fields. Ancient Kommos is perched on low limestone cliffs 20 m above the Libyan Sea, on a northsouth stretch of coast at the mouth of the Pitsidia Valley (Pl. 4.10). The site extends along the edge of the sea cliffs which slope southward to the beach. It is overlooked on the southeast by Vigles, a limestone hill (elevation 80 m; Pl. 4.1). The Pitsidia Valley, flanked by low limestone hills (elevation 100-150 m) is about 2 km long and 1 km wide and is drained by an intermittent stream. The village of Pitsidia is at the head of the valley. A blanket of sand covers the Kommos site, the low coastal hills to the north, and the lower part of the Pitsidia Valley. The sand is up to 14 m deep in places and, although much of it is recent, there is evidence that some was present in Minoan times (Chap. 3, Section 2, under "The Recent Sand Cover").
96
Flora
Soils The intermittent sand accumulation as well as erosion of the surrounding hills has in places inhibited soil development. Thus, soils on the sand and on steeper slopes tend to be shallow and low in fertility (Chap. 6). Bottom-land soils are deeper and moderately fertile. These variations can be traced to several factors. The fertility of a soil depends upon interactions between the climate, geological parent material, and the associated plant and animal life. For example, plant growth may be limited by a lack of adequate soil moisture and nutrients, such as nitrogen and phosphorus. These limitations apply to the soils sampled around Kommos. The composition of weathered parent material affects the availability of plant nutrients. The predominant limestone and marl bedrock is rich in calcium and magnesium carbonates. Erosion of this bedrock ensures that the soil will be rich in these elements (Table 4.2). However, high concentrations of carbonates, by raising soil alkalinity (pH), reduce the availability to plants of other nutrients such as phosphorus and nitrogen (Barber 1984). Consequently soils in limestone areas in Greece tend to be low in these elements (Allbaugh 1953) and only potassium is usually available in adequate amounts. As pH increases, the availability of these key nutrients declines (Barber 1984). Kommos soils, with an average pH of 8, are relatively low in nitrogen and phosphorus (Table 4.2). Soil moisture and nutrients are both affected by the texture of the soil. A soil that is rich in clay and silt can hold much more water than one rich in sand (Barber 1984). By this criterion, most of the soils in the plant communities we studied have a low to moderate moistureholding capacity. Sandy areas (sampling sites 3, 4, 9, and 15) have soils with 90% sand and only 10% silt and clay (Table 4.2). Rocky sites with shrublands have as much as 50% silt and clay and can hold more water, although these rocky soils also tend to be shallow and are usually less than 0.3 m deep. The surface area of the soil particles is important for holding nutrients as well as water. Clay particles, for example, may have more than ten thousand times the surface area per unit volume than those of sand (Barber 1984). This difference in surface area in part explains why the sandy soils are lower in nitrogen, potassium, calcium, and magnesium than those in the rocky shrublands (Table 4.2). The amount of plant cover and the litter it produces also play a role in providing soil moisture and nutrients. Decomposing plants contribute to the soil organic matter, which is responsible for at least half of a soil's nutrient-holding capacity (Brady 1984). Soils rich in organic matter can also hold much more moisture than mineral soils. Organic matter in the Kommos soils is low, averaging only 2% (Table 4.2), less than half the average for soils in Mediterranean France (Rapp and Lossaint 1981). Conductivity is an overall measure of nutrient availability. This measure also shows that the soils around Kommos are nutrient poor with an average conductivity of 0.4 mScm^ 1 (micro Siemens per centimeter; Table 4.2). Soils on the sea cliffs are somewhat higher in conductivity, doubtless due to the effects of salt spray. In sum, soils of the plant communities
2.6 ± 2.2
All Sites Combined
*One sample only
49.5 ± 26.9 39.9 ± 21.4 10.2 ± 6.2
3.0 ± 2.2
Shrublands 41.4 ± 21.4 46.1 ± 17.0 11.9 ± 5.3 1, 2, 5, 6, 7, 8, 10, 11, 12, 13, 14, 16, 17, 18, 19, 20, 22, 23
0.9 ± 0.2
2.6 ± 1.6
2.1 ± 1.0
4.5 ± 2.6
2.7 ± 1.0
Nitrate Nitrogen Available (NO 3 ) Phosphorus PPm (P) P P m
0.8 ± 0.6
9.3 ± 5.8
Clay