The Human Use of Caves 9780860548591, 9781407349770

Proceedings of a conference covering many aspects of human use of caves: from Palaeolithic carvings in France to present

220 79 36MB

English Pages [229] Year 1997

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

The Human Use of Caves
 9780860548591, 9781407349770

Table of contents :
Front Cover
Title Page
Copyright
Table of Contents
Contributors
Preface
Convenient Cavities: Some Human Uses of Caves and Rockshelters
Cave Versus Open-Air Settlement in the European Upper Palaeolithic
The Human Use of Caves in the Belgian Palaeolithic
Cave as Context in Palaeolithic Art
Dancing in the Dark: Probing the Phenomenon of Pleistocene Cave Art
Settlement Patterns, Cave Sites, and Locational Decisions in Late Pleistocene Central Europe
Human Use of Caves in the Postglacial: Archaeological Evidence from the Framconia Alb, Southern Germany
The Weidental Cave: Changing Use in Changing Times
Changes in the Use of Caves in Cantabrian Spain During the Stone Age
Back Home! Neolithic Life and the Rituals od Death the the Portuguese Ribatejo
The Human Use of Caves in East-Central Italy During the Mesolithic, Neolithic, and Copper Age
Symbolism of Space in Guanche Culture Cave Sites of the Canary Islands
Cult Caves of Latvia
Artifical Caves Used as Human Dwellings and Out-Houses in the Baranja Region of Croatia
Working in the Dark? Caves as Workshops in Roman Britain
Archaeology and the Ethnohistory of Cave Dwelling in Scotland
The Human Use of Caves and Rockshelters in the Mountain Crimea (Ukraine) in Prehistoric, Classical, and Medieval Times
Human Use of the Northern Urals Caves from the Late Pleistocene to Modern Times: An Arachaeozoological Perspective
Human Use of Caves in the Caucasus
Palaeolithic Cave Sites in Gubs Canyon, Northern Caucasus
The Natufian Use of el-Wad Cave, Mount Carmel, Israel
Nahal Qanah Cave: A Unique Chalcolithic Burial Site in West Samaria
Changes in the Significance of a Site: The Klasies Cave Complex in the Middle and Later Stone Ages
Secrecy and Creativity: The Use of Rockshelters by the Nyau Secret Society, Malawi
Australian Aboriginal Use of Caves
Precolumbian Cave Utilization in the Maya Area
Rockshelters and Ritual Activities in the Atacama Desert of Northen Chile
The Human Use of Caves

Citation preview

BAR S667 1997

BONSALL & TOLAN-SMITH (Eds): THE HUMAN USE OF CAVES

B A R

The Human Use of Caves Edited by

Clive Bonsall and Christopher Tolan-Smith

BAR International Series 667 1997

Published in 2016 by BAR Publishing, Oxford BAR International Series 667 The Human Use if Caves

© The editors and contributors severally and the Publisher 1997 Volume editor: Rajka Makjanic The authors' moral rights under the 1988 UK Copyright, Designs and Patents Act are hereby expressly asserted. All rights reserved. No part of this work may be copied, reproduced, stored, sold, distributed, scanned, saved in any form of digital format or transmitted in any form digitally, without the written permission of the Publisher.

ISBN 9780860548591 paperback ISBN 9781407349770 e-format DOI https://doi.org/10.30861/9780860548591 A catalogue record for this book is available from the British Library BAR Publishing is the trading name of British Archaeological Reports (Oxford) Ltd. British Archaeological Reports was first incorporated in 197 4 to publish the BAR Series, International and British. In 1992 Hadrian Books Ltd became part of the BAR group. This volume was originally published by Archaeopress in conjunction with British Archaeological Reports (Oxford) Ltd/ Hadrian Books Ltd, the Series principal publisher, in 1997. This present volume is published by BAR Publishing, 2016.

BAR

PUBLISHING BAR titles are available from:

EMAIL

PHONE FAX

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

Contents

List of Contributors

V

vii

Preface Lawrence Straus

Convenient Cavities: Some Human Uses of Caves and Rockshelters

1

Pavel Dolukhanov

Cave vs Open-air Settlement in the European Upper Palaeolithic

9

Marcel Otte

The Human Use of Caves in the Belgian Palaeolithic

14

Ann Sieveking

Cave as Context in Palaeolithic Art

25

Paul G. Bahn

Dancing in the Dark: Probing the Phenomenon of Pleistocene Cave Art

35

Berit Valentin Eriksen

Settlement Patterns, Cave Sites and Locational Decisions in Late Pleistocene Central Europe

38

Bettina Stoll-Tucker

The Human Use of Caves in the Postglacial Period: the Archaeological Evidence from the Franconian Alb (Bavaria, Germany)

50

Erwin Cziesla

The Weidental Cave (Germany): changing use in changing times

52

Manuel Gonzalez Morales

Changes in the Use of Caves in Cantabrian Spain to the Palaeolithic/ Neolithic Transition

63

Luiz Oosterbeek

Back Home! Neolithic Life and the Rituals of Death in the Portuguese

��

m

Robin Skeates

The Human Uses of Caves in East-Central Italy during the Mesolithic, Neolithic and Copper Age

79

Michael R. Eddy

Symbolism of Space in Guanche Culture Cave Sites of the Canary Islands

87

Yuris Urtii.ns

Cult Caves of Latvia

90

Zarko Spanicek

Artificial Caves Used as Human Dwellings and Out-houses in the Croatian Region of Baranja

101

Keith Branigan

Working in the Dark? Caves as Workshops in Roman Britain

114

Roger Leitch & Christopher Tolan-Smith

Archaeology and the Ethnohistory of Cave Dwelling in Scotland

122

Grigoriy M. Burov

Human Use of Caves and Rockshelters in the Crimean Mountains (Ukraine) in Prehistoric, Ancient and Mediaeval Times

127

Alexander Borodin & Pavel Kosintsev

Human Use of Caves in the Northern Urals from the Late Pleistocene to Modern Times

136

Vassily P. Lubin

Human Use of Caves in the Caucasus

144

111

Elena Beliaeva

Palaeolithic Caves in the Gubs Canyon (Northern Caucasus)

150

Mina Weinstein-Evron

The Natufian Use of el-Wad Cave, Mount Carmel, Israel

155

Avi Gopher & Tsvika Tsuk

Nahal Qanah Cave: a Chalcolithic Burial Site in West Samaria

167

Zoe Henderson & Johan Binneman

Changes in the Significance of a Site: the Klasies Cave Complex in the Middle and Later Stone Ages

175

Secrecy and Creativity: the Use of Rockshelters by the Nyau Secret Society, Malawi

185

Josephine Flood

Australian Aboriginal Use of Caves

193

Andrea Stone

Pre-Columbian Cave Utilization in the Maya Area

201

Penny Dransart

Rockshelters and Ritual Activities in the Atacama Desert of Northern Chile

207

The Human Use of Caves

217

Yusuf M. Juwayeyi

Christopher Tolan-Smith & Clive Bonsall

iv

Contributors

PaulG.Bahn 428 Anlaby Road, Hull HU3 6QP, United Kingdom.

Berit Valentin Eriksen Department of Prehistoric Archaeology, University of Aarhus, Moesgaard, DK-8270 H!2Sjbjerg, Denmark.

Elena Beliaeva Institute of the History of Material Culture of the Russian Academy of Sciences, Dvortcovaja nab. 18, 191186 St Petersburg, Russia.

Josphine Flood Ffynnon Bedr, Llanbedr Y Cennin, Conwy, Gwynnedd LL32 8YZ, United Kingdom.

Johan Binneman Albany Museum, Somerset Street, Grahamstown 6140, South Africa.

and 31 Durack Street, Downer, Canberra, ACT 2602, Australia.

Clive Bonsall Department of Archaeology, University of Edinburgh, Old High School, Infirmary Street, Edinburgh EHl lLT, United Kingdom.

Manuel R. Gonzalez Morales Departamento de Ciencias Hist6ricas, Universidad de Cantabria, E-39071 Santander, Spain.

Alexander Borodin Institute of Plant and Animal Ecology, Russian Academy of Sciences, 202 8th March Street, 620144 Ekaterinburg, Russia.

Avi Gopher Department of Archaeology & Near Eastern Studies, University of Tel Aviv, Ramat Aviv, Tel Aviv 69978, Israel.

Keith Branigan Department of Archaeology & Prehistory, University of Sheffield, Northgate House, West Street, Sheffield Sl 4ET, United Kingdom.

Zoe Henderson

National Museum, P.O. Box 266, Bloemfontein 9300, South Africa. Yufus M. Juwayeyei Commissioner for Culture, Ministry ofY outh, Sports & Culture, Lilongwe 3, Malawi.

Grigoriy M. Burov Department of Ancient and Medieval History, State University of Simferopol', 4 Y altinskaya Street, 333036 Simferopol', Ukraine. Erwin Cziesla Martin Wurzel Archaologie GmbH, FasanenstraJ3e 25 b-d, D-14532 Stahnsdorf, Germany.

Pavel Kosintsev Institute of Plant and Animal Ecology, Russian Academy of Sciences, 202 8th March Street, 620144 Ekaterinburg, Russia.

Pavel Dolukhanov Department of Archaeology, The University, Newcastle upon Tyne NEl 7RU, United Kingdom.

Roger Leitch 58 Seafield Road, Dundee DDl 4NW, United Kingdom.

Penny Dransart Department of Archaeology, University of Wales, Lampeter, Ceredigion SA48 7ED, United Kingdom.

Vassily P. Lubin Institute of the History of Material Culture of the Russian Academy of Sciences, Dvortcovaja nab. 18, 191186 St Petersburg, Russia.

Michael R. Eddy Department of Art History & Archaeology, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom.

Luiz Oosterbeek Centro Europeu de Investiga9ao Pre-Hist6rica do Alto Ribatejo, Laborat6rio de Pre-Hist6ria do lnstituto

V

Politecnico de Tomar, Campus da Quinta do Contador edificio M, Estrada da Serra, 2300 Tomar, Portugal.

Andrea Stone Department of Art History, University of Wisconsin-Milwaukee, P.O. Box 413, Milwaukee, Wl 53201, U.S.A.

Marcel Otte Service de Prehistoire, Universite de Liege, Place du XX Aofit 7-Al, B-4000 Liege, Belgium.

Lawrence G. Straus Department of Anthropology, University of New Mexico, Albuquerque, NM 87131-1086, U.S.A.

Ann Sieveking Peartree Farm, Gray's Lane, Wissett, Halesworth, Suffolk IP19 OJR,United Kingdom.

Christopher Tolan-Smith Department of Archaeology, The University, Newcastle upon Tyne NEl 7RU, United Kingdom.

Robin Skeates School of World Art Studies & Museology, University of East Anglia, Norwich NR4 7TJ, United Kingdom.

Tsvika Tsuk Department of Archaeology & Near Eastern Studies, University of Tel Aviv, Ramat Aviv, Tel Aviv 69978, Israel.

Zarko Spanicek Regionalni Zavod za Zastitu, Spomenika Kulture, F. Kuhaca 27, 54 000 Osijek, Croatia.

Yuris Urtans State Inspection for Heritage Protection of Latvia, Klostera iela 5/7, Riga, Latvia, LV-1014.

Bettina Stoll-Tucker Landesamt flir archaologische Denkmalpflege Sachsen-Anhalt, Landesmuseum flir Vorgeschichte, Richard-Wagner-StraBe 9-10, D-06114 Halle, Germany.

Mina Weinstein-Evron The Zinman Institute of Archaeology, University of Haifa, Haifa 31905, Israel.

The participants in The Human Use of Caves Conference photographed at Henderson Hall, University of Newcastle upon Tyne, 8 July 1993.

vi

Preface

The 28 papers which form this volume are the outcome of an International Conference on the theme The Human Use of Caves held at the University of Newcastle upon Tyne from 6-9 July 1993. Throughout their evolutionary history humans have made frequent use of natural shelters for a variety of purposes. Thus caves and rockshelters constitute a major resource for archaeologists. Too often, however, these sites are treated simply as depositional environments and are investigated with a view to answering questions about the wider world and human behaviour in general. More rarely is the question asked: 'what were the humans doing in the cave in the first place?' . We trust that this book goes some way toward redressing the balance. The idea for the Conference emanated from our collective archaeological experience of cave and rockshelter sites in coastal areas of Western Scotland. However, the meeting was never intended as a vehicle for our own research or cave archaeology per se, and we invited specialists in a variety of disciplines - anthropologists and ethnohistorians, as well as archaeologists - from throughout the world to address a number of themes in terms of the evidence for cave and rockshelter use in their areas. Those themes include caves as residential sites, defensive sites, waste disposal areas, storage facilities, ossuaries, theatres of ritual, and art galleries. The Conference was participated in by over fifty specialists from five continents. With three exceptions, all of the original papers presented at the Conference are published here, and four new papers are included. Most of the papers have been revised since the Conference, although several authors wished to emphasize that their basic texts were prepared in 1993 and reflect their thinking at that time; where this is the case it is acknowledged in a note at the end of the paper. We have adopted what we feel is the most logical format for a volume of this kind. The major subdivision is geographical and within geographical areas the papers are

arranged chronologically. The contributions cover a very broad temporal range, from the early Stone Age to the present day, and inevitably the basis of the chronology is variable. For purely practical reasons, we have not adopted a calibrated timescale for the prehistoric period; when they are based on uncalibrated radiocarbon dates or dates derived by some other radiometric technique, 'ages' are quoted in years BP. The use of BC or AD indicates that the chronology is normally derived from historical sources. Typesetting was done 'in house' on an Acom RiscPC computer fitted with a StrongArm processor, using Computer Concepts' Impression Publisher DTP software. We wish to thank the British Academy, the British Council, the Prehistoric Society, and the Universities of Edinburgh and Newcastle upon Tyne for providing the financial support which made possible the Conference and the publication of this volume. We acknowledge with gratitude the financial support of the Centre for the Archaeology of Central and Eastern Europe in funding the participation of two delegates and providing a reception. We would also like to thank Basil Butcher, John Chapman, Kris Strutt and Myra Tolan-Smith for practical help in running the Conference and Stewart Smith for managing the bookshop, Kim Wilson who re-typed several of the manuscripts onto disk, Gordon Cook and Simon Gilmour who re-drew or modified some of the illustrations, Joe Rock for technical help in the latter stages of production of the book, and Rajka Makjanic of Archaeopress for her advice throughout the editing and typesetting process. Special thanks go to Pavel Dolukhanov who provided invaluable assistance with the editing of those papers which deal with cave use in parts of Russia and the Ukraine. However, we owe our major debt to our contributors who have waited so patiently for their work to be published.

Clive Bonsall Christopher Tolan-Smith

vu

May 1997

Convenient Cavities: Some Human Uses of Caves and Rockshelters Lawrence Guy Straus

Abstract Caves and rocksheltersare convenientcavities on many landscapesthat have been (and still are being) used by humans for a wide variety of purposes. Sometimescave role or functionhas remainedmore or less the same throughoutthe period of use by humans.But other caves have witnessedmajor changes in their uses. Many have been alternatelyused by hominids,carnivoresand raptorial birds, adding to the natural complexityof formationand disturbanceprocessesin these cavities.Archaeologistsmust make the most of caves and rocksheltersas resources,not only for developingchronostratigraphies,but also to interpret the regional adaptationsof prehistoric humans - despite the fact that caves may represent only a relatively small portion of the places used by humans on any given landscape.In this paper I discuss the chief aspects of the use-historiesof a number of caves and rocksheltersthat I have excavated. Changesin the use of some caves and lateral variationsin the spatial organizationof activities in others are examined. The examples include La Riera Cave (Spain), l'Abri Dufaure (France),le TrouMagrite (Belgium),and Vidigal,Algarao da Goldra and Casa da Moura (Portugal). While caves and rocksheltershave a special place in the history and practice of prehistoric archaeology,their place in the lives of prehistoricpeople must be demystifiedsomewhat.They were simply convenientcavities for whateverneeds humans may have had at some times and in some places, for ritual, graphic expression,shelter,work,storage,and disposalof trash and of their dead.

in El Quintana!, a cave art site in eastern Asturias. The Cueva Mayor at Atapuerca (long before it gained fame as a Middle Pleistocene hominid living and dying site) was, in the late years of Franco, a favourite meeting place of local teenage boys and girls, sheltered from the disapproving gaze of parents, priest and police - as Geof Clark and I found when we interrupted this function by digging its Bronze and Copper Age site. The cave at Vidigal (northwest Alentejo, Portugal) presented evidence of very recent (poor, landless) human occupation and burial, some too fresh for comfort when I tested it in 1988. La Riera Cave was used as a machine gun nest during the Battle of El Mazuco in 1937 - the last major fight in the northern front before the fall of Oviedo. As Clark discovered to his horror, nearby Coberizas was the scene of an execution during the Spanish Civil War. Igrejinha dos Soidos and Toxugueira ('badger hole') are caves, where I discovered that modern Algarve peasants continue to hunt badgers by burning them out of their tunnels. The Abric Romani in Catalonia was used as a paupers' cemetery and nearby Consagraci6 is presently used as an opportunistic dump by inhabitants of Capellades. Too common to enumerate are the caves of Europe that have been (or still are) used to stable goats or other livestock. They include, for example, the minor cave art site of Trauno at the edge of the Picos de Europe. I have seen the classic Middle and Upper Palaeolithic cave of Spy in Belgium being used in a Boy Scout 'treasure hunt', with skull-and-cross-bones symbols pasted to its walls. And Le Trou Magrite, despite barbed wires and 'no trespassing' signs, is in frequent use as a picnic site (with a camp fire repeatedly rekindled, always in the same spot in the centre

Having twice earlier written in general terms about the role of caves and rockshelters as archaeological resources (Straus 1979, 1990), I would like to use this opportunity to present an overview of some human uses of these cavities, based on excavations I have directed or co-directed over the last 20 years in Spain, France, Portugal and Belgium. My point of departure is the belief that caves and rockshelters are features of some landscapes that have repeatedly been used by humans on an opportunistic basis for a wide variety of purposes, and that they represent only a relatively small fraction (albeit situationally variable) of the 'places' so used, while sometimes constituting a disproportionately large part of the surviving archaeological record for many prehistoric periods. Caves and rockshelters were (and are) convenient cavities on the landscape for a multitude of human uses, but we should neither overstate nor glorify their role simply because they (either as dumps or as art sanctuaries) have often tended to survive to be rediscovered archaeologically more frequently than many open-air sites.

Caves as sites in the modern world of Western Europe Without pretending to impose any sort of ethnographic analogies on the prehistoric past, it does nonetheless seem interesting to me to reflect on the variety and nature of uses to which caves and rockshelters are put in the modern world. At least this may help demystify these human sites on the landscape. One of my first 'caving' experiences in Cantabrian Spain involved stumbling over the fairly fresh carcass of a burro

1

Straus of the vast, sheltered cave mouth), as well as the locus for rock climbing and other modem activities. In the very recent past someone painted the cave of Zubialde (Alava, Spain) with crude images imitating the styles of several great Palaeolithic cave art sites, from Lascaux to Ekain. In short, caves and rockshelters continue to be important, living (or dying) places on the landscapes of 20th century people. We must think of them in often less than grandiose terms in Stone Age prehistory too. The task of the archaeologist is to try to establish why certain caves/rockshelters were chosen for human use and under what circumstances (permanent [i.e. geological, geographical] and situational [i.e. climatic, seasonal, activity-related]). It is not immediately obvious in all cases why a cave/rockshelter was used (or not used). The fact that in some regions of the world (e.g. Vasco-Cantabria, Perigord, the South Cape of Africa) the vast majority of Old Stone Age sites are in caves/rockshelters, does not necessarily mean that humans mainly established their sites in such cavities. It might mean (i) that certain types of sites were established there for situational/activity reasons under particular circumstances, and (ii) that erosion (or deep burial) has differentially made the record of open-air sites inaccessible to 20th century archaeologists, while favouring the preservation (and discovery) of occupation residues in caves. We must never forget this (though it is easy to do) when interpreting maps of surviving site distributions and when evaluating the apparent range of site functions.

rockfall - a common occurrence that can pulverize exposed materials, distort underlying strata, and totally reconfigure (or terminate) the habitable area of a cavity. All these facts must be considered when excavating caves/rockshelters and when interpreting their archaeological record. Yet these kinds of sites are often most or all of what we have, and are too important to despair in the fact of palimpsests, disturbances and destruction. There is a substantial record of human activity and behaviour in caves/ rockshelters. The task of the geologist, palaeontologist and archaeologist is to unravel that record from unusually complex sedimentary deposits.

Residential use of caves and rockshelters The following is a discussion of human uses of two caves and one rockshelter that I have excavated. The focus is on how those uses changed through time and varied across space within the sites.

La Cueva de la Riera La Riera Cave (Vega del Sella 1930; Straus & Clark 1986) is part of an extensive karstic system in eastern Asturias (Spain). It is through this karst that water drains from the Sierra de Cuera and coastal plain under the La Llera ridge and into the Cantabrian Sea. La Riera is only one of a large number of cave mouths along the sheltered, south face of La Llera that were repeatedly chosen for human use throughout the course of the Upper Palaeolithic and Mesolithic (ca 30,000-7000 BP). Under Last Glacial conditions, rather than being located within a half-hour walk of the shore as it is under present interglacial conditions of high sea level, La Riera was in the centre of the coastal plain, a two-hour walk from the shore and from the Sierra de Cuera crest. Then, as now, fresh water was immediately available in the Rio Calabres at the point at which it enters the karst. The food (and lithic) resources of the littoral, estuaries, streams, coastal plain and steep Sierra de Cuera slopes were all close at hand from the 'central places' at La Riera, Cueto de la Mina, Bricia, Tres Calabres, Balmori and other La Llera sites. A nearby gorge of the Rio Bed6n through the Sierra de Cuera provides easy access to the east-west intermontane valley of eastern Asturias and to the Picos de Europa. In short, La Llera was a good spot for hunter-gatherers to be located, irrespective of its caves. They provided an additional, non-negligible benefit, especially in glacial times - shelter. Among the first archaeologically visible human uses of La Riera were a series of limited special-purpose occupations. Apparently small groups of people bringing non-local lithics in the forms of tools, weapons (Solutrean points) and blanks, came to the cave on several occasions to hunt mainly ibex, presumably on the Sierra de Cuera slopes. There is no evidence of constructed features during these limited-function, ephemeral occupations of the site, at least in the portion of the site excavated by us. At this time La Riera would have had a capacious vestibule (before substantial sedimentary infilling), but the cave entrance itself may have been considerably more restricted than was later

Issues of formation, disturbance and destruction Despite their excellent qualities as shelters and as traps for often deeply stratified sediments, caves and rockshelters pose significant interpretive problems, some of which have been recognized since the earliest era of prehistoric archaeology. They are often subject to a wide variety and great intensity of disturbance agencies: carnivores, running water, rockfall and intrusive excavation. Because, in all but the largest ('Haua Fteah'-size) cavities, cave mouth/rockshelter space is a constraining factor for most human uses, earlier occupation residues are often levelled, dug into, or, at the very least, conflated with the most recent residues (hence the often more serious problem of 'palimpsests' in caves/rockshelters, due to the frequent and tightly confined re-use of such places). While hearth, roasting and storage pits, drainage ditches, burials and other human modifications may disturb earlier levels at unbounded, nonarchitectural open-air sites, they are common in the frequently re-used spaces of caves/rockshelters (as well as in the walled confines of architectural open-air sites). A great diversity of large and small carnivores tunnel, den or hibernate in caves, often causing much disturbance and mimic human activities in the acquisition and modification of other animal remains. Because true caves are parts of karstic systems, running water (during times of high water table) can periodically flush out or at least disturb sedimentary deposits. Also often opening out onto river valleys, caves and rockshelters are sometimes flooded by high, running water, and hence flushed or disturbed. Finally, a major destructive agency in caves/rockshelters is

2

Convenient Cavities: Some Human Uses of Caves and Rockshelters the case, after centuries of Pleniglacial and Tardiglacial rockfall. Then use of La Riera changed, as shown by the construction of hearths and roasting pits (with hints of partly paved - as well as densely littered - living floors), by the high numbers and diversity of cores, debitage and finished tools and weapons (plus bone/antler artifacts including points, awls and needles), and by the great variety and wealth of terrestrial and aquatic faunal remains. The later Solutrean and Magdalenian levels seem to represent large-scale, multi-purpose, longer-term, residential occupations of the cave, possibly by larger groups of humans (perhaps more evenly divided between the sexes and among age segments) vis avis the basal levels. The presence of human remains in three of the middle levels of the site is also testimony to the more substantial human utilization of the cave, as is the fact that the burials had apparently been disturbed and those remains scattered as a result of frequently repeated occupation of the site in the period centred on 18,000 BP. Rock-lined hearths (for light, cooking and heating) were also observed in nearby Cueto de la Mina by Vega del Sella (1916) for this same extremely cold Solutrean time period. It was probably in this time that paintings were made on a wall at La Riera. A period of flooding of the cave in late Magdalenian times caused both erosion (presumably under a highvelocity regime of karstic rejuvenation) and silty clay deposition (presumably under conditions of ponding). Human occupations seem to have become more intermittent. By this time (Azilian) the mouth of La Riera had become so filled up with highly anthropogenic (and natural) sediments, that it was becoming difficult to stand erect in many places. Finally, La Riera (like many other Upper Palaeolithic sites along this coast) had become physically uninhabitable. Yet the sunny, sheltered south face of La Llera, in the rincon formed by the mouths of La Riera and Cueto de la Mina and bounded by the Rio Calabres, remained an attractive place for humans to camp, especially when they were heavily engaged in fishing and mollusc gathering along the now nearby early Postglacial shore. Humans responsible for the deposition of 'Asturian' shell middens did not usually live in the same caves (or in any caves), but the caves were often at hand and served as convenient holes into which to dump the noxious, bulk debris created by their shellfishing and other activities. Out of sight, out of mind and out of nose(!) La Riera became so full that, when discovered (by intuition) by the Conde de la Vega del Sella in 1916, its mouth was completely blocked. (This sounds like the predecessor to my city's problem of overstuffed landfills.) La Riera's roles had changed significantly throughout its 13,000-year run as a humanly used site. Part of this was a result of its changing position on a changing landscape, but much of it was caused by changing human activities and depositions.

time. In contrast, although there were 5000 years of archaeological levels at the Abri Dufaure, the Dufaure Prehistoric Project (Straus, Akoshima et al. 1988; Altuna et al. 1991; Straus 1992, 1995) concentrated on 'horizontal' problems of site formation and activity structure within the major terminal Magdalenian level (ca 11,000 BP), albeit in the context of the Tardiglacial adaptive system in the French Pyrenean region. The Dufaure site consists of a small rockshelter and a talus slope at the base of the prominent, limestone Pastou cliff, near the confluences of the Oloron, Pau and Adour rivers at the southern end of the Aquitaine Basin, 60km from the crestline of the Basque Pyrenees in south-west France. It and the other Pastou rockshelter sites (Duruthy, le Grand Pastou and le Petit Pastou) were located at a particularly strategic spot along a natural avenue of communication between the Pyrenees, Gascon lowlands and Atlantic coast. A permanent ford crosses the Gave d'Oloron right in front of the sites and the Pastou cliff would serve io deflect reindeer migrations and to help humans slaughter large numbers in their passage across the river. The cliff also would provide shelter from north and north-west winds, as well as excellent solar heating along its white, south-facing wall. Good-quality nodular and tabular flints are locally available in separate sources on opposite sides of the Gave d'Oloron. All these features make it easy to understand why Magdalenian groups repeatedly chose to occupy the base of the Pastou Cliff in winter - often at Duruthy, sometimes at Dufaure or the other rockshelters, perhaps depending on such circumstantial factors as vermin and garbage build-up (or even death-site taboos?). Stratum 4 (terminal Magdalenian) at Dufaure consists of pavements of cobblestones constructed by humans during a long series of cold-season occupations of the site during the fairly temperate, very humid Aller!lld climatic phase. It is culturally, chronologically, structurally and seasonally identical to Couche 3 at nearby Abri Duruthy, where huge investments were repeatedly made in paving extensive areas of the rockshelter and talus surface (Arambourou 1978). Pavements had been uncovered in 1900 within the Dufaure rockshelter by Breuil and Dubalen (1901). There they were associated with abundant amounts of charcoal and ash, indications of hearths, probably used for cooking and for heating the sheltered sleeping area of the site. Based on large amounts of limestone blocks in overlying (Azilian) Stratum 3 on the narrow terrace right in front of the present rockshelter, it seems likely that the covered area had been as much as twice as deep in Magdalenian times as it is today. But two huge blocks inside the extant abri, as observed by Breuil and Dubalen, were already in place by Magdalenian times and served to partition the rockshelter and to separate its then inner and outer areas. The space available to Aller!lld-age inhabitants would, thus, have included a rockshelter of moderate size (40-50m 2) and an open talus slope with a typical sigmoidal profile. We excavated the remaining flat area of about 3m2 at the top of the slope and about 20m2 of intact Stratum 4 pavements on the upper slope. Extensive refitting and statistical analyses (computer mapping of burned objects) were carried out by M. Petraglia (1987, 1995). This work,

L' Abri Dufaure The La Riera Paleoecological Project was essentially a study in changing site use and regional adaptations through

3

Straus by revealing 'latent structures', added to our field observation of many fragile faunal remains (some in nearanatomical connection) and antler artifacts, and of possible single-carcass bone units associated with dense, spatially delimited clusters of lithic artifacts (possible residues from individual butchering episodes). All these analyses and finds combine to demonstrate that the terrace and upper slope areas of the site had remained essentially intact, despite the steepness of the slope and despite devastating erosion to the east, south and west, as documented in other excavation trenches and pits. K. Akoshima (1993, 1995) has done extensive site structural analyses of the 22.5m 2 of remaining intact Stratum 4 deposits at Dufaure. He first defined the site framework in terms of the cliff wall, the rockshelter overhang and terrace, the areas of boulder outcrop already present in terminal Magdalenian times, and microtopographic variations in the utilized area of the talus slope (i.e. changes in slope angle from metre to metre - partly controlled by the underlying bedrock contour of earlier clifffaces ). These elements of the site topography were very important in determining how the site space was very specifically used at different times. By means of statistical and cartographic analyses, Akoshima has been able to show that the excavated area of Stratum 4 contained evidence of two distinct patterns of structured activities. For the first half of the long period of the terminal Magdalenian occupations, large fires were repeatedly built in the centre of the upper slope (L & M rows), possibly having to do with mass roasting of reindeer meat. In a micro-topographic depression to the south (downslope) of the burning area, there is a possible 'dump' area consisting of dense, unstructured masses of lithic debris and faunal remains through five pavements in the N-P rows. Retouched tools, especially burins, are much more frequent on the flat terrace at the front of the rockshelter than downslope. The terrace also has debris (partly refitted by Petraglia) from a clearly delimited episode of chalcedony knapping. Petraglia found other instances of in situ core reduction _in the I row at the outer edge of the terrace - the only part left over from the 1900 excavation. The terrace seems to have been a specialized area for toolmaking and using. A small possible hearth there may be indicative of the kinds of hearths once found within the present abri and used for light, cooking and space-heating. Akoshima has found that this tripartite subdivision of the intact excavated area broke down in the later Stratum 4 occupations. On the last four pavements to be constructed, the bonfire area disappeared and retouched tools are more evenly distributed. There are small fire areas at the foot of the upper break-in-slope (K row) and near the lower part of the upper slope (0 row). However, in general (and with the caveat that the excavated area represents only a small portion of the original site area) it seems as if activities became less spatially compartmentalized during the most recent Magdalenian occupations. This apparent change in site structure (coincidentally or causally?) was correlated with a significant decline in the relative importance of reindeer and its steady replacement by red deer as game at Dufaure. Whether the change in spatial organization was

related to a decline in truly mass processing of carcasses with the declining importance of reindeer slaughter, remains to be demonstrated. At any rate, Dufaure is an excellent example of how a complex palimpsest deposit in a rockshelter can be studied, (i) to demonstrate relative intactness, and (ii) to demonstrate the existence of structured activity areas resultant from the redundant use of the topographically circumscribed space in such a land feature.

Le Trou Magrite Recently, in association with M. Otte (University of Liege, Belgium), I excavated the open-air Gravettian site of Huccorgne and remnant Aurignacian and Mousterian deposits in le Trou Magrite (Straus, Otte et al. 1992; Otte et al. 1992). The latter is a historic karstic cave site in the lower Lesse valley near its confluence with the Meuse in southern Namur Province. The archaeological contents of the inner cave were entirely excavated in the 19th (notably by E. Dupont [1873]) and early 20th centuries, while most of the sloping cultural deposits in the high-roofed outer cave were also removed - both for construction of a scenic promenade in the 1830s and by subsequent archaeological trenching in the axis of the cave mouth. A moderate-size area (ca 15m2) of truncated intact archaeological deposits was, however, discovered as a result of our 1991-92 excavations in le Trou Magrite. These deposits lay at the western edge of the outer cave. Although presently situated just outside the dripline, the human occupations (particularly in Mousterian times) would have been under the cave overhang, since huge blocks testify to massive collapses and retreat thereof during the Last Glacial. In the area of remnant intact Palaeolithic deposits at le Trou Magrite (Trench C), the top of the Aurignacian (Stratum 2) outcrops essentially at the present ground surface, covered only by a thin layer of humus and mixed modem and prehistoric material (Stratum 1). In short, deposits radiocarbon-dated to ca 30,000 BP have apparently been lying exposed but untouched on the surface since the 1830s. Not a single intrusive object was found in the area of intact, stratified Palaeolithic deposits. Stratum 2 is composed of fine angular eboulis (cryoclastic gravels) with a silt matrix. Despite this fact and despite its 160-year exposure, faunal remains were relatively well preserved and undisturbed. They included several cervid dental series (including cases of deciduous teeth) in anatomical order despite the absence of alveolar bone. Even preliminary observations indicate the presence of lithic refits in Stratum 2. Stratum 3, AMS dated to 38,000-40,000 BP is similar in archaeological content (e.g. keeled and nosed end-scrapers, retouched blades, leaf points) and in geological origin (i.e. cryoclastic) to Stratum 3, but also contains larger roof-fall blocks. Preliminary computer back-plotting by A. Martinez indicates that there are distinct lenses of faunal remains and artifacts within Strata 2 and 3. Although there were clear localized hints of the existence of such 'living floors' during excavation of particular metre squares or subsquares, their horizontal extent only really becomes evident with retromapping. This is because of the presence of large blocks (in

4

Convenient Cavities: Some Human Uses of Caves and Rockshelters

Stratum 3) that frequently interrupt the lenses and because excavation techniques had to emphasize vertical stratigraphic control at the expense of broad horizontal exposures. Artifact analyses show that the full gamut of lithic manufacturing stages was conducted in situ, although there are clear differences in the treatment of local indurated limestone versus imported flints. Most formal tools are made on the latter, suggesting the existence of both expedient and curated aspects of the Trou Magrite Aurignacian lithic technology. The heavy use of local limestone implies a substantial residential use of the cave at this time, with flints from ca 70km distant being procured either logistically by special-purpose expeditions to the sources or as an 'embedded' aspect of hunting trips, or as a result of longterm, wide-ranging residential moves, or through exchange with other groups. The large, diverse faunal assemblages of Strata 2-3 (including reindeer, horse, rhinoceros, bison and ibex - plus a variety of carnivores, according to identifications by A. Gautier) reinforce the impression that the cave was used for fairly large, multi-purpose residential occupations during the Aurignacian. These were cold season occupations according to dental cementum analyses by A. Stutz. The first of those occupations found a cave mouth surface containing many huge roof-fall blocks that had fallen from the overhang during a period of only occasional ephemeral human use of le Trou Magrite: Stratum 4, probably dating to the Wiirm Interpleniglacial (P. Haesaerts, pers. comm.). These boulders outcropping from Stratum 4 would have provided significant elements of the site framework (and good windbreaks) for at least the initial Stratum 3 inhabitants. At the base of Stratum 4 there is a patchy layer of clayey, calcium carbonate-rich silt extremely rich in microfaunal remains, studied and dated to oxygen isotope stage Sb by J-M. Cordy. This is almost certainly the aggregate result of the deposition of owl regurgitation pellets (Andrews 1990). Before the big collapses during Stratum 4 times, the cave mouth overhang extended southward by several metres and must have served as an owl roost. Stratum 5 per se, consisting of colluvial or alluvial silt and some blocks, contains rare lithic artifacts of Mousterian aspect and scattered faunal remains, including cave bear, wolf and lynx. At this time (early Last Glacial) the cave seems to have been used as a carnivore lair, with only brief, ephemeral visits by hominids (Neandertals). Faunal analyses determined that the ungulate remains in Stratum 5 were often the result of carnivore activity or natural deaths (cf. Brain 1981; Straus 1982; Gamble 1983). The first surviving sedimentary deposit - in contact with bedrock - is fluviatile. Coarse sands, gravels, pebbles and cobbles (often very large) constitute Stratum 6, which is completely devoid of faunal or cultural remains. The surviving 3-metre stratigraphy at the entrance of le Trou Magrite thus contains evidence of highly varied natural, animal and anthropogenic depositional processes. It also attests to very different types of animal and human uses of the site, the latter including both insubstantial short-term visits and substantial long-term, multipurpose residences.

Caves as stables, dumps, burial grounds and living sites Some caves, because of their location and by their artifactual and faunal contents (and sometimes lack of constructed features), clearly had very specialized functions that were maintained throughout long periods of time. Good examples of such sites are the ibex-hunting locations of Les Eglises in the Pyrenees, Bolinkoba and Rascafio in the Cantabrian Cordillera, and Collubil in the Picas de Europa (Straus 1987, with references). These were classic 'logistical locations' (sensu Binford 1982). In contrast, other caves that I have excavated were used for wide ranges of human purposes. The sinkhole-like vestibule of la Cueva Mayor at Atapuerca (Burgos, Spain) yielded a stone slab-lined hearth in an early Bronze Age level, evidence that the cave was actually inhabited by humans at that time. Above this, however, the sediments are organic-rich clays with patches of ash. They are rich in remains of domesticated sheep/goat, cattle and, to a lesser extent, pig. Most of these animals were culled as juveniles or neonates. The cave, with its very steep entrance, probably served as a stable/slaughterhouse in late Bronze Age/early Iron Age times (Oark et al. 1979). This function may have been associated with a large open-air site on the slope just outside the cave. The Algariio da Goldra (Algarve, Portugal), an aven with relatively difficult access, yielded disarticulated remains of seven humans (3 young adults, an adolescent, and 3 children) from the basal cultural horizon (middle Neolithic, ca 5000 BP) in a 2xlm sondage. Associated materials included faunal remains, scattered ash and charcoal, ceramic sherds, and a few stone and bone artifacts, all of which might suggest that the human carcasses had been simply heaved into the aven along with the rest of the 'trash'. However, the discovery of a small, whole, incised bowl in this level suggests that the midden might actually represent in situ human habitation - as well as inhumation. After a period of abandonment and then a minor Chaieolithic occupation, Goldra became a major owl roost, causing the formation of a regurgitation lens composed almost entirely of rodent and insectivore remains (Straus, Altona et al. 1992; Crispim et al., 1993). The classic site of Casa da Moura (Estremadura, Portugal) was first dug by J.F. Nery Delgado in 1866, resulting in a most extraordinary study of the formation processes of a cave infilling (Delgado 1867). My work in this aven in 1987 confirmed his fundamental conclusions. The basal sands, under a massive flowstone, are absolutely sterile of faunal (even microfaunal) and archaeological remains; they were laid down by dissolution of the limestone before the cave was open to the surface. The first occupation of the cave was by carnivores: a wolf jaw embedded in the top of the flowstone yielded an AMS date of ca 25,000 BP. Then there was a human occupation during the Solutrean/Last Glacial Maximum (unfortunately completely dug out by the earlier excavations). Again Casa da Moura was abandoned and finally re-used as an early Neolithic (ca 6000 BP) and Chalcolithic ossuary (Straus, Altona et al. 1988; Zilhiio, in

5

Straus changing places on the human landscape, even if they may remain fixed spots for millennia. And as places used by humans, they continue to live.

press). Similar use of caves as ossuaries is common in the Portuguese (and Spanish) Chalcolithic; we found human remains in association with the sherds from much of a large bowl in Buraca dos Mouros (Bocas, Rio Maior, Estremadura, Portugal) (Straus, Altuna et al. 1988). Finally, the cave of Vidigal is another example of widely varying human uses of such cavities. In this case, the 'cave' is a deep rockshelter in an indurated sand dune near the Atlantic coast of northern Alentejo (Portugal). The cave is literally under the large open-air Mesolithic shell midden site of the same name (Straus et al. 1990). Two test pits (1 x Im and 2 x Im) dug to 'bedrock' in Vidigal Cave showed the stratigraphy to have been considerably disturbed by rabbits and humans. Fish, ovicaprine and human remains (burnt and unburnt) were relatively common (Le Gall et al. 1994). The infilling included layers consisting almost solely of mollusc shells. There were also patches of ash, charcoal and (in upper spits) unburnt leaves and sticks. The few patently modern artifacts tended to be in upper layers, despite the presence of large intrusive pits. But there were a few simple cobble grinding stones throughout the deposit and on the cave surface. Near contact with basal sterile sand (i.e. disintegrated 'bedrock') in one of our sondages, we found human infant remains associated with sherds from three different vessels, a chalcedony blade, 3 flint blades, a quartzite flake and a flint flake. These probably date to the middle Neolithic. The striking fact about Vidigal is that it seems to have been used as a temporary shelter, shellfish, fish and vegetal food processing and consumption locus, and human burial site from Neolithic times up to the present. Landless people are known to have occupied this area at least into the 1970s, according to a local historian (A. Quaresma, pers. comm.). These people ('malteses') practised a subsistence foraging economy, including shellfish and nut collection. Our survey of a transect between the Cereal Mountains and the Atlantic coast, centred on Vidigal, yielded slab grindstones and cobble handstones - and a recently occupied brush shack not far from Vidigal. Did the occupants of this lean-to in the middle of the wild, exposed coastal maquis sometimes take shelter, eat and die in Vidigal Cave? Maybe. It was convenient - and there was no room at the inn.

Epilogue Since this paper was written in 1993, I have excavated three more, very different cave/rockshelter sites: Abri du Pape and Grotte du Bois Laiterie (Namur Province, Belgium) in collaboration with M. Otte, and Cueva del Mir6n (Cantabria, Spain) in collaboration with M. Gonzalez Morales. These physically very different sites also manifest very diverse uses and run a broad gamut in terms of archaeological problems and potential. Pape is a small rockshelter at the foot of a 100-metre vertical cliff of the Meuse River canyon. Its 7 .5-metre geoarchaeological stratigraphy built up quickly since initial Holocene times by rapid scree deposition and spans virtually every cultural period between the early Mesolithic and the Middle Ages, with uses ranging from ephemeral campsites to ossuaries (the post-Mesolithic components were excavated under the direction of J-M. Leotard (1989), and include middle and late Neolithic, Iron Age and Roman horizons - despite the unimpressive size of the usable area at Pape). The 9000 and 8000 BP Mesolithic components seem to represent little more than informal, short-term, limited-activity bivouacs in a hollow bounded by the rear of the shelter, two lateral scree cones and the talus slope descending to the banks of the Meuse (Straus et al., in press). Bois Laiterie (Otte & Straus 1997) is a tiny, uncomfortable cave (no more than ca 3Sm2 of minimally habitable space that is drafty and faces due north). It is far (30-70km) from sources of good flint, but it dominates the gorge of a tributary that is a key avenue of communication between the Sambre-Meuse interfluve plateau and the Meuse itself. Besides being an excellent location for ambush hunting and for fishing, BL is 'en route' between open-air Magdalenian sites on flint sources in Middle Belgium and major Magdalenian cave sites (such as Chaleux) on the edge of the Ardennes uplands. We were able to demonstrate that BL was a highly specialized, warm-season campsite used only a few times right around 12,650 BP during the B{11lling climatic optimum, with minimal human investment in infrastructure. When not occupied by humans, the cave was used as a den by arctic and common foxes. Yet this insignificant stopover spot was part not only of a complex Magdalenian settlement-subsistence system established during B!1illing times in W allonia, but also of a broad network of social contacts and exchanges that interlinked the Magdalenian community of north-west Europe, as manifested by our finding of eight Miocene fossils (most of which are perforated) whose source is probably the Paris Basin, the location of such famous penecontemporaneous sites as Pincevent and Verberie. Similar fossils have also been found at Chaleux and several other Belgian sites. When next used by humans, BL became an early Mesolithic burial cave, one of several in the Upper Belgian Meuse-Sambre Basin to date to 9000-9500 BP (by direct AMS dating of human bones) and to have no (or virtually

Conclusions Examples from my archaeological experience could be multiplied, but hopefully I have illustrated at least a small part of the range of human uses of caves through time. The human uses of individual cavities changed, as their physical configuration changed, as their place in the landscape changed, or simply as human needs, numbers or adaptive strategies changed. Other cavities seem to have maintained the same basic role or function throughout long periods of time. The walls, roofs, rockfall blocks, talus slopes and other physical features of caves/rockshelters served in each instance as elements of a framework to structure and organize human occupations. However, these features are not static; nor are such phenomena as the water table and the possibility of karstic rejuvenation or flooding from external streams. Caves and rockshelters are ultimately

6

Convenient Cavities: Some Human Uses of Caves and Rockshelters no) associated cultural materials (Toussaint et al. 1996). Both Pape and BL (plus many other Belgian cave sites) had been only recently discovered through the efforts of just one highly insightful, persistent prospector, Ph. Lacroix (a.k.a.,'Bibiche'), despite the fact that many prehistorians had written off Wallonia as a region long-explored and archaeologically 'exhausted'. Much remains to be found! The same is true in Cantabria. El Miron was archaeologically discovered at the same time as the adjacent cave art sites of Covalanas and La Haza in 1903 but, despite its huge size (130m deep and 20m wide by 12m high at the mouth) and strategic, west-facing location dominating the confluence of three gorges and the main intermontane valley of eastern Cantabria, it has been archaeologically written off as 'disturbed' ever since. El Miron, which I first visited in 1973, is so large that any number of human activities could be and have been accommodated in it, including recent livestock corrals, stone cabins, gypsy camps and possible use as an artillery emplacement. Our test excavations in 1996 (Straus & Gonzalez Morales 1996) have already revealed not only a series of functionally diverse occupations dating to the Medieval, Bronze Age, Chalcolithic, Neolithic, Azilian, Upper and Middle Magdalenian and Solutrean periods, but also the presence of stratigraphically datable rupestral engravings that had escaped detection by a century's worth of archaeological visits to the cave. And we have only tested to a depth of at most 2.5m in what could well tum out to be an infilling of several times that thickness . . . This will give us a great opportunity to extensively sample the changing uses of a cave in an interior, montane setting and one so vast as to represent sheltered but practically unconstrained space over a very considerable span of time and environments during much of the Quaternary.

BINFORD, L.R. 1982. The archaeology of place. Journal of Anthropological Archaeology 1(1):1-31. BRAIN, R. 1981. The Hunters or the Hunted. University of Chicago Press, Chicago. BREUIL, H. & DUBALEN, P. 1901. Fouilles d'un abri a Sordes en 1900. Revue de !'Ecole d'Anthropologie de Paris 11:251-268. CLARK, G., STRAUS, L., BURTON, S. & CLARK, V.J. 1979. The North Burgos Archaeological Survey: an inventory of cultural remains. In The North Burgos Archaeological Survey, edited by G.A. Clark, pp. 18-157. Anthropological Research Papers no. 19. Arizona State University, Tempe, Arizona. CRISPIM, J., POVOAS, L. & STRAUS, L. 1993. Further studies of Algarao da Goldra and Igrejinha dos Soidos: archaeological cave sites in the Algarve. Algar 4:31-44. DELGADO, J.F. NERY 1867. Da Existencia do Homem en Tempos mui Remotos Provada pelo Estudo das Cavernas: I. Noticia acerca das Grutas da Cesareda. Comissao Geologica de Portugal, Lisbon. DUPONT, E. 1873. L' Homme pendant !es Ages de la Pierre dans !es Environs de Dinant-sur-Meuse. Muquardt, Brussels. GAMBLE, C. 1983. Caves and faunas from Last Glacial Europe. In Animals and Archaeology: 1. Hunters and their Prey, edited by J.Clutton-Brock & C.Grigson, pp. 163-172. BAR International Series S-163. British Archaeological Reports, Oxford. LE GALL, 0., ALTUNA, J. & STRAUS, L. 1994. Les faunes mesolithiques et neolithiques de Vidigal (Alentejo, Portugal). Archaeozoologia 7(1):59-72. LEOTARD, J. 1989. Occupations prehistoriques a l'Abri du Pape. Notae Praehistoricae 9:27-28. OTTE, M. & STRAUS, L. 1995. Le Trou Magrite. Etudes et Recherches archeologiques de l'Universite de Liege 69. Liege. OTTE, M. & STRAUS, L. 1997. La Grotte du Bois Laiterie. Etudes et Recherches archeologiques de l'Universite de Liege 80. Liege. OTTE, M., STRAUS, L., LEOTARD, J., GAUTIER, A. & HAESAERTS, P. 1992. Fouilles dans le Paleolithique moyen et superieur de Belgique meridionale. Rapport 1991. Notae Praehistoricae 11:3-28. PETRAGLIA, M. 1987. Site Formation Processes of the Abri Dufaure. Unpublished Ph.D. dissertation, Department of Anthropology, University of New Mexico, Albuquerque. PETRAGLIA, M. 1995. Processus de formation du gisement. In Les Derniers Chasseurs de Renne du Monde Pyreneen, edited by L.G. Straus, pp. 53-74. Memoires de la Societe Prehistorique Fran9aise 22. Paris. STRAUS, L.G. 1979. Caves: a palaeoanthropological resource. World Archaeology 10:331-339. STRAUS, L.G. 1982. Carnivores and cave sites in Cantabrian Spain. Journal of Anthropological Research 38:75-96. STRAUS, L.G. 1987. Upper Palaeolithic ibex hunting in SW Europe. Journal of Archaeological Science 14:163-178. STRAUS, L.G. 1990. Underground archaeology. In Archaeological Method and Theory, edited by M.B. Schiffer, vol. 2, pp. 255-304. University of Arizona Press, Tucson. STRAUS, L.G. 1992. L' Abri Dufaure et la falaise du Pastou dans le systeme adaptatif regional des Pyrenees au Magdalenian. In Le Peuplement Magdalenien, edited by H. Laville, J. Rigaud & B. Vandermeersch, pp. 335-343. Comite des Travaux Historiques et Scientifiques, Paris. STRAUS, L.G. (Editor) In press. Les Derniers Chasseurs du Renne le Long des Pyrenees. L'Abri Dufaure: Un Gisement Tardiglaciaire en Gascogne. Memoires de la Societe Prehistorique Fran9aise, Paris.

Acknowledgements: My research in Spain, France, Portugal and Belgium has been generously supported by many grants each from the National Science Foundation, the L.S.B. Leakey Foundation, the National Geographic Society and the University of New Mexico. I wish to thank all my cave archaeologist colleagues in all four countries for sharing their underground realms with me. Any and all errors of interpretation in this paper are, however, mine alone.

References AKOSHIMA, K. 1993. Site Structure Analyses of the Abri Dufaure and Mill Iron Sites. Unpublished Ph.D. dissertation, Department of Anthropology, University of New Mexico, Albuquerque (in preparation). AKOSHIMA, K. 1995. Analyse des structures d'habitat de la Couche 4. In Les Demiers Chasseurs de Renne du Monde Pyreneen, edited by L.G. Straus, pp. 165-180. Memoires de la Societe Prehistorique Fran9aise 22. Paris. ALTUNA, J., EASTHAM, A., MARIEZKURRENA, K., SPIESS, A. & STRAUS, L. 1991. Magdalenian and Azilian hunting at the Abri Dufaure, SW France. Archaeozoologia 4(2):87-108. ANDREWS, P. 1990. Owls, Caves and Fossils. University of Chicago Press, Chicago. ARAMBOUROU, R. 1978. Le Gisement Prehistorique de Duruthy. Memoires no. 13. Societe Prehistorique Fran9aise, Paris.

7

Straus STRAUS, L. & CLARK, G. 1986. La Riera Cave. Anthropological Research Papers no. 36. Arizona State University, Tempe, Arizona. STRAUS, L., AKOSHIMA, K., PETRAGLIA, M. & SERONIEVIVIEN, M. 1988. Terminal Pleistocene adaptations in Pyrenean France: the nature and role of the Abri Dufaure site. World Archaeology 19:328-348. STRAUS, L., AL TUNA, J., JACKES, M. & KUNST, M. 1988. New excavations in Casada Moura and at Bocas, Portugal. Arqueologia 18:65-95. STRAUS, L., AL TUNA, J. & VIERRA, B. 1990. The concheiro at Vidigal: a contribution to the late Mesolithic of southern Portugal. In Contributions to the Mesolithic in Europe, edited by P. Vermeersch & P. Van Peer, pp. 463-474. Leuven University Press, Leuven. STRAUS, L. & GONZALEZ MORALES, M. 1996. Preliminary excavations in El Miron Cave. Old World Archaeology Newsletter 20(1):14-18.

STRAUS, L., OTTE, M., LEOTARD, J., GAUTIER, A. & HAESAERTS, P. 1992. Middle and early Upper Palaeolithic excavations in southern Belgium: a preliminary report. Old World Archaeology Newsletter 15(2): 10-18. STRAUS, L., ALTUNA, J., FORD, D., MARAMBAT, L., RHINE, J.S., SCHWARCZ, H. & VERNET, J-L. 1992. Early farming in the Algarve: a preliminary view from two cave excavations near Faro. Trabalhos de Antropologia e Etnologia 32:141-172. TOUSSAINT, M., RAMON, F. & DEWEZ, M. 1996. L'ossuaire mesolithique ancien de la grotte de Claminforge a Sambreville. In Actes de la Quatrieme Journee d' Archeologie Namuroise, edited by J. Plumier & M. Corbiau, pp.19-32. Ministere de la Region Wallonne, Namur. VEGA DEL SELLA, Conde de la 1916. Paleofftico de Cueto de la Mina. Memoria 13. Comisi6n de Investigaciones Palaeontol6gicas y Prehist6ricas, Madrid. VEGA DEL SELLA, Conde de la 1930. Las Cuevas de la Riera y Balmori. Memoria 38. Comisi6n de Investigaciones Palaeontol6gicas y Prehist6ricas, Madrid. ZILHA.O, J. In press. The Upper Palaeolithic of Portugal: past research and current perspectives. In Recent Research on the European Palaeolithic, edited by E. Webb. British Archaeological Reports, Oxford.

STRAUS, L., OTTE, M., LACROIX, P., LEOTARD, J., GAUTIER, A., SUMMERS, J., ORPHAL, J. & PERNAUD, J. In press. L' Abri du Pape. In Epipaleolithique et Mesolithique en Europe, edited by P. Bintz. Comite des Travaux Historiques et Scientifiques, Paris.

8

Cave versus Open-Air Settlement in the European Upper Palaeolithic Pavel M. Dolukhanov

Abstract The maximum of the Last Glaciation (25,000-15,000 BP) was marked by a significant increase in population density throughout periglacial Europe. Three zones of major concentration of Upper Palaeolithic population are distinguished: (1) Atlantic Europe characterized mainly by cave and rockshelter sites; (2) Central Europe - in which open-air sites predominate; and (3) the East European Plain - where only open-air sites occur. Comparative locational analysis of the sites in Zones 1 and 3 reveals many common features. In both areas the sites were located along water courses, and there is good evidence that caves and shelters were occupied seasonally. The main distinction between the two zones is the occurrence of dwelling structures at many East European sites. These structures have no direct analogues in Atlantic Europe. Soffer (1985) has suggested that large dwellings made of mammoth bones found at large complex base camps were 'ideotechnic artifacts' which reflected ritual behaviour. It is significant that in the Perigord, there exists a direct relationship between large Magdalenian cave sites and the occurrence of portable art (White 1985). It is concluded that a substantial degree of similarity in social behaviour existed between Upper Palaeolithic populations in western and eastern parts of Europe. Both populations consisted of small, seasonally-migrant groups. The existing cave- and open-air sites in both areas are the remains of short-term occupations. The larger sites were predominantly centres of ritual behaviour and of ideological power. A large proportion of sites located at low topographic levels have been destroyed by erosion.

The coldest stages of the Last Glaciation (25,000-15,000 BP) were marked by a notable increase in population density throughout periglacial Europe, evidenced by an increased number of Upper Palaeolithic sites. Three zones of major concentrations of Upper Palaeolithic population may be distinguished (Fig. 1): (1) Atlantic Europe, including Franco-Cantabria; (2) Central Europe, including the loessic plains in the Danube, Vistula and Odra basins, the Central European uplands and the Carpathians; and (3) the East European Plain, including the valleys of the Dniepr and Don and their tributaries. Zone 1 featured predominantly cave and rockshelter sites, with a limited number of open-air sites. Zone 2 contained mainly open-air sites and a limited number of caves. In contrast, Upper Palaeolithic deposits in zone 3 are found almost exclusively in the form of open-air sites The only exceptions are the cave sites in the extreme east and north-east of the Russian Plain - in the southern Urals (e.g. Kapova decorated cave) and the valley of the Pechora river (Bear Cave). Comparative analysis of the Upper Palaeolithic sites in zones 1 and 3 reveals numerous common elements.

(ca 19,200 BP) and Lascaux (ca 18,000-16,630 BP) interstadials - forests with willow, alder and lime spread in some areas, while steppe-like vegetation remained generally dominant (Leroi-Gourhan 1980). On the Russian Plain, as palynological evidence (Velichko 1973) shows, the principal area of concentration of Upper Palaeolithic sites corresponded to a so-called periglacial-forest zone which combined periglacial habitats (arctic steppe with rare trees like dwarf birch) with forests along water courses. Detailed analysis (Spiridonova 1991) shows that during interstadial episodes, the forests on the floodplains included pine, birch, alder and willow. The interfluve areas were taken up by treeless steppe with the participation of wormwood and Chenopodiaceae. Thus, taking into consideration local climatic peculiarities - Atlantic Europe was considerably wetter - the general pattern of vegetation was similar in both zones. There is evidence that periglacial landscapes in Europe had a very high ungulate biomass.

Vegetation

Palaeolithic cave sites and rockshelters in the Perigord were located mainly in limestone rocks facing the river valleys. As White notes, 'only 17 of the 272 Upper Palaeolithic occurrences were situated more than 1000 meters from the nearest river. For the Magdalenian, only 6 out of 86 occurrences fall beyond this 1000 meter limit' (White 1985:90). It is significant that the same observation applies to open-air sites; only 5 out of 46 open-air occurrences are situated more than 1000m from the nearest river. Summarizing the

Geomorphology

Let us start with zone 1 as exemplified by the Perigord, the classic Upper Palaeolithic area in south-west France. Palaeobotanical evidence shows that during the Upper Palaeolithic (corresponding roughly to Wurm III and IV) the vegetation of this area was characterized by a varying combination of steppe and forest. During the comparatively short-lived warmer climatic episodes - such as the Laugerie

9

Dolukhanov

Figure I

Upper Palaeolithic Europe. Key: shaded area - extent of the ice sheet at glacial maximum; A - sites in southwestern France; B - sites in Central Europe; C - sites in the Upper Dniestr area; D - sites in the Middle Dniepr area; E - sites in the Kostenki-Avdeevo area; F - Sungir area; G- Bear Cave area; H- Kapova Cave. 1 -zone 1; 2- zone 2; 3 - zone 3. After 0. Soffer 1985.

available evidence, White concludes that 'the available data support the idea of a river valley focus in the settlement strategy during the Upper Palaeolithic in the Perigord' (White 1985:93). It is particularly significant that in the Perigord the sites were usually located near fords. According to White, proportionately more Magdalenian and especially Solutrean occupations are located within 1000m of a known ford than in any preceding period (White 1985:119). All major Upper Palaeolithic sites on the East European Plain tend to be located in areas where river valleys formed chains of lakes with numerous crossings that could be used by aggregations of herd animals. It has been noted that all Upper Palaeolithic sites on the Russian Plain are linked to the

hydrological network. Soffer (1985) has described the linear patterning of Upper Palaeolithic settlements related to water courses. She noted that the sites usually clustered on the banks of lake-like widenings of river valleys via which the meltwater from glaciers was channelled to the sea. Thus the area of Kostenki on the River Don, which has an abnormally high concentration of Upper Palaeolithic sites, coincides with a large depression which originally was 35km wide (Krasnov 1982). Kvasov (1979) concluded that during the Late Pleistocene the valleys of large rivers formed 'chains of lakes' the level of which fluctuated seasonally. A similar situation occurred in the vicinities of major Palaeolithic sites in the basin of the Dniepr Dobranichevka, Mezherich, Timonovka, Khotylevo and

10

Cave versus Open-Air Settlement in the European Upper Palaeolithic

juveniles made up nearly 50%, and females were twice as numerous as males among the adults. A close similarity with the population of Berelekh may be noted. The main distinction is the higher proportion of adults in the Kostenki sites. In general the faunal evidence from all parts of Europe demonstrates the great diversity of the animal world and the high ungulate biomass. The location of Palaeolithic sites in the immediate vicinity of fords and river-crossings suggests that a considerable proportion of the carcasses of large animals were procured not by hunting but by scavenging the corpses of naturally perished animals.

others (Velichko et al. 1981). Considerable areas in the immediate vicinity of the sites were taken up by swampy floodplains with dunes and numerous oxbow lakes. It is probable that fords and crossings occurred there. It is equally probable that large animals often drowned while trying to cross the river. In this respect the case of the Berelekh mammoth 'cemetery' is highly significant. According to Vereshchagin (1977), the cemetery was located in the middle stretches of the Berelekh river, in Y akutia (Central Siberia) at 71° O' N and 145° O' E. The bones occur in a 2m-thick sequence of laminated muds with abundant plant remains and broken by ice wedges. Based on the character of deposition it was suggested that the 'cemetery' was formed in an oxbow lake. The alternating layers indicate repeated changes in the sedimentation. The excavations carried out during one month in 1970 brought to light some 8500 mammoth bones belonging to at least 140 individuals. Of these 60% were females. The faunal remains also included about a thousand bones of wolf, glutton, cave lion, hare, horse, woolly rhinoceros, reindeer, bison and willow grouse, with juveniles being most commonly represented. Table 1

Seasonality There is increasing evidence (White 1985; Koetje 1987) that Upper Palaeolithic caves and rockshelters in south-west France were occupied seasonally. There is evidence to suggest that this was also the case with the sites in zone 3. Based on archaeological and archaeozoological sources, Soffer (1985) concludes that all the Upper Palaeolithic sites of the Russian Plain were 'seasonally occupied and do not represent permanent settlements occupied for years and decades' (Soffer 1985:348). A similar view was taken earlier by Vereshchagin who wrote: 'Upper Palaeolithic dwelling sites were of seasonal character; only winter camp-sites were relatively stable ... This is proven by the absence of juvenile individuals of arctic fox, wolf, fox, hare and reindeer. The skeletons and skulls of wolves at the site of Kostenki 12 belonged to 8-9 months old individuals. Consequently they were killed in the middle of winter' (Vereshchagin 1977:85). Summarizing the existing evidence, Soffer distinguished several types of Upper Palaeolithic occupations on the Russian Plain (Table 2).

Minimum numbers of mammoths represented at Upper Palaeolithic sites in Zone 3.

Site

Yeliseyevichi Mezin Gontsy Kirillovka Dobranichevka Mezherichi Anosovka Kostenki 8 Kostenki 11 Kostenki 14

Number of individuals 60 116 25 25 18 95 32 10 42

Table2

9

Types of Upper Palaeolithic occupations on the Russian Plain (after Soffer 1985). Cold-weather occupations

Economy

complex base camps simple base camps hunting base camps

The economy of Upper Palaeolithic settlements was based on the hunting of large herd herbivores. The Perigord sites show a clear predominance of reindeer (Rangifer tarandus). Apart from this dominant species, the Upper Palaeolithic fauna was quite varied; it included mammoth, woolly rhinoceros, wild ass, wild horse, saiga antelope, wild boar, roe deer, red deer, aurochs, Pyrenean ibex, arctic fox, wolf, and many other mammals, as well as fish and birds. The Upper Palaeolithic open-air sites of the Russian Plain show a clear dominance of mammoth. The minimum number of individuals of mammoth established at Upper Palaeolithic sites in zone 3 is shown in Table 1. Approximately 59% of the mammalian bones from sites deposited in the upper loess of the Kostenki region (ca 22,000-15,000 BP) are from mammoths. Apart from this species, the hunting prey of Upper Palaeolithic sites included reindeer, woolly rhinoceros, horse, wild ass, wolf, polar fox, red deer, wild boar, bison, musk-ox and other mammals, as well as fish and birds. Among the mammoths,

Warm-weather occupations

simple base camps lithic workshops hunting stations bone-collecting stations

'Dwellings' The main cultural distinction between the two zones lies in the presence/absence of 'dwelling structures' that have been identified by Russian prehistorians (Efimenko, Rogachev, Praslov, Grogoriev) at numerous East European sites since the 1930s. These structures have no direct analogues in Atlantic Europe. The 'dwellings' are usually either ovalshaped or circular. 'Complex 1' at Kostenki 1 is 30m long x 15m wide. The structure may include exterior and interior hearths, exterior storage pits, and numerous finds of archaeological material. In some cases a clear patterning or sorting of mammoth bones may be observed. Thus structure 1 at Mezherich features an 'outside facing' of mammoth skulls, while 95 mandibles stacked one on top of

11

Dolukhanov the other formed a 'herringbone' pattern (Soffer 1985:369). A distinct patterning of bones is observable at several sites in the area of Kostenki (Praslov & Rogachev 1982). According to Soffer, dwellings combined with storage pits occurred only in the cold-weather complex base camps. There are some remains which have been interpreted as evidence for the existence of dwelling structures within caves and rockshelters in France. Two hearths in Lazaret Cave have been interpreted as evidence of a tent (Lumley 1976). Five regularly-spaced hearths at Abri Pataud were interpreted as the remains of a longhouse (Leroi-Gourhan & Brezillon 1972). In these and similar cases there are serious arguments against such interpretations (Gamble 1986). There are several unclear signals relating to the existence of dwelling structures at open-air sites in zone 1. The most notorious case is that of Pincevent, where Leroi-Gourhan & Brezillon (1966) recognized indications of circular tents around three hearths. On several occasions, so-called stone pavements were identified at Magdalenian sites. These include Solvieux and other Magdalenian sites in the Perigord, sites in the Paris Basin, as well as sites in the Rhineland in Germany (e.g. Gonnersdorf). In all these cases, serious difficulties were met when attempts were made to interpret them as dwelling structures. As Koetje wrote:

containing four or more compositions; there occur small groups consisting of 2-3 compositions, as well as isolated pictures. It is highly significant that the excavations in one of the galleries of the first level revealed cultural deposits with a maximum thickness of 15-20cm which contained fragments of charcoal and 201 artifacts manufactured from flint, jasper, limestone and sandstone. The tools included points, end-scrapers, and retouched blades. A series of three small serpentine beads is remarkable. The fauna included mammoth, cave bear, European hare, wolf, and arctic fox. The pollen content reveals an environment of periglacial forest-steppe. Two radiocarbon dates were obtained for samples of charcoal: 14,680±150 BP (LE-3443) and 13,390±300 BP (GIN-4853). Hence, the age, the palaeoenvironment, the topography, the content and the technique of cave art at the Kapova Cave may be compared to those of Upper Palaeolithic Franco-Cantabria. The Kapova Cave is not a unique phenomenon. There are indications of the existence of more decorated caves in the Urals and other regions of zone 3 (Stolyar 1985).

Ceremonial sites For a long time there has been abundant evidence for the use of Palaeolithic caves as places of ritual and ceremony. Suggestions were made that concentrations of specifically arranged bones of cave bear at several caves suggest the existence of a 'bear cult'. Such caves were first identified in the Swiss Alps (Drachenloch and others) as well as in France (Regourdou). Abnormally high concentrations of bear bones at the Caucasian cave sites are also seen as evidence for the existence of a similar cult. At one of the cave sites (Tsutskhvaty, western Georgia) the skulls of cave bear were apparently intentionally arranged around the walls (Stolyar 1985). Indications of ritual use are particularly common in the case of decorated caves. Numerous footprints discovered at the Niaux cave complex have been interpreted as evidence of 'initiation ceremonies'. This was 'achieved through the art and associated ritual that included the journey and most probably dancing and music' (Gamble 1986:235). It is also significant that in the Perigord there exists a direct relationship between large Magdalenian cave sites and occurrences of portable art (White 1985). Recently Soffer (1985) put forward the suggestion that 'monumental architecture' in the form of 'dwellings' made of mammoth bones at 'large complex base camps' on the Russian Plain were 'ideotechnic artifacts' which reflected ritual behaviour. The arguments in favour of such an interpretation include both the highly-patterned architectural details and the topography of the sites. Also, these sites normally include considerable amounts of portable art and decorated bone pieces, some of which have been interpreted as ritual musical instruments (Bibikov 1969). If we extend Soffer's arguments, which seem fairly convincing, to a wider corpus of evidence, it may be suggested that both the 'large complex base-camps' and large decorated caves are the 'ideotechnic artifacts' i.e. the remains of ceremonial centres, the sources of ideological power.

'It was also not possible to unambiguously identify potential hearths. In several cases there are essentially identicallysized 'circular cobble features' that resemble traditionally identified hearths; in most of these cases the cobbles in the features show signs of having been burned. Unfortunately, such is the case for many of the cobbles at these sites. In the absence of other evidence there is no compelling reason to accept them as hearths ... It is very possible that whatever hearths once existed did not survive, or were located beyond the boundaries of the excavated area. In no case however, does the extant spatial patterning of tools suggest a restricted portion of a larger pattern along the lines of Binford's or Yellen's camp site structures' (Koetje 1987:166-167).

Art The occurrence of decorated caves is traditionally regarded as one of the particular cultural features of zone 1. The greatest density of decorated caves is observed in the Dordogne and seems to cluster within 30km of Les Eyzies (Gamble 1986:232). The discovery of decorated caves in the Urals proves that this phenomenon occurred also in at least this region of zone 3. The paintings in the Kapova Cave (Shchelinski 1989) were found in four separate galleries at distances between 170m and 300m from the entrance of the cave. Three galleries are located on the first level (Domed Hall, Hall of Signs, Hall of Paintings); the fourth (Hall of Chaos) is on the second level. The paintings vary in length from 0.58 to 1.06 metres. All the paintings were made in ochre applied directly on the limestone surface. Among the animals mammoth, horse and rhinoceros could be identified. Signs and geometric symbols (trapezes, triangles, a truncated cone, a square with loops) are more numerous. Paintings form large groups, each

12

Cave versus Open-Air Settlement in the European Upper Palaeolithic Suomalainen Tiedeakatemia, Helsinki. LEROI-GOURHAN, A. 1980. Interstades wiirmiens: Laugerie et Lascaux. Bulletin de l' Association Franqaise pour l' Etude du Quaternaire, 1980-83:95-100. LEROI-GOURHAN, A. & BREZILLON, M. 1966. L'habitation magdalenienne no. 1 de Pincevent, pres Montereau. Gallia Prehistoire 9:263-285. LUMLEY, H. de (ed.) 1976. La prehistoire franqaise, vol. 1. CNRS, Paris. PRASLOV, N.D. & ROGACHEV, A.N. (eds) 1982. Paleo/it Kostenkoysko-Borshevskogo raiona an Donu. 1879-1979. Nekotorye itogi poleyyh issledovanii (The Palaeolithic of the Kostenki-Borshevo region on the River Don 1879-1979. Results of field studies). Nauka, Leningrad. SHCHELINSKY, V.E. 1989. Some results of new investigations at the Kapova Cave in the southern Urals. Proceeding of the Prehistoric Society 55:181-191. SPIRIDONOVA, E.A. 1991. Evolicija rastitel'nogo pokrova basseina Dona v verhnem pleistocene-holocene (The evolution of the vegetational cover in the region of the Don in the Upper Pleistocene and Holocene). Nauka, Moscow. SOFFER, 0. 1985. The Upper Palaeolithic of the Central Russian Plain. Academic Press, Orlando. STOLJAR, A.D. 1985. Proiskhozhdenie izobrazitelenogo iskusstva (The origin of figurative art). Iskusstvo, Moscow. VELICHKO, A.A. 1973. Pririodnyi process v pleistocene (The natural process in the Pleistocene). Nauka, Moscow. VELICHKO, A.A., LUBIN, V.P. , PRASLOV, N.D. & KURENKOVA, E.I. (eds) 1981. Archaeology and Palaeogeography of the Late Palaeolithic of the Russian Plain. Nauka, Moscow. VERESHCHAGIN, N.K. 1977. Berelehskoe 'kladbishche' mamontov (The Berelekh mammoth 'cemetery'). In 'Mamontovaja fauna Russkoi Rayniny i Vostochnoi Sibiri' (The mammoth fauna of Eastern Europe and eastern Siberia), edited by 0.A. Skarlato. Proceedings of the Zoological Institute 72:5-50. WHITE, R. 1985. Upper Palaeolithic Land Use in the Perigord. A Topographic Approach to Subsistence and Settlement. (BAR International Series 253). British Archaeological Reports,

Conclusions In this paper it has been suggested that there existed a substantial degree of similarity in the social behaviour of Upper Palaeolithic populations in both the western and eastern parts of Europe. In both cases the population consisted of small, seasonally migrant groups. The majority of existing cave and open-air sites in both areas are the remains of comparatively short-term occupations. The larger sites - both the open-air sites with 'dwellings' and the decorated caves - were predominantly centres of ritual behaviour and ideological power. It may also be suggested that a large proportion of the sites located at low topographic levels were subsequently destroyed by erosion.

References BIBIKOV, S.N. 1969. Drevneishii muzykal'nyi kompleks iz kostei mamonta (The most ancient musical instruments of mammoth bones). Naukova Dumka, Kiev. GAMBLE, C. 1986. The Palaeolithic Settlement of Europe. Cambridge University Press, Cambridge. KOETJE, T.A. 1987. Spatial Patterns in Magdalenian Open-air Sites from the Isle Valley, Southwestern France (BAR International Series 346). British Archaeological Reports, Oxford. KRASNOV, I.I. 1982. Geologo-geomorfoligicheskoe stroenie doliny Dona i razmeshchenie paleoliticheskih pamiatnikov (The geological-geomorphological structure of the Don Valley and the setting of Palaeolithic sites). In Paleo/it Kostenkoysko-Borshevskogo raiona an Donu. 1879-1979. Nekotorye itogi polevyh issledovanii (The Palaeolithic in the Kostenki-Borshevo Region on the River Don. 1879-1979. Results of Field Investigations), edited by N.D. Praslov & A.N. Rogachev, pp. 37-41. Nauka, Leningrad. KVASOV, D.D. 1979. The Late Quaternary History of Large Lakes and Inland Seas of Eastern Europe. Annales Academiae Scientiarum Fennicae, Ser. A III, Geol-Geogr., 127.

13

The Human Use of Caves in the Belgian Palaeolithic Marcel Otte (Translated by L.G. Straus)

Abstract

In Belgium caves, found only in the southernpart of the country (Wallonia),were used as shelters by Palaeolithichumans. In contrast, the central region has both abundant flint sources and plateaux with open steppes rich in Pleistoceneherbivorousgame. Palaeolithic sites are known in both regions, but they are radically different in nature. Among many of the caves, the relative distance to lithic raw material sources caused intensive use and reuse of the tools that, of course, varied by period. Living space was naturally limited in the caves, leading to the cumulativedepositionof very dense cultural and fauna/debris and to the blurring of site structural organization.In contrast, the open-air sites are self-structuredand the activity organizationof surfaces is archaeologicallymuch more visible and more common than among cave sites of the same periods. The importancethat has been accordedto Palaeolithicsites in caves is obviously linked to the history of the discipline,since such sites are much easier to find than open-airloci. However, the phenomenonof 'findability'must have operatedalso during the Pleistoceneand the great accumulationof residential debris in caves may have been naturallyprovoked by the appeal of these shelters in what was then a largely treeless landscape. The extent of cave use by Palaeolithicpeople in Belgiumis accentuatedby their ease of discoveryin past and present. In addition, the nature of these occupationsis deformedby the constrainingcave walls. Thesehumancave occupationswere in some respectssimilar to those of hyaenas and bears whose remains are so often mixed in with those of humans and their culture. So far, in Belgium we have no evidence of deep penetration of caves by humans (possiblyfor artistic, magico-religiousreasons, as was common in southern France and Spain).

known abroad mainly for their Neolithic flint mines, such as Spiennes in the west and Sainte Gertrude in the east, to cite just the largest ones. There are no caves or rockshelters in this region. The chalk deposits have been extensively covered by Pleistocene loess to a thickness ranging from a few to ten metres or more. In this region, they were transformed into clay or loam soils by Holocene weathering, and are nowadays heavily exploited for agriculture. This began in the Early Neolithic at the end of the eighth millennium BP, when Bandkeramik invaders first settled on these fertile soils. However, many open-air Palaeolithic sites embedded in these loess deposits at different depths and of different ages, from Acheulean to Mesolithic, are known. It is then interesting to compare, for each period or each 'cultural tradition', the archaeological remains from regions 1 and 2. Moreover, for long periods of the Ice Age the plateaux of Middle Belgium were characterized by an open landscape dominated by grasses. Therefore, they provided ideal terrain and food supply for herbivores, and thus excellent hunting territories for man. The third strip, to the north, adjacent to the present North Sea, is covered by thick sandy deposits of younger age, which neither facilitate discovery nor favoured Pleistocene human occupation. However, it must be borne in mind that this region was widely open in the glacial age to the northern European territory, linking the Continent to Great Britain across the North Sea and the Channel, then above sea level. It has thus a crucial importance for the study of cultural contacts between different provinces that are now widely separated.

Though small, Belgium possesses a long and complex geological sequence covering many of the stages found in the European continent (Heinzelin 1949). For this presentation, the territory may be divided into three parts corresponding to three parallel strips running from southwest to northeast and descending in altitude from south to north. The southernmost and highest strip is formed of old Primary rocks: quartzites and sandstones in the extreme east, then extensive limestone deposits, folded along this strip and sometimes divided by schist formations. Inside the last folds, the main rivers flow northwards: the Meuse and the Sambre, sometimes creating high cliffs that are very typical of this landscape (Fig. 1). The cliffs and the ridges have been truncated by numerous smaller rivers, crosscutting the axes of the main streams. This situation has led to the formation of a large number of rockshelters, caves and, in general, karstic formations either opening out into the flanks of the rivers or very deep and difficult to enter even by speleological means. Many of these natural shelters have been occupied by man or animals, or both. It is important to stress the great diversity in the topographical settings of these 'caves', in their altitude, proximity to water, accessibility, dimensions, orientations, and so forth. As there is no preferential model caused by natural processes, so diverse in their action, we should not expect any kind of regularity in their occupation either. The second region or 'strip' is called 'Middle Belgium' and consists of plateaux formed by thick chalky deposits of the Cretaceous era. These formations contain abundant high quality flint nodules. The flint outcrops have, of course, been intensively exploited throughout prehistory, but are

14

The Human Use of Caves in the Belgian Palaeolithic

.

..

..,,..-··-..~ r

\......· ··,~

f;;::,::.::;, .' ' -~ ~

-,,,.~"""""'~~~)

a~-~

I - -~

€::::r~ 73

Oosterbeek . :·;:_

- . - - - .,. .. . --- -·... ·. · .. ·.

-~ - .

~~.;-

,

~1;i.U::::,.I""-"

-

. :..--=~

..

~-~_,.=,,:~ .

~;y•

~-·· ...:-:-- _;~~~~~~ ·.. . : . ·'..

g;~

\

'

Figure 4

Late Neolithic (5th millenniwn BP) pottery from the Gruta dos Ossos (Tomar).

earlier times no definite pattern existed in the selection of the caves, in the late Neolithic caves tend to be grouped in small areas, defining necropoleis. This is the case in the Nabao valley (Oosterbeek 1991; Arsenio & Batata 1992), where all the caves (and only the caves) in a small part of the valley (called the 'canteir6es') were used for burial. The same occurs in other valleys, such as the Almonda (Zilhao et al. 1991), the 'Rio Maior' (Senhora da Luz - Gon9alves 1978), the Ribeira dos Crastos (Ferreira et al. 1977), or around Alcanena (Sa 1959; Gon9alves & Pereira 1977; Oosterbeek 1988), although isolated cave burials also exist. The necropoleis of the 'canteir6es' of the Nabao valley have been systematically surveyed. The Nabao flows roughly north-south, except for this 600-metre portion where it flows from west to east before resuming its 'normal' orientation. All the caves on both sides of the river were used for collective burials in the fifth millennium BP. This tendency for grouping of the necropoleis is clearly of megalithic influence (Jorge 1986, 1989) and is to be observed also with passage graves, false corbelled chambers (Leisner 1959) or artificial caves (Leisner et al. 1961). In the Nabao, all the burials share their collective nature, the technology of the artifacts associated with them, and most of the artifactual assemblages. Yet, significant differ-

regional distribution and different artifactual assemblages, and this other Neolithic pattern, dominant inland (in contrast to the coastal caves) is the key to understanding the whole Neolithic process. · The region we are concerned with also has megaliths and constitutes the meeting area of both traditions (Oosterbeek 1991). One of these, Anta 1 de Val da Laje, is a small passage grave beneath an oval mound. It was built on top of an early Neolithic habitation. Its low passage (0.80m high) was approached through an external atrium with two altars and a narrow entrance 0.50m wide (Drewett et al. 1992; Oosterbeek et al. 1992). This hierarchy of spaces, restricting access to the sacred area and even conditioning the physical behaviour of visitors who had to bend before entering the chamber, is a prototype for the kinds of cave burials that are found in the fifth millennium BP, thus providing an insight into the relations between these two Neolithic traditions, to which I shall return.

The Late Neolithic In the early fifth millennium BP individual burials in caves were replaced by collective ones (Zilhao 1984). Another major change relates to their spatial distribution. Whereas in

74

Back Home! Neolithic Life and the Rituals of Death in the Portuguese Ribatejo

artefacts assembl.

burial single

Early Neolithic

Middle Neolit.

• •

Late Neolithic

Chalcolithic

collective

• •

similar

variable

• •

Beakers

Early B.A.

Figure 5



• above 75% •

25% to 75%

closed space

restricted passage

• • •• •• @

restricted atrium

mound or cave

• • • ••• •

under 25%

Typology of late prehistoric burials from the Portuguese Ribatejo: 1. Early Neolithic - individual burial in pit with artifacts; 2. Early Neolithic - individual burial in stone structure with artifacts; 3. Middle Neolithic - collective burial in megalithic chamber with artifacts; 4. Middle Neolithic - collective burial in passage grave with artifacts; 5-6. Late Neolithic - collective burial in monumental passage grave with artifacts; 7. Late Neolithic - collective burial in cave with artifacts; 8-10. Beaker/ Early Bronze Age - individual burial in stone structure with no, few or many artifacts.

ences in these assemblages are evident if we look at decorated vessels and other 'prestige goods'. Discoidal beads are an example; these are present in all assemblages, but made of bone in one case, and of different types of stone in each of the others. Some types are exclusive to certain burials, such as textile implements (found only in Gruta do Cadaval), or flutes and Atlantic shells (in Gruta dos Ossos), or red deer bones (in Gruta do Morgado). These particularities may reflect the status of the people buried in those caves, and hence representing a further stage in social differentiation. The rituals also become.more complex and, interestingly, show strong parallels to megalithic ones. A good example is Gruta dos Ossos (Fig. 4), radiocarbon dated to the second quarter of the fifth millennium BP (ICEN-465: 4630±80 BP). In this cave, no previous or later occupations existed and natural disturbance was minimal, thus enabling a very detailed record of the fossil evidence of the original

structure. It is a small cave, 20m2 in area with a small entrance 1.20m wide and 0.90m high, facing south. The dead were originally buried near the back wall of the cave, but were later re-buried near the entrance. The base of this ossuary was levelled with stones and sand, defining a platform over which the pelvis and sacrum bones of the individuals were deposited. This first layer was then covered with sand, and a second layer comprising the femur, tibia, and other leg bones. After a new layer of sand, a third deposition included the humerus, radius and cubitus, also covered with sand. Finally, the skulls and mandibles ·were placed in an arc and left uncovered. A line of blocks was then built around this deposit and the artifacts deposited outside it in the entrance (Oosterbeek & Cruz 1988; Cruz 1992). This scenario is very similar to the one mentioned above for the passage grave. The low entrance had its access restricted by a ritual 'wall', with an external small atrium,

75

Oosterbeek

1

2

-------,

'

I :_______:: '

:

•□

6

3

4

5

I

I

I

-■

--

I

I

[i] □ □ 8

Figure 6

9'

10

Matrix of burial typology criteria.

and access to the ossuary was not possible without bending (or even crawling!). However, not all collective burials are ossuaries, and even the ossuaries differ one from another. What seems clear is the shift towards collective burial.

The use of caves also changes. Still used for burying individuals, likely to be usually of high status (after the rich artifact assemblages associated with them), they are also used for episodic 'camping' (like Gruta do Cadaval - Cruz & Oosterbeek 1985), or even for permanent habitation (like Gruta da Avecasta, a very large cave that included part of a Bronze Age village - Mateus 1982; Santos 1985). All in all, caves lose a large portion of their previous symbolic importance. Their modern uses for goods storage or as treasure hiding places, are already present in the fourth millennium BP. The questions that arise now are - why did these changes occur, and what do they mean?

The Early Bronze Age This dominance of 'death' over 'life' structures starts to change in the fifth millennium BP, and this represents a major change in the social structure of the living communities. In the late fifth millennium BP, associated with Beakers and, later, with the Early Bronze Age, individual burials reemerge and gradually replace collective ones (Gom;alves 1972; Oosterbeek & Cruz 1992). They are different from the early Neolithic ones, however. Found in caves in small numbers, they can also be grouped in large open-air necropoleis, and the individual ranking becomes very clear. This new approach to death is associated with a more stable settlement pattern and is a result of a major economic revolution, referred to by Sherratt as the Secondary Products Revolution (Harrison 1985). A new perception of the landscape is achieved with the possibility of systematic exploitation of all kinds of agricultural soils. Territories are no longer marked by funerary monuments (caves, megaliths, etc.), but by stone-built settlements.

Neolithic life: where is home? Human perceptions are a result of sets of negative assumptions. In this sense, there is not a great difference between the history of human beliefs and the present scientific method. We assimilate data input to existing structures (cognitive, social, economic, etc.), and the same happened in prehistory. The importance of caves cannot be understood outside the social context of Neolithic life. The study of the Alto Ribatejo region, together with the most recent research in the Iberian Peninsula, suggests that at least one of the early Neolithic inputs was associated with

76

Back Home! Neolithic Life and the Rituals of Death in the Portuguese Ribatejo

the Cardial Impressed Ware spread. This early Neolithic phenomenon, radiocarbon dated in Tomar to the midseventh millennium BP is (and this is essential) what decades ago was called the 'full Neolithic package' including domesticated plants and animals and the basic technological innovations. This Neolithic combined a marginal agriculture, largely opportunistic, with the survival of previous predatory strategies. The Alto Ribatejo is a region that combines in a relatively small area all the main geomorphological units of the peninsula, hence providing the ideal location for this eclectic strategy. The archaeological record suggests that this Neolithic was so successful that it enabled population growth without major social or economic changes. As a result, whereas the region was prosperous, it was also conservative, and by the early fifth millennium BP it went into decline, clearly seen at the technological level. Further south, in the Alentejo, another early Neolithic, possibly of Andalucian origin and unrelated to the Cardial spread, confronted by a more homogeneous landscape, was condemned to the full Neolithic specialization, and thus generated an equally successful but less stable process from which megaliths and, later, the Beakers and the Bronze Age emerged. Caves existed and were in use in the first of these processes, but their function was determined by the dominant, megalithic process. I think that despite retaining a relative isolation, the coastal Neolithic was still a part of the whole Neolithic system, and its ideological guidelines were common. The Cardial tradition of the use of caves for habitation was broken by this prevailing model according to which the rhythm of life is focused in death. This is the 'why' of the use of caves for burials. But what do its changes signify? Again, the explanation must be sought on a broad scale including the Iberian Peninsula and the west Mediterranean. Across this entire region a single Neolithic tendency can be discerned made up of numerous specific processes - the unity of diversity. My explanation is based on a series of assumptions:

(v) the major collective investment of production surplus was made in burials (in caves or megaliths), these being collective (corresponding to the lineages); (vi) by the fifth millennium BP, the increasing complexity of the power network in some areas (e.g. Millares, Zambujal) enabled some of the lineages to impose a settlement ranking, entering a new stage in the investment of the surplus, this time also in structures for the living, real centres of power - the hillforts of Mediterranean type; burials were still collective, but a greater degree of ranking can be seen among them; (vii) at the end of the fifth millennium BP, and throughout the fourth millennium, power relations became more stable, and a third and decisive stage was entered as far as the surplus is concerned - monumental structures ceased to be built; internally ranked settlements made their appearance, together with individual burials. Thereafter, a new social structure was achieved (or imposed!) - what is usually termed the state. By the middle of the fourth millennium BP caves had lost their previous significance This reflects the new social structure and is a result of a slow but clear process of substitution of the lineages ranking by the ranking of individuals. Acknowledgements: Most of the fieldwork and subsequent data processing were carried out in association with Ana Rosa Cruz who also read earlier drafts of this paper. Many of the ideas put forward in the paper were also discussed with Peter Drewett of the Institute of Archaeology, University College, London, and Vitor Oliveira Jorge of Oporto University, while I was working on my PhD. Finally, my participation in the Human Use of Caves conference would not have been possible without the support of the British Council and the Instituto Politecnico de Santarem where I teach.

References ARSENIO, P. & BATATA, C. 1992. 0 desenvolvimento da espeleologia na Regiao de Tomar. Boletim Cultural da Camara Municipal de Tamar, 16:12-29. BONIFAY, E. 1955. Methode d'etude du remplissage des grottes. Bulletin de la Saciete Prehistorique Franqaise 52:144-145. BONIFAY, E. 1956. Les sediments detritiques grossiers dans le remplissage des grottes - methode d'etude morphologique et statistique. L'Anthropologie 60:447-461. BOSCH-GIMPERA, P. 1932. Etnologia de la Peninsula Jberica. Barcelona. BOUILLON, M. 1972. Decouverte du Monde Souterrain. Robert Laffont, Paris. BOULE, M. 1892. Notes sur le remplissage des cavernes. L'Anthropologie 3:19-36. BREUIL, H. 1952. Quatre Cent Siecles d'Art Parietal. Centre d'Etudes et de Documentation, Montignac. CASTRO, M.J.M. 1973. Subs{dios para a Carta Arqueol6gica do Concelho de Tamar. Unpublished BA dissertation, University of Lisbon. COUTINHO, A.P. 1985. Gruta das Andorinhas. Boletim de divulgaqiia de actividades espeleol6gicas 3/4:1. CRUZ, A.R. & OOSTERBEEK, L.M. 1985. A Gruta do Cadaval: elementos para a Pre-Hist6ria do Vale do Nabao. Arquealogia na Regiiio de Tamar, suplemento ao Boletim Cultural e Jnformativa da Camara Municipal de Tamar 1:61-76.

(i) the Neolithic was successful to the point that it enabled population growth (regardless of the fact that it probably resulted from a previous population increase), and was built on a pre-existing information network already in existence in the Mesolithic (represented by lithic typology, settlement and burial structures and art); (ii) the more complex economic activities resulted in a· more complex social structure, and some regions of the west Mediterranean began to generate a power network on top of the previous one; (iii) in this early stage social differentiation was structured in lineages, and the relative power of each one of these, its rank, was still of a collective (though restricted) nature, where the status of the living was anchored in the dead; (iv) settlements were small, in the open air, and made of perishable materials; they could also move easily, within territories in which boundaries were defined by the necropolis;

77

Oosterbeek CRUZ, A.R. 1992. Estudo preliminar do ossuario Neolftico da Gruta dos Ossos. Revista de Ciencias Hist6ricas da Universidade Portucalense 6:91-127. DREWETT, P., OOSTERBEEK, L., CRUZ, A.R. & FELIX, P. 1992. Anta 1 de Val da Laje 1989/90 - the excavation of a passage grave at Tomar (Portugal). Bulletin of the Institute of Archaeology 27:133-148. EL AVESTA. 1974. Textos Relativos al Mazdefsmo o Zordatrismo. Clasicos Bergua, Madrid (Traducci6n: Juan B. Bergua). ELIADE, M. 1969. Le mythe de l' eternel retour. Editions Gallimard, Paris. FERREIRA, O.V., NORTH, C.T. & LEITAO, M.. 1977. 0 Esp6lio Arqueo16gico das Grutas da Ribeira dos Crastos (Caldas da Rainha). Comunicar;oes dos Servir;os geol6gicos de Portugal 61:5-11, 2 plates. FERREIRA, O.V. 1982. Cavernas com Interesse Cultural encontradas em Portugal. Comunicar;oes dos Servir;os geol6gicos de Portugal 68: 285-298. GON30,000year-old occupation in central Australia (Puritjarra), the Nullarbor Plain, and north Queensland (Nurrabullgin). In temperate regions in south-eastern Australia much new archaeological research has explicated the rationale for the choice and use of particular caves as habitation sites. Finally, in Tasmania occupation has now been found extending back throughout the glacial maximum to 35,000 BP, and hand stencils are present in dark chambers deep inside some of these subterranean limestone caves, suggesting ritual as well as habitational use.

A cave is defined in Chambers' English Dictionary as 'a hollow place in a rock', and the term will be used here in this broad sense. The Oxford English Dictionary gives a narrower definition of a cave as 'a hollow place opening more or less horizontally under the ground, a cavern, den, habitation in the earth, hollow place, cavity'. Then there are 'external caves' (Renfrew & Bahn 1991:204) which are rockshelters at the entrance to a cave. A rockshelter may be defined as 'a naturally-formed overhang in a cliff, outcrop or boulder, sheltering a floor area' (Flood 1990:357). The morphology of caves depends on local geology, and in a continent as vast as Australia, varies widely. Whilst they differ markedly in form, depth and amount of natural light, traces of past habitation are common both in rockshelters, usually of sandstone, quartzite or granite, and limestone 'caves', both at the entrance and in caverns far underground. Exclusion of all 'caves' from consideration except those which conform to the O.E.D. definition would be artificial and unduly restrictive, so for the purposes of this paper all three types, underground 'caves' or caverns, external 'caves' and rockshelters will be included, although the morphology of the 'cave' in each case will be made clear. An unbroken tradition of Aboriginal use of caves in Australia extends from more than 50,000 years ago till the present day. Whilst significant environmental changes occurred in the continent during this period, their impact was relatively localized and moderate. Likewise, although

there were changes in technology and other aspects of society, Australian Aboriginal culture is marked by a basic stability and continuity of lifeways and traditions. Clearly, care must be taken to avoid imposing ideas derived from recent societies onto those in the distant past, but, given this caveat, it seems valid to use a combination of ethnographic and archaeological evidence to try to elucidate the traditional use of caves by Australian foragers. At the time of first European settlement in Australia (1788) the principal function of caves was observed to be as camping places, but they were also used as sources of water and minerals, as burial sites, keeping places for sacred objects, and as ritual and mythological sites (Bednarik 1986). Ethnographic and historical evidence on Aboriginal use of caves, however, is regrettably scarce, because of the rapid disruption of traditional lifeways, so this paper will focus on ethnographic rather than archaeological evidence from one of the very few regions where such information can still be obtained. Examples of the traditional use of caves in the ethnographic present will be taken mainly from one Aboriginal group, the W ardaman of the Victoria River District of the Northern Territory, at approximately latitude 15 degrees south of the equator. The Wardaman still practise rock art (at least retouching of paintings), know the mythology associated with the sites, retain some stone tool making skills, and lived in caves during the wet season as recently as the 1940s. Our principal informants are now in their sixties or seventies, but recall moving round the

193

Flood

country in the traditional way from rockshelter to rockshelter each wet season, exploiting local food resources. I have been carrying out fieldwork (from 1988-1992) in the Katherine region as Principal Investigator of the Lightning Brothers Project, involving multi-disciplinary research into past Aboriginal lifeways, the prehistory, rock art and associated mythology of Wardaman country. Their traditional use of caves is described here for the first time, including reasons for the selection of particular caves for habitation and/or other uses and for the presence, location and type of rock art (petroglyphs and pigmented art) in such sites.

ca 7000 BP for the lowest organic material (charcoal) present in the site (Mulvaney 1975:184-188). It may be no coincidence that the excavated rockshelter where the longest and most abundant human occupation in W ardaman country has been discovered is also the closest to a large, permanent waterhole, and has such a wide overhang that it provides good shade in winter as well as summer. In other words, because of its convenient location close to good fishing and its large overhang, it may well have been used in winter as well as summer, unlike most other W ardaman rockshelters, which we were repeatedly told were used in the wet but not the dry season.

Caves as habitation sites

Aspect

Seasonality

Particularly favoured W ardaman sites provide deep shade, have a soft earth floor for comfortable sleeping, and are large enough to house the whole 'hearth group', the foraging unit which typically numbered 20-25 people. An excellent example of this is Gamawala I, a large rockshelter with wide overhang, facing south, where intensive occupation going back to 5500 BP was found (David et al. 1994; Flood & David 1994). On the opposite side of the same mesa-like outcrop is Garnawala II, a huge shelter but north facing, with many rock ledges but a comparatively small area of earth floor protected by the overhang. However, it must be remembered that in these tropical latitudes (ca 15° south) in high summer the angle of the sun is from the south rather than the north, so north-facing rockshelters provide better shade at that time of year. We excavated here because the immense number of stone artifacts lying on the surface in the dripline area (>2000 in 2.5m 2 ) showed considerable use of the shelter at least during the more recent past, and the presence of extremely weathered petroglyphs on the walls suggested that occupation might have some antiquity. The excavation (by Bryce Barker and Bruno David of the University of Queensland) is not yet complete, but occupation continues well below the assemblage of points, which typically appear in the west of the Northern Territory around 3000 years ago. Out of our six excavated rockshelters, it may be significant that it is north-facing Garnawala II which contains the largest surface assemblage of artifacts, the longest and probably oldest sequence of occupation and the largest body of petroglyphs and pigmented art. This shelter is far too shade-less and hot for comfortable daytime occupancy in winter, but seems to have been a favoured habitation in the summer wet season. Thus physical data correlate well with ethnographic evidence to indicate that this rockshelter was occupied in summer, not winter. Unfortunately bone and other organic material, which might shed light on seasonality questions, does not survive long in these sandstone rockshelters, probably mainly due to the numerous termites. Aspect of rockshelters is of most importance in temperate Australia, and in the cold regions it has been found that the majority of occupied rockshelters face east, to catch the early morning sun. Although the people of south-eastern Australia wore cloaks of kangaroo or possum skin, temperatures frequently drop below freezing in winter.

In the tropical north of Australia the primary use of caves was for habitation and shelter from monsoonal rain and the fierce sun in the summer wet season (from approximately November till April). Wardaman tribal territory receives an annual average rainfall of 45 inches (114cm), but usually concentrated into a few weeks of short, sharp tropical downpours, when caves provided the best means of keeping dry, although bark huts on stilts were used in other parts of the Northern Territory which lacked caves. Shade provided by rock overhangs was of equal or even greater importance in both the tropics and arid heart of Australia, where daytime temperatures can reach 50°C in summer and trees tend to be either absent or narrow-leaved eucalypts of savannah woodland type, which provide minimal shade. Because the rock retains the heat, rockshelters also provide some protection against the cold of the night, when temperatures often drop below freezing point. Proximity to water and food or other resources was not a major factor for Wardaman people in the selection of caves to use for summer wet season camping, since there was water everywhere at that time, and food such as yams and resources like ochre and stone were fairly evenly spread across the tribal territory. However, in the winter dry season occupation focused tightly on rivers and large waterholes, where both water and food such as fish, turtles, freshwater crocodiles, shellfish and water lily roots and other edible plants were available, and kangaroos and other game would come to drink in the evening. Winter camps were traditionally on the sandy banks or channels of seasonal creeks and rivers. If there happened to be caves located conveniently close to permanent water, these might be used in winter as well as summer. This may well be the case in Ingaladdi I (Aboriginal name: Yingalarri), a huge sandstone rockshelter with an earth floor sheltered by a wide overhang, located ca 1.5km from the excellent fishing of Ingaladdi waterhole, which is a permanent lagoon some eighteen metres wide and over 1.5km long. The site bears dozens of paintings and thousands of petroglyphs, including many extremely weathered ones in an early style. It was partially excavated by John Mulvaney in the 1960s, who found a very large assemblage of stone artifacts and some buried petroglyphs. The oldest occupation is significantly earlier than the radiocarbon date of

194

Australian Aboriginal Use of Caves

Spatial patterning within caves

about 20 W ardaman people were making a lunchtime camp to cook the food they had just caught or gathered, there would be two or three hearth groups. In the evening there would also be a separate young men's camp and hearth. Rarely has much intra-site spatial patterning been discovered in Australian cave excavations, because most are exploratory, test excavations of only a small area of a not very large cave. One outstanding exception was Rhys Jones' discovery in the quartzite south cave at Rocky Cape in Tasmania of a small inner chamber sealed off 6,800 years ago by midden accumulation at its mouth.

The aspect, size, rock art and other features of some 200 W ardaman rockshelters have now been recorded. Detailed analysis is not yet complete, but some relevant points have already emerged. More than 90% of rockshelters have stone artifacts on their earth floors and, in those which have been excavated, there are sub-surface stone tools, charcoal, ochre and other occupational debris. The numbers of surface artifacts vary greatly, but at least the larger assemblages can be taken to signify that the shelter was used for camping (those with very few artifacts may reflect use as occasional stopping places on a journey or 'dinnertime camps' whilst out foraging). Wardaman rockshelters which contain rock art but no occupational material generally have rocky floors unsuitable for sleeping. Although large pieces of soft bark were traditionally stripped off the paperbark trees and used as sleeping mats, they do not provide much padding, but are used more like a modem groundsheet. Moreover, paperbark trees grow only in moist areas usually along riverbanks, so would not be readily available in many areas. Thus caves with soft earth floors were much preferred over sleeping on rubble or rock slabs. Spatial patterning within caves has also been investigated, and what we have learnt from our own observations and our Aboriginal informants is that there was no rigid earmarking of particular parts of a rockshelter for particular activities. Unlike subterranean caves, the largest rockshelter is rarely more than 30m long and 10m wide. We learned that ceremonies were not carried out in the shelters, even those bearing large, elaborate paintings of Ancestral Beings, but out in the bush. This correlates with much other ethnographic and archaeological evidence that almost all the rockshelters were family camping shelters, rather than special purpose or 'secret-sacred' sites. The largest part of the earth floor would be used for sleeping, preferably with at least two yards cleared around the sleeping area, to discourage snakes, which occasionally frequent rockshelters, especially those close to water some of the world's deadliest snakes, like the taipan and king brown, live in tropical Australia; they hibernate in winter but would be much more in evidence in summer, and were an additional reason why rockshelters were preferred for summer camping over river banks, where snakes are numerous. Small fires would be lit around and even in between the sleepers for warmth (no skin cloaks or rugs were used in the tropics or arid centre as far as I know). This practice generates considerable ash. We observed it in an open-air Aboriginal camp near us, and when we recorded all features of the camp left behind after the Aboriginal people left, we discovered that the largest ash pile was by the sleeping place of the oldest man, who was on his own and felt the cold! Cooking might be done outside the rockshelter, or inside if protection from the elements was needed. Each 'hearth group' would have its own fireplace; a Wardaman hearth group tends to be the nuclear family, plus any elderly male or female relatives who do not have a spouse present, plus grandchildren without parents present. Typically, when

Piles of big abalone shells were dumped around the walls, but in the centre the floor had been swept clear for comfortable sitting round the fireplace. Here five small ashy hearths, placed close to the rock wall for maximum heat reflection, were found. Food refuse included bones of seals, fish, a few birds and small mammals, shells of rocky coast species, bracken fem stems, a lily tuber and split sections of the pith of the grass-tree. Stone scrapers and a stone mortar, with pestle neatly placed on top, had been left behind by the last occupants for their next visit, a visit that never took place (Flood 1995:196-197).

The surface of this inner chamber has been recorded, but the site has deliberately been left unexcavated, in view of the continual development of ever more sophisticated and non-destructive archaeological techniques.

Women's and men's sites Grinding hollows, ground surfaces and/or grooves resulting from the fashioning of stone axes or other artifacts are also common in rockshelters containing occupational debris. W ardaman women informed us that they, but not the men, used to grind up plant food, particularly the tubers of the water lily, a favourite, sweet food, in such hollows, which are regularly 7.5-lOcm in diameter. However, similar hollows were produced by grinding of ochre into powder, and this was done by both men and women. Ground surfaces (patches of more or less horizontal rock worn smooth by grinding) were said to be the result of women grinding seeds of grass or wattle or other plants into flour. Their presence in a rockshelter is therefore important evidence of the presence of women in a site. Other indicators of family camping places are stencils of babies' feet or children's hands. Stencils are rare in Wardaman country, but measuring the size of hand stencils and comparing with data on the size of Aborigines' hands recorded by physical anthropologists has proved useful elsewhere, as in a study of a large body of stencilled art in rockshelters in Koolburra in Cape York Peninsula (Flood 1987). Both archaeological evidence and ethnographic information indicate that the vast majority of W ardaman rockshelters were used by both men and women as family camping places. Men's presence is indicated by the occurrence of stone spear-points, adzes or ground-edge axes, which were traditionally used only by men (women's basic equipment was the wooden digging stick-cum-club and dilly-bag for collecting food, and her use of stone tools was largely confined to millstones used for grinding seeds). A

195

Flood

few remote shelters situated high above the valley floor on inaccessible clifflines were 'men only' sites, and usually have paintings of sacred Dreamtime Ancestral Beings and stone spear-points, but no evidence of use by women or children (see discussion below under Caves as ceremonial sites). Teenage boys are traditionally kept apart from the rest of the group when a 'band' is camping. In 1988 a group of about twenty Wardaman camped near us for ten days, and it was interesting to observe that the older boys camped some 30 metres away from the rest of the group of older men, women, girls and children. When visiting the Yingalarri I rockshelter, it was explained to us by senior traditional owners that the family group used to camp in the main, large rockshelter, but the young men were sent round the comer to a much smaller and less favourable shelter. Certain rockshelters were also 'women only' sites. In W ardaman country only one has been definitely identified, and this is Murdu-ya, a women's dancing place. Women traditionally danced naked, and this shelter provides privacy because its rock walls form a square enclosing an earth 'dancing ground', with a narrow entrance at both ends. There are a few, but very few, stone artifacts on the floor and a few paintings on the walls, including one of a woman at one end. Adjacent to it is a large, south-facing rockshelter, containing many spear-points and much, varied rock art, which was used as a family camping place.

Sacred and secular sites The vast majority of rockshelters in W ardaman country are secular, habitation sites. Sacred sites tend not to be rockshelters or caves, but features of unusual or dramatic form out in the open such as rocks resembling huge eggs, a phallus or turtle shell, a large rare tree or deep-cut waterhole. There are likely to be prohibitions on access to such sites, and they are often gender specific, such as a women's site where the visitor is guaranteed to become pregnant. Whilst spectacular features in rockshelters such as natural windows or archways will almost invariably be ascribed a sacred origin, and are said to have been made by a Creative Dreaming Being like an Ancestral Giant Kangaroo, who leapt through the rock, these features do not prevent access to the site for· family camping. For example, one deep rockshelter in the Gamawala outcrop has a dramatic chimney-like hole in the ceiling which was said to have been made by the Rainbow Serpent, but there are many stone artifacts on the floor, and no restrictions on access, at least currently.

Caves as art 'galleries'

range from gigantic figures of Dreaming Beings to European sailing ships (some painted, some engraved - i.e. both techniques were in use in modem times). Rock painting has ceased now, but retouching of Dreaming figures was done as recently as 1986 ('to make them look good' in preparation for the making of the film Land of the Lightning Brothers (Film Australia, Sydney). Rock art here is primarily a male activity, but we were continually told that 'anyone can do rock painting - men, women and children'. Rock art cannot readily be used to distinguish sacred from secular sites. The rock art of the W ardaman people is clearly used by those who know the associated oral traditions as a powerful method of communicating their religion, traditional belief system and cosmology to the next generation, but it is also used for education, telling a story, recording one's visit to a site (particularly by a hand stencil or print or an abraded groove) or simply decorating the 'living room' walls. The latter seems to be the primary reason for the mass of 'art' found in W ardaman rockshelters. Children's 'art' was pointed out to us in various rockshelters. This tended to be small rather crude drawings or paintings low on the back wall, the drawings usually in black charcoal and the paintings in white kaolin, the most widely available pigment. Subjects were mostly humans, birds or animals. Some Aboriginal children also rub straight line grooves on boulders or the back wall when visiting a rockshelter, for example at Yingalarri I in 1988 (Flood et al. 1992). Whilst the concepts of 'art for art's sake' and 'art galleries' seem quite foreign to Australian Aboriginal society, big caves offering expanses of relatively smooth walls almost always contain pigmented or engraved marks, ranging from linear finger-markings to complex figurative paintings of sacred or secular subjects (Bednarik 1986). Such sites also tend to exhibit many superimpositions of one picture on another. The great majority contain occupational debris, usually in close association with the 'art'. At some there may be prohibitions on touching the paintings or 'looking at them too long', in case harm should come to the onlooker. There is, nowadays, some embarrassment about sexually explicit scenes, and our senior male traditional owner said that he could visit and camp in one such site, but joked that he had to put his hands over his eyes in front of one particular scene of a man copulating with two women. Many rockshelters all over Australia bear signs of pounding, rubbing or scratching the wall with another rock, particularly along the rims of ledges or prominent natural features. The Wardaman say this is to bring out the 'power' in the rock.

Making marks or pictures on walls seems to be a primary human instinct, and availability of large smooth rock surfaces appears to have been an important factor in site selection. Virtually every accessible, Wardaman rockshelter of medium to large size with an earth floor bears both signs of occupation and some rock art. This comprises paintings (in wet pigment), drawings (in dry pigment), petroglyphs, stencils and prints. Subjects are both sacred and secular, and

Caves as ritual, ceremonial or mythological sites A few remote rockshelters, typically situated high above the valley floor at the base of an escarpment, were sacred, 'men only' sites, usually with paintings of sacred Dreamtime Ancestral Beings. These were used to instruct boys undergoing initiation and were also used for camping by the men and boys involved. They tend to contain prints and stencils

196

Australian Aboriginal Use of Caves

of men's or what seem to be adolescent boys' hands, but no evidence of the presence of women or children. Some relatively inaccessible rockshelters, such as those on a high cliffline, were used as 'boys' seclusion sites', where a few boys undergoing the various stages of initiation might be taken by a tribal elder for instruction into tribal religion and law. Remote rockshelters were traditionally used for this purpose, and small shelters with only a little occupational debris and a few small hand stencils on the walls may reflect this use. Elsewhere in the Northern Territory certain rockshelters were used only for ritual or ceremonial purposes, for example, Sleisbeck Cave in the extreme south of Kakadu National Park, which contained a sacred arrangement of crocodile bones (Mulvaney 1975:plate 78), or the sacred men's initiation shelter in the base of Uluru (Ayers Rock) in central Australia (Mountford 1977). And in some subterranean limestone caves on the Nullarbor Plain in South Australia stone arrangements, traditionally used in ceremonies, have been built in the outer chambers. Certain caves of dramatic form are sacred mythological sites, which exhibit no visible traces of past Aboriginal presence or modification, but are of great significance in Aboriginal religion, as made by an Ancestral Being. For example, on Beecroft Peninsula south of Sydney Devil's Hole, an 80-metre-deep abyss connected with the sea through a series of caves, is believed to house the spirit of a Creator Ancestor, Bundoola (Flood 1990:287). In Wardaman country the site of Garnyiwarnyirr, a high, steep rock outcrop, is associated with a Rainbow Serpent, who went into a cave on the south face and blasted its way to the top, making a massive 'crater' inside the outcrop. Such dramatic features in the landscape will often be a focus for Aboriginal visits, and hence occupational debris, which is found along with rock paintings and petroglyphs and human burials in rockshelters around the base of Garnyiwarnyirr.

prehistoric period. The presence of visible human skeletal material in a rockshelter did not inhibit its use for camping; it was treated surprisingly casually, whether the remains were of an unknown individual or 'Aunt Annie'.

Caves as sources of mineral resources Some caves were important sources of mineral resources, such as stone for tool-making. This particularly applies to subterranean limestone caves, such as Koonalda Cave on the Nullarbor Plain in South Australia, where people camped and mined flint more than 24,000 years ago (Wright 1971). Koonalda had another vital resource in an arid area drinking water in an underground lake. The prehistoric inhabitants also made linear engravings on the soft limestone walls in chambers and passages far beyond the reach of daylight, and there is strong circumstantial evidence that these are at least 20,000 years old (Flood 1995:152-154).

Antiquity and caves in Australia

The earliest sites The earliest human occupation yet found in Australia is in rockshelters at the foot of the western Arnhem Land escarpment in the Northern Territory (Fig. 1). A series of thermoluminescence (TL) dates on artifact-bearing quartz sands in three rockshelter deposits suggest the arrival of people in northern Australia between 50,000 and 60,000 years ago (Roberts et al. 1990, 1993). The TL dates were obtained from sandy footslope deposits at occupation sites, which contain stone artifacts in their primary depositional setting. Comparative radiocarbon and luminescence dates obtained from the upper occupation layers of all three rockshelters, Malangangerr, Malakunanja II and Nauwalabila I, show close correspondence. The large rockshelter of Malangangerr, near the East Alligator River in Kakadu National Park, yielded a sequence going back to 32,000±7900 BP (KTL-126), the lowest artifacts being at a depth of two metres. Malakunanja II, a rockshelter with a shallow overhang at the foot of the western escarpment of the Arnhem Land plateau, gave a basal TL determination of ca 100,000 years. The layers containing the oldest human occupation were TL dated to between 52,000 and 61,000 BP. Human occupation appears to have commenced abruptly at a depth of 2.6m below present ground surface, dated to 61,000±9130 BP (KTL-162). There was dense occupation from 2.3-2.Sm depth, between 45,000±6900 BP (KTL-164) and 52,000±7110 BP (KTL-158), with more than 1500 artifacts in this occupation layer. Cultural material included corescrapers, flakes and amorphous artifacts made of silcrete, quartzite and white quartz, a grindstone, pieces of dolerite and ground haematite, chlorite and mica, together with red and yellow ochre. Whilst allowing for the possibility that artifacts were trodden into soft, sandy, older sediments by the first occupants, the researchers state that a conservative estimate places initial occupation at ca 52,000 years (Roberts et al. 1990). Below this, a sterile sand sheet,

Caves as keeping places Caves were also used as hiding places, where sacred items used in rituals were cached for safe keeping. They might be sacred churinga (engraved boards) made of wood or stone, pearl shell pendants, circumcision knives, quartz crystals or other sacred items, often brought from many hundreds of miles away for use in secret ceremonies. In a rockshelter they would generally be wrapped in paperbark and hidden on a high ledge or cached under a boulder.

Caves as burial sites The W ardaman dispose of their dead by secondary burial, exposing the corpse on a platform, later returning to collect the bones and then placing them on a ledge under a rock overhang in their own clan estate. Many skulls still lie on ledges in rockshelters or niches in cliffs 'looking out over their own country'. Long bones are often found with them, wrapped in paperbark. Human bones are not buried in the floors of W ardaman shelters, nor have any been found in excavations, but this practice has occurred in some other Northern Territory rockshelters, both recently and in the

197

Flood

Matenkupkum Matenbek'

NEW GUI

-=

j/J

0

Nawamoy •• Malangang Malakunanja U Nauwalabila

r

Widglngarri

.

•Miriwun

alkunder A~ch oColless Creek •Cuckadoo

AUSTRALIA • Puritjarra • Mt Newman

Kenniff Cave. Native Well•

• Puntutjarpa

Koon aid a• Allens Cave•

Hawker Lagoon



•Upper Swan



Lime Springs •

Willandra Lakes



Cheetup Kal2an Hall

PLEISTOCENE

Menindee Lakes

SAHUL

"=jJ

Cave Bay Cave

>:Mannalargenna

Land Exposed 18,000 years ago. (-200m sea level). 0

Figure 1

100

200

Mackintosv• Kutikina

300 Mllos

OP.P.n,1RP.An~

-TASMANIA ORS 7

• ::~.

...

Beginner's

Luck

Major archaeological sites of the Late Pleistocene in Australasia.

immigrants. In the Kakadu region and elsewhere they made extensive use of fire, presumably in the course of the food quest, for geomorphological and palynological evidence reveal vegetation modification and ecosystem alteration on a continental scale commencing around 55,000-60,000 years ago. It also now appears (from amino acid dates on extinct Genyornis eggshell from Lakes Mungo, Arumpo and Victoria, and other evidence) that megafaunal extinction occurred essentially simultaneously with human immigration, making likely a causal relationship (Miller 1992). Some form of art was also practised in caves by these early settlers. Ground and striated haematite was found at the base of the cultural sequence in Malakunanja II (Roberts et al. 1990), indicating that paint was being made more than 50,000 years ago, although we do not know whether the

completely devoid of any artifacts, was deposited steadily from around 110,000 years ago (Jones 1992). Preliminary dating by optically stimulated luminescence (OSL) dating - a more sophisticated version of TL - of the lowest levels at Nauwalabila I or the Lindner site, a rockshelter on Deaf Adder Creek in Kakadu National Park, has revealed human occupation at 56,000±11,000 BP, with artifacts occurring even below this date (Roberts et al. 1993). Again optical dating is consistent with radiocarbon dating and with depth. The coincidence of first occupation at Nauwalabila I and Malakunanja II rockshelters and the absence of artifacts in deposits older than 60,000 years in the prime site of Malakunanja II strongly suggest that humans first arrived in this region 50,000-60,000 years ago. It seems then that rockshelters were used for camping, and stone tool manufacturing, by Australia's first

198

Australian Aboriginal Use of Caves

pigment was used to decorate the cave walls or for other purposes such as body decoration. The walls of this rockshelter, along with thousands of others, are now decorated with paintings, but their age is unknown. At several rockshelters organic material apparently contained within pigment has yielded AMS radiocarbon dates of Pleistocene antiquity, such as Laurie Creek in the Northern Territory, where red ochre was dated to ca 18,000 BP (C. Chippindale, pers. comm.), and Sandy Creek rockshelter in Cape York Peninsula which gave an AMS date of ca 24,000 BP (Ang-Gnarra Aboriginal Corporation, pers. comm.). It is not absolutely clear whether these dates relate to paintings or merely to organic deposits on the walls, but petroglyphs in the Olary region of South Australia have yielded dates of ca 40,000 years, by a combination of AMS radiocarbon dating of microscopic organic remains found in the rock varnish covering the engravings, cation-ratio dating of the varnish, and analysis of the varnish stratigraphy (Nobbs & Dom 1993). The dated petroglyphs (Karolta and Panaramitee) are in the open air rather than in caves, which are exceedingly scarce in the Olary region, but the motifs of circles, lines, bird and kangaroo tracks, and many dots and small pits, also occur widely in Australian caves and may well be of similar age. The impression that the oldest art in Australia was engraved rather than pigmented may be entirely due to the lesser durability of pigmented art. The presence of used ochre in the floor deposits of a number of Pleistocene occupational caves like Malakunanja II suggests that painting was carried out just as early as engraving - there are virtually no petroglyphs in the northern part of Kakadu National Park, where Malakunanja, Malangangerr and Nauwalabila lie, probably because of the difficulty of engraving quartzite, a much harder rock than sandstone. The art of stencilling is also probably of great antiquity, for hand stencils have been found in Pleistocene contexts deep inside Tasmanian underground limestone caves, such as W argata Mina Cave in the Southern Forests, where subsequent growth of stalactites has almost covered them (Cosgrove & Jones 1989; Loy et al. 1990). Currently some Australian caves bear only petroglyphs, some only paintings and/or drawings and/or stencils, some a combination, whilst others were apparently occupied but undecorated. The present distribution of rock art techniques and styles depends on many cultural and natural factors, such as relative hardness of the rock and consequent ease or difficulty of engraving. Projection of the current pattern back into the past is extremely difficult, and complicated by questions of the relative durability of pigmented and nonpigmented art.

from the W ardaman people, who lived in caves during the wet season only 50 years ago, shows that their primary use was as wet season habitation sites, but they were also used for burials, and some were men's or women's sacred sites, used in the course of rituals, ceremonies or single-sex activities. In arid environments the importance of site morphology and location and the proximity of water and other resources as factors in human use of caves for habitation and/or ritual purposes, has been demonstrated by recent excavations in several rockshelters, for example in central Australia at Puritjarra (Smith 1987, 1989) and in north Queensland at Fern Cave and Nurrabullgin (David 1991, 1993). In temperate regions in south-eastern Australia much new archaeological research has elucidated the rationale for the choice and use of particular rockshelters as habitation sites, and aspect has proved a particularly important factor in site selection. Finally, in Tasmania occupation has now been found extending back throughout the last glacial maximum to 35,000 BP, and hand stencils are present in dark chambers deep inside some of these subterranean limestone caves, suggesting ritual as well as habitational use. Acknowledgements: The information on which this article is based comes from W ardaman traditional owners and custodians, to whom we owe a great debt of gratitude for allowing us to look into their culture and publish the results. The main people who participated in the project were the late Ruby Alison, Riley Burdun, July and June Blutcher, Daisy Girnin, Lily Gin.gina, Billie Harney, Queenie Ngabijiji, Tarpot Ngamunagami, Elsie, Oliver, Barbara, Michael, Lindsay, Tilley and Jason Raymond. We would also like to extend our thanks to the owners and managers of Innesvale, Willeroo and Delamere. Earthwatch provided the major funding for five field seasons, and Earthcorps volunteers carried out high quality fieldwork in difficult conditions. Dr Francesca Merlan of Sydney University, the Australian Heritage Commission, Australian Institute of Aboriginal and Torres Strait Islander Studies, and the North Australia Research Unit of the Australian National University provided a variety of help, for which we are most grateful. Finally, my heartfelt thanks go to my fellow researchers, in particular to Bruno David and to my husband Nigel Peacock.

References BEDNARlK, R.G. 1986. Cave use by Australian Pleistocene man. Proceedings of the University of Bristol Speleological Society 17:227-245. COSGROVE, R. & JONES, R. 1989. A subterranean Aboriginal painting site, southern Tasmania. Rock Art Research 6(2):96-104. DAVID, B. 1991. Fern Cave, rock art and social formations: rock art regionalisation and demographic models in southeastern Cape York Peninsula. Archaeology in Oceania 26:41-57. DAVID, B. 1993. Nurrabullgin Cave: preliminary results from a pre-37,000 year old rockshelter, North Queensland. Archaeology in Oceania 28(1):50-54. DAVID, B., McNIVEN, I., ATTENBROW, V., FLOOD, J. & COLLINS, J. 1994. Of Lightning Brothers and white cockatoos: dating the antiquity of signifying systems in the Northern Territory, Australia. Antiquity 68:241-251. FLOOD, J. 1987. Rock art of the Koolburra Plateau, north Queensland. Rock Art Research 4(2):91-126. FLOOD, J. 1990. The Riches of Ancient Australia. A Journey into Prehistory. University of Queensland Press, Brisbane.

Conclusions A continuous tradition of habitation in caves by Australian foragers extends from more than 50,000 years ago till today. As habitation sites, caves were important in hot arid central Australia for shade, in temperate regions such as south-eastern Australia for shelter against rain and cold, and in the tropical north for both shade and shelter in the summer wet season. In the Northern Territory evidence

199

Flood FLOOD, J. 1995. The Archaeology of the Dreamtime, 3rd ed. Angus and Robertson, an imprint of Harper Collins, Sydney, London & New York. FLOOD, J. & DAVID, B. 1994. Traditional systems of encoding meaning in Wardaman rock art, Northern Territory, Australia. The Artefact 17:6-22. FLOOD, J., DAVID, B. & FROST, R. 1992. Dreaming into art: Aboriginal interpretations of rock engravings, Yingalarri, Northern Territory (Australia). In Rock art and Ethnography, edited by M.J. Morwood & D.R. Hobbs, pp. 33-38. Occasional AURA Publication no. 5. Australian Rock Art Research Association, Melbourne. LOY, T.H., JONES, R., NELSON, D.E., MEEHAN, B. & VOGEL, J. 1990. Accelerator radiocarbon dating of human blood proteins in pigments from late Pleistocene art sites in Australia. Antiquity 64:110-116. MILLER, G. 1992. Human immigration and megafauna extinction in Australia dated by amino acid racemization in ratite eggshell and palaeoclimatic implications. Paper presented at 'Australia Day. A Meeting of the Old and the New in Australian Archaeology', Cambridge University Museum of Archaeology and Anthropology, Cambridge. MOUNTFORD, C.P. 1977.Ayers Rock. Rigby, Melbourne.

MULVANEY, D.J. 1975. The Prehistory of Australia. Penguin, Melbourne. NOBBS, M. & DORN, R. 1993. New surface exposure ages for petroglyphs from the Olary Province, South Australia. Archaeology in Oceania 28:18-39. RENFREW, C. & BAHN, P. 1991. Archaeology. Theories, Methods and Practice. Thames & Hudson, London. ROBERTS, R.G., JONES, R. & SMITH, M.A. 1990. Thermoluminescence dating of a 50,000-year-old human occupation site in northern Australia. Nature 345: 153-156. ROBERTS, R.G., JONES, R. & SMITH, M.A. 1993. Optical dating at Deaf Adder Gorge, Northern Territory, indicates human occupation between 53,000 and 60,000 years ago. Australian Archaeology 37:58-59. SMITH, M.A. 1987. Pleistocene occupation in arid central Australia. Nature 328:710-711. SMITH, M.A. 1989. The case for a resident human population in the Central Australian Ranges during full glacial aridity. Archaeology in Oceania 24:93-105. WRIGHT, R.V.S. (ed.) 1971. Archaeology of the Gallus Site, Koonalda Cave. Australian Institute of Aboriginal Studies, Canberra.

200

Precolumbian Cave Utilization in the Maya Area Andrea Stone

Abstract The limestone-rich Maya area is endowed with an abundance of caves. The human use of caves has continued unabated since before the present millennium until the present day. The Maya exploited water and mineral resources in caves and also used them extensively as religious shrines. Archaeological evidence of ritual cave use, mainly dating from the Classic period (AD 200-900), includes modification of cave walls with architectural features, pottery and other portable artifacts, human skeletal remains, and painted and carved wall art. The cultural context of this material rests on a cognitive model of space in which caves were imbued with supernatural power. Ancient forms of Maya ritual activity in caves can be reconstructed from the available archaeological and ethnological data.

The Maya area, comprising Belize, Guatemala, southern Mexico, and western Honduras and El Salvador, is a remarkable repository of archaeologically important caves. A full 70% of the terrain consists of karstified limestone, and so caves abound in virtually all comers of this tropical region (Brady & Veni 1992). The ancient Maya paid a great deal of attention to caves, explored them assiduously, and viewed them as valuable cultural and material resources. Hence, the idea that caves are marginal to the social and economic processes of complex societies is not true in the case of the Prehispanic Maya. It is important to point out that Maya cave utilization extends to caves of immense proportions, and not just to shallow caves. The largest caves utilized by the ancient Maya, which also happen to be the largest caves in all of Central America, are located in a part of Belize known as the Chiquibul, named after a river that drains the Maya Mountains. In terms of volume, the Chiquibul caves rank as one of the largest cave systems in the world (T. Miller, pers. comm. 1986). Four caves separated by collapsed roofs, Actun Kabal, Actun Tunkul, Cebada and Xibalba, reach lengths of ten to fifteen kilometers each, with a total length of fifty-five kilometres (Miller 1996: fig. 15). The passages frequently exceed one hundred metres in width; they have an average width of 30-40 metres (Miller 1984). The Belize Chamber of Actun Tunkul measures an astounding 150 by 350 metres. The Caves Branch River Valley in central Belize is also home to some of the biggest caves known in the Maya area, such as Mountain Cow, Petroglyph (ReentsBudet & MacLeod 1986), and Footprint Cave (Graham et al. 1980). Large caves are also found in the department of Alta Verapaz in Guatemala, notably the Candelaria cave system which has thirty kilometers of large passage in a series of caves (Gatica 1980; Carot 1984). The longest Yucatecan cave, Actun Kaua, has seven kilometers of surveyed passage (Reddell 1977:260). The deepest point of penetration into any of these caves by the ancient Maya stands at about three kilometers (at Footprint Cave, Belize, at ca AD 800; MacLeod & Puleston 1980:72). The limiting factor in the Maya's quest to explore

deep caves was illumination, typically provided by wood torches, though in Yucatan modem Maya make torches from dried cactus and grass stalks (Mercer 1896:fig. 35), a practice which may have Precolumbian antecedents. Rarely the remains of torches have been found; for instance, un unburnt wood torch in extremely fragile condition was discovered at Footprint Cave (A. Stone, unpublished data). Late Classic (ca AD 700) ceramic torch holders (ceramic tubes closed at one end and flared at the other) have been found in several Belizean and Guatemalan caves (Graham et al. 1980:169; Brady 1989:257-258). Maya caves commonly have charcoal marks on the walls and flecks of charcoal on the ground. Lithics found in association with Pleistocene megafauna at Loltun, Yucatan, indicate that human utilization of caves in the Maya area extends back to the late Pleistocene, around 8000 BC (Velazquez Valadez 1980). However, the late Pleistocene antedates occupation by ethnic Maya as defined linguistically, as Proto-Maya is usually placed at ca 2400 BC (Kaufman 1976:107). The earliest ceramics known from any Maya cave are Middle Preclassic (ca 900-400 BC) and have mainly been reported from the Yucatan Peninsula (e.g. Gonzalez Licon 1986:70; Zapata Peraza et al. 1991:26). Their presence constitutes the earliest evidence of cave use by the ancient Maya which later reached a zenith during the Late to Terminal Classic period (AD 600-900). Most of the Belizean caves excavated by Pendergast (1969, 1970, 1971, 1974), for instance, yielded only Late to Terminal Classic material, and the majority of caves studied thus far have a Late or Terminal Classic component, even when other occupational periods are represented. The longest cultural sequence for any cave used by the Prehispanic Maya is at Loltun which yielded ceramics spanning the Middle Preclassic to Late Postclassic periods (ca 800 BC-AD 1500).

Practical exploitation of caves by the Maya The Maya made some practical uses of caves, though habitation was not among them, and certainly the tropical

201

Stone weather never warranted using caves as permanent shelters. Historical records indicate, though, that in extraordinary circumstances the Maya took temporary refuge in caves, as they did in Loltun during the nineteenth century Caste War (Thompson, J.E.S. 1975:xli). In 1934, Actun Spukil, Yucatan, served as a hiding place for outlaws (Hatt et al. 1953:21). Yucatecan farmers working in fields distant from their homes are reported to live in caves for brief periods (Hatt et al. 1953:15). Certain masonry walls in cave entrances have been identified as hunters' blinds (Mercer 1896; Hatt et al. 1953:15-19). Again, all of these practices may have Precolumbian antecedents. Some rather sketchy evidence indicates that the Maya extracted clays and minerals from caves. For instance, modern Yucatec Maya obtain pottery temper from caves in the form of crystalline calcite (presumably from speleothems) as well as other tempering material (Thompson, R.H. 1958:70). Elsewhere in Yucatan gypsum was extracted from caves, as at Actun Jih (Hatt et al. 1953:16), and a red earth was mined from Actun Spukil (Hatt et al. 1953:23). Attapulgite was also retrieved from caves. This white mineral, an ingredient of Maya Blue, was mined in great quantities from a deep, cave-like cenote (a type of sinkhole) in Yucatan (Arnold & Bohor 1978). While such mining activity can be difficult to date archaeologically, certain examples are definitely Prehispanic. The attapulgite mining seems to be that. Another case of Prehispanic mining was discovered at Footprint Cave, Belize, where a stone mortar with a white mineral ground on the surface was found along with other Late to Terminal Classic artifacts (MacLeod & Puleston 1980:72). In addition, many Prehispanic Maya cave paintings utilize pigments that appear to derive from clays found on the nearby cave floor. But, as Brady (1989:33) observes, the extent to which the Maya engaged in mining activities in caves is difficult to assess, as this practice has not been adequately investigated. The use of bat guano by the Maya has not been reported to date; however, the Aztecs hauled bat guano in baskets on the backs of slaves from caves as far away as Morelos, which they used to fertilize their chinampas (Werner 1992: 10). The Aztecs also put bat guano to use in another curious way. The sixteenth century Florentine Codex, a treatise on Aztec culture compiled by Bernardino de Sahagun (1963:233), mentions the gluing of jade mosaics with bat excrement, an odd use, indeed, for bat guano extracted from caves. Without doubt, the most important resource caves provided to the ancient Maya was water, especially in the northern Yucatan Peninsula (including the Mexican states of Campeche, Yucatan, and Quintana Roo) where surface water features are absent owing to the highly fractured nature of the limestone. Here caves (called actun and ch'en in Yucatec) and cenotes (Yucatec ts'onot) have long provided the Maya with a source of water during the dry season. Thompson asserts that the Maya's persistent interest in caves was driven by their need for water; however, this could only be argued for the northern Yucatan Peninsula where access to water from caves and cenotes was so crucial that it even determined settlement patterns

202

(Thompson, J.E.S. 1975:x). Moreover, cave water was often extracted at considerable effort. For instance, at the Gruta de Chae, south of the Puuc site of Kabah, 'water had to be carried on hands and knees for almost a kilometer and then 65m up an almost vertical series of ladders to the entrance' (Andrews, E.W. 1965:7). The tortuous passage to the pool of water is littered with thousands of broken water jars, many dating back to the Early Classic (ca AD 300-600). The mid-nineteenth century explorer John Lloyd Stephens witnessed an extraordinary scene of water collection at Xtacumbilxunaan near Bolonchen, Campeche. To reach the water table, the Maya had to negotiate a 70-metre descent, the final stretch being made on an immense wooden ladder (Catherwood 1844:pl. 20).

Caves and Maya religion While the Maya's practical exploitation of caves was not inconsequential - especially in the northern Yucatan Peninsula in light of their acute water problem - their importance from a purely cultural standpoint far outweighs any material benefits that may have been derived from them. In fact, the religious and ritual role of caves among the Maya motivated nearly the entire spectrum of Prehispanic utilization. This can be determined from a substantial body of evidence: an extensive archaeological record supplemented by a rich ethnological record. The Maya were using caves at the time of Spanish contact and continue to do so to this day. Cultural continuity in this case is quite strong, and so ethnographic analogy represents a valid avenue of interpretation. Ancient Maya art and writing also provide a window into ancient ritual practices and religious beliefs associated with caves. What can be gleaned from these sources is that, essentially, the role of caves in Maya culture is a function of religious beliefs and ritual practices generated by sociospatial concepts. Caves constituted a form of sacred space in Maya culture, just as they did for most ancient Mesoamericans (see Heyden 1975, 1976, 1981 for a discussion of the cave cult in non-Maya parts of Mesoamerica). The cave's sanctity hinged on a cognitive model of space in which rugged, remote features of the natural landscape, a category of social space that might be termed 'wilderness', were construed as an inversion of mundane space, and, as such, assumed sacred qualities (Stone 1992, 1995). Places in the wilderness were considered ideal for petitioning deities who were believed to reside in natural landforms, not just caves, but also mountain crests, springs, lakes, and other dramatic features of the landscape. Such wilderness locales, meeting grounds with the gods, evolved into topographic shrines, many of which, perhaps the majority, consisted of caves. In Maya thought caves were a conduit into the bowels of the earth, a dangerous but supernaturally charged realm, often referred to as the 'underworld' in current literature or by the Quiche term, Xibalba. Herein dwelt the ancestors, rain gods, various 'owners' of the earth, culture heroes, nefarious death demons, animal and wind spirits. The Maya made pilgrimages to caves to propitiate these beings; sometimes this entailed long journeys of over a hundred

Precolumbian Cave Utilization in the Maya Area kilometers (this is known both archaeologically and ethnographically). Post-contact sources tell us that cave ceremonies usually concerned rain and other agricultural interests, hunting, ancestor worship, renewal/New Year rites and other calendrically-timed ceremonies, and petitions for various personal needs (e.g. health problems). Caves were also used by brujos (witches) to cast spells.

entrance (Andrews, E.W. 1970:fig. 2). These constnctlons and blockages reverberate with the notion of the wilderness as a remote and inaccessible place. It would seem that passages difficult of access, blocked by walls and so forth, were perceived as having heightened sacred qualities. The narrowing of cave passages also underscores the fact that sacred space for the Maya was, in its most fundamental form, enclosed, intimate, and remote. This is also apparent in Maya temple architecture with its windowless, small rooms elevated on towering platforms. The narrowing of rooms and passages in caves also echoes the private nature of cave ritual which often entailed highly personal rites of passage, including ritual bloodletting. Altar-like constructions have also been found in Maya caves. In the entrance of the Chiquibul cave, Actun Kabal, is a low pile of stones topped by two stone 'rabbit ears' (A. Stone, unpublished data); this structure probably served as an altar. One unequivocal altar was found in a hidden passage at Naj Tunich (Stone 1989a): a pile of limestone rocks topped by a long, tapering stone supporting two ollas rims. Small fires had been lit on the altar, and on the nearby floor lay a smashed polychrome plate; a hieroglyphic text is painted on the back wall. Another curious altar-like construction found in a deep, hidden chamber at Naj Tunich consists of a box-like form, about a metre across, made from broken speleothems with six broken stalactites set on top (Brady et al. 1992:fig. 3). Stalactites and other speleothems were sometimes removed from their original setting to be used, seemingly, as objects of veneration. For instance, a stalactite was set up stela-like in the entrance of Petroglyph Cave, Belize (Reents-Budet & MacLeod 1986). Cave formations were used in similar ways at surface sites. One striking example is a stalactite carved with the portrait of the ruler Bird Jaguar IV which stands in front of his temple, Structure 33, at Y axchilan, Mexico. Stalactites were also set up among a number of other carved stelae in front of Structure 41, Yaxchilan, a building commissioned by Shield Jaguar, Bird Jaguar IV's father (Tate 1992: 240).

Maya cave archaeology The incredibly rich and varied material remains found in Maya caves rest upon this cultural foundation. For the remainder of this paper I will briefly survey the nature of the archaeological record, though I cannot, for lack of time and space, cover this topic in anything approaching a comprehensive fashion. I would direct the reader to more detailed studies for this and highly recommend two articles by J. Eric S. Thompson (1959, 1975), the first, and still among the most successful, attempts by any Mayanist to define the role of caves in Maya culture. More recently, Brady (1989) has addressed Maya cave archaeology systematically based on an exhaustive search of the literature, and McNatt (1996) has produced a major synthesis of Belizean cave archaeology, incorporating new data based on unpublished reports. Also see Bonor Villarejo (1989), Bassie-Sweet (1991) and Stone 1995).

Stone constructions Stone masonry constructions are commonly found in Maya caves, usually in the entrance but also in recessed areas. Cave entrances often contain low walls. Their function is unknown, but they seem to demarcate precincts; some have been identified as hunters' blinds, as noted above. Walls were used to modify natural topography, especially to rectify incoherent jumbles of stones. For instance, a massive breakdown slope at Naj Tunich, Guatemala, was transformed into rough terraces by the addition of masonry walls. Some stone constructions found in Maya caves are quite elaborate. They are sometimes associated with plaster floors, though plaster floors also occur in Maya caves independent of ot;her architectural features (e.g. Las Cuevas, see Anderson 1962:327). Structure 2 in the entrance of Naj Tunich is a rectangular enclosure constructed from ashlar blocks and rubble core. The walls stood over two metres high and probably supported a perishable roof. This building seems to have served during the Late Classic period as a tomb (Brady 1989:132-137). The cave of Quen Santo in highland Guatemala houses a unique plastered masonry building with a post and lintel portal resting on a platform, reminiscent of a Maya temple (Seler 1901:fig. 239). Masonry walls were also used to block passages partially or completely. Constricted passages can be seen, for example, at Las Cuevas (Anderson 1962), Eduardo Quiroz Cave (Pendergast 1971), and Quen Santo; in the latter case walls narrow the passage leading to the temple-like building at the back (Seler 1901:fig. 236 & 238). The cave of Balankanche, located near Chichen ltza, was completely sealed by a masonry wall located some 250m in from the

Ceramics Ceramics comprise the most prolific artifact category found in Maya caves. They are sometimes recovered from obscure nooks and crannies and are frequently mutilated with 'kill holes' or battered rims or leg supports. Potsherd middens are also found in Maya caves. This at times is the result of accidental breakage, as at the Gruta de Chae where sherds from thousands of jars represent vessels broken in the Maya's quest for water. However, at other times piles of broken pottery have a purely ritual function. Thompson describes pottery middens from caves in Belize (Thompson, J.E.S. 1975:xviii-xix), and similar pottery dumps have been found in Naj Tunich (Brady 1989:83) and Miramar (Breuil 1986). The smashing of pottery in caves may have been a type of sacrificial offering, just as some highland Maya today smash pottery at ancestor shrines on mountain tops (Tedlock 1982:fig. 13). The repetition of such acts over successive generations could have generated cave pottery dumps.

203

Stone Pottery found in Maya caves is diverse and includes crude domestic wares as well as fine polychromes, the most famous of the latter being the Hokeb Ha Vase (Palacio 1977) and the Actun Balam Vase (Pendergast 1969), both found in Belizean caves. While cave pottery usually reflects regional surface site collections, nonetheless, certain kinds of pottery have a high incidence in cave artifact inventories. Brady (1989:238) has noted, for instance, that for the Maya area three-quarters of the provenanced 'shoe-pots', a roughly shoe-shaped vessel found across the Americas, have been found in caves. He therefore concludes that shoepots had a special use associated with cave ritual, at least among the Maya. Certainly, one of the most common ceramic forms found in Maya caves is the olla or water storage vessel. While at times water collection was motivated by pragmatic needs, at other times it was collected strictly for ritual use. Thompson first proposed that the Maya sought suhuy ha or 'virgin water' from caves (Thompson, J.E.S. 1959), an idea which has been borne out archaeologically. Ollas were set under active drips in areas of such difficult access that they could not represent the exploitation of water resources for merely practical purposes. Virgin water was also collected in Yucatan in a stone trough called a haltun. These crudely cut blocks of stone with a shallow depression are frequently encountered in Yucatecan caves under ceiling drips. Colonial sources describe the use of 'virgin water' for rites of purification, and it is believed that this water was often collected from caves (Thompson, J.E.S. 1975:xxi). Ceramics also served as incense burners. The burning of incense was and remains a standard fixture of Maya cave ritual (as it is for virtually all forms of Maya ritual). In Rio Frio Cave C stacked bowls with hardened copal were found (Reents-Budet & MacLeod 1986). Sometimes copal was burnt in a special, decorated incense burner; perhaps the most famous examples of this from any Maya cave are the Terminal Classic incense burners adorned with the visage of the Mexican god Tlaloc from Balankanche which contained ashes, presumably from burnt incense (Andrews, E.W. 1970). However, incense was also burnt in plain domestic pottery. Brady (1989:212) found, for instance, that utilitarian pottery sherds from Naj Tunich, mainly low bowls, were imP,regnated with copal smoke, and one of the paintings from the cave shows a figure sitting in front of a bowl of smoking incense (Stone 1989b: fig. 22-18).

Special artifacts Artifact collections from Maya caves often include manos and metates (maize-grinding implements). A particularly interesting example of this occurs at Balankanche where 232 miniature manos and metates were discovered (Andrews, E.W. 1970:11). In Actun Kabal, Belize, over twenty full-sized metates were scattered in a remote upper level passage, hauled up at considerable effort (A. Stone, unpublished data). These maize-processing implements appear to be offerings tied to agricultural rites held in caves. Caves also commonly contain spindle whorls, while obsidian blades and bone needles were probably used for auto-sacrifice.

204

Human skeletal remains Thompson's inventory of major categories of Maya cave use includes 'burials, ossuaries, and cremations' (Thompson, J.E.S. 1975:xiv). Human remains are frequently encountered in Maya caves, though the bones are often disturbed, making it difficult to assess the original context of the deposits. Some caves seem to have functioned as ossuaries wherein bones were collected and sometimes separated into piles of skulls, long bones, etc. At the Gruta de Xcan, Yucatan, 11 skulls were found cached together, while long bones were found in another pile (Marquez de Gonzalez et al. 1982). Brady (1989:344) notes that the charred bones from Cave 3 of Copan, filling a 25m x 6m chamber to a depth of nearly a metre (also see Gordon 1898:143-144), is the largest deposit of human bones found in a Maya cave. Cremated human remains are reported by Blom (1954) for a group of caves in Chiapas. Ethnohistoric sources suggest that ossuaries functioned as ancestor shrines. Nunez de la Vega, describing the customs of eighteenth century Maya of highland Chiapas, mentions an ancestor cult centered on a collection of bones kept in a cave (Thompson, J.E.S. 1975:xxxiii). Burials also comprise some of the human remains found in Maya caves. Structures in the entrance of N aj Tunich, which contained fragments of human bone, have been identified by Brady (1989) as tombs. Though the tombs were looted, jade and other luxury goods were found in the debris. Some human skeletons are the remains of sacrificial victims, and this seems especially likely for the skeletons of children. For instance, in Petroglyph Cave, Belize, the skeletons of six infants between the ages of six months and one-and-a-half years were found in such a way that it was clear that they had been abandoned in a pool of water (Reents-Budet & MacLeod 1986). Brady (1989:348-363) notes that the seven children's skeletons found at Naj Tunich probably were sacrificed (one had an unhealed wound in the cranium). Based on an analogy with Aztec practices, the sacrifice of children in Maya caves was probably tied to a rain cult.

Artificial caves The importance of caves in Maya culture can also be gauged by the remarkable practice of constructing manmade caves (Brady & Veni 1992). This has been documented in highland Guatemala at four sites where manmade caves range between 8m and 68m in length. In two cases, at La Laguna and Utatlan, the caves were excavated out of volcanic ash deposits. At Mixco Viejo, the cave is an enlargement of an existing natural fissure, while the cave of Esquipulas is completely man-made. The Utatlan and Esquipulas caves are today highly popular pilgrimage shrines.

Cave art The Maya decorated caves with painted, drawn sculpted, and imprinted motifs. Though considerably smaller than the Palaeolithic corpus, Maya cave art is at times quite

Precolumbian Cave Utilization in the Maya Area spectacular. Sculpted images were created by incising, basrelief, and modelling. To date only two fully modelled figures have been reported in the literature. A grotesque face was modelled in clay in a wall of Footprint Cave (Graham et al. 1980:155; McNatt 1996:fig. 7), and an extraordinary life-size figure of a deity was modelled in clay in a cave in the Peten but apparently has been destroyed (Stuart & Stuart 1977:53). Both sculptures date to Late Classic times. Most cave petroglyphs are pecked images of ladder-like forms or rough frontal faces. They are often difficult to identify, let alone date or interpret. However, one outstanding bas-relief flanks an entrance to Loltun and can be dated stylistically to the Protoclassic period (Andrews, A.P. 1981). Other, more modest, Classic style petroglyphs are incised into a soft limestone wall at Caactun in central Yucatan (Stone 1995). Generally, the painted art is more interesting, though it is usually monochrome (black), and around twenty caves with paintings or drawings are known from the Maya area. The earliest paintings are also found at Loltun and are contemporary with the aforementioned Protoclassic bas-relief (Stone 1989b ). Some undatable paintings from Loltun may be earlier, given the antiquity of the archaeology. In fact, Velazquez Valadez (1980) suggests that some of the Loltun paintings date to the Late Archaic period (ca 2000 BC), based on their linear, schematic style, but this idea has no firm foundation. The inventory of painted art can be divided into two stylistic groups. One is a schematic, stick-figure style which, again, does not lend itself easily to interpretation. The most impressive cave art in the schematic style is Actun Dzib in southern Belize. The cave houses two panels of stick-figure quadrupeds, spirals, and T-shaped designs, curiously reminiscent of some rock art of the American Southwest (Stone 1995). The other style group, which I refer to as Greater Classic, more closely corresponds to art of surface sites. It is far more open to interpretation given its strong art historical context. For instance, a Late Classic three-figure scene from Actun Ch'on, Yucatan, may depict the presentation of a nude captive prior to his sacrifice in the cave (Stone 1989c). Clearly, the most spectacular of the Greater Classic cave art sites is Naj Tunich, Guatemala, which houses about ninety paintings, including dozens of hieroglyphic texts, all dating to the eighth century AD (Stuart 1981; Brady & Stone 1986; Stone 1995). Many of the texts record the visits of elite pilgrims to the cave. In some cases we know that individuals from different sites joined together for these pilgrimages, suggesting that the cave played a role in regional political integration. The figurative paintings portray the pilgrims themselves, sometimes shown sitting in a calm, meditative state, at other times engaged in various ritual acts. A few paintings depict deities. The Popol Vuh Hero Twins, Hunahpu and Xbalanque, are represented, as is a sun god. One depiction of a dwarf reflects the widespread belief in the Maya area that dwarf-spirits, often associated with rain, live in caves. About a dozen Maya caves have been reported with handprints. The largest collection of handprints, 135 in all,

can be found at Actun Acum, one of the Puuc Hill caves (Strecker 1982). The handprints are both positive and negative and occur primarily in red and black. Some of the negative handprints are quite unusual. One from Loltun shows two hands with a thin rod touching the index fingers. Animal heads are made in a stencil technique at Acum. Negative handprints using both hands are also found at Caactun.

Conclusion Maya cave archaeology has much to tell us about the mental processes of a fascinating Precolumbian people, particularly how they attributed extraordinary powers to space in the natural environment. Our ability to make sense of the archaeological data is aided considerably by the customs of living Maya who utilize caves in some of the same ways as their forebears: they continue to petition deities, bum incense, and perform blood sacrifice in caves, but today only domestic fowl are offered up to the gods. Scholars outside the field of Mesoamerican studies could benefit from looking at the Maya material in considering broader questions of how humans have thought about caves in the past and incorporated them into material and intellectual aspects of their lives.

References ANDERSON, A.H. 1962. Cave Sites in British Honduras. In Akten des 34 Internationale Amerikanisten-Kongresses, Wien, pp. 326-331. Verlag Ernst Berger, Horn, Vienna. ANDREWS, A.P. 1981. El guerrero de Loltun: comentario analitico. Boletfn de la Escue/a de Cfencias Antropl6gicas de la Universidad de Yucatan 8-9 (48--49):35-50. ANDREWS, E.W., IV 1965. Explorations in the Gruta de Chae, Yucatan, Mexico. Publication 31. Middle American Research Institute, Tulane University, New Orleans. ANDREWS, E.W., IV 1970. Balankanche, Throne of the Tiger Priest. Publication 32. Middle American Research Institute, Tulane University, New Orleans. ARNOLD, D.E. & BOHOR, B.F. 1978. Attapulgite and Maya Blue: an ancient mine comes to light. Archaeology 28(1):23-29. BASSIE-SWEET, K. 1991. From the Mouth of the Dark Cave: Commemorative Sculpture of the Late Classic Maya. University of Oklahoma Press, Norman. BLOM, F. 1954. Ossuaries, cremations, and secondary burials among the Maya of Chiapas, Mexico. Journal de la Societe des Americanistes 43:123-36. BONOR VILLAREJO, J.L. 1989. Las cuevas mayas: simbolismo y ritual. Universidad Complutense de Madrid, Madrid. BRADY, J.E. 1989. An Investigation of Maya Ritual Cave Use with Special Reference to Naj Tunich, Peten, Guatemala. Unpublished Ph.D. dissertation, Department of Anthropology, University of California at Los Angeles, Los Angeles. BRADY, J.E. & STONE, A. 1986. Naj Tunich: entrance to the Maya underworld.Archaeology 39(6):18-25. BRADY, J.E. & VENI, G. 1992. Man-made and pseudo-karst caves: the implications of subsurface features within Maya centers. Geoarchaeology 7(2):149-167. BRADY, J.E., VENI, G., STONE, A. & COBB, A. 1992. Explorations in the new branch of Naj Tunich: implications for interpretation. Mexican 14(4):74-81.

205

Stone BREUIL, V. 1986. Registro de las cuevas de la region de Xcochcax: informe de trabajo de la temporada 1986. Ms. on file, Centro Regional del Sureste del Instituto Nacional de Antropologia e Historia, Merida, Yucatan. CATHERWOOD, F. 1844. Views of Ancient Monuments in Central America, Chiapas, and Yucatan. Federick Catherwood, London. CAROT, P. 1989. Arqueologfa de las cuevas de! norte de Alta Verapaz. Centre d'Etudes Mexicaines et Centramericaines, Mexico. GATICA TREJO, R. 1980. El lugar de los ritos mayas amenezado ahora por el hombre. El Grafico, Guatemala City, June 29. GONzALEZ LICON, E. 1986. Los mayas de la gruta de Lo/tun, Yucatan a traves de sus materiales arqueol6gicos. Instituto Nacional de Antropologia e Historia, Mexico. GORDON, G.B. 1898. Caverns of Copan, Honduras. Peabody Museum of American Archaeology and Ethnology, vol. 1, no. 5. Harvard University, Cambridge, Massachussets. GRAHAM, E., McNATT, L. & GUTCHEN, M.A. 1980. Excavations in Footprint Cave, Caves Branch, Belize. Journal of Field Archaeology 7(2):153-172. HATT, R.T., FISHER, H.I., LANGEBARTEL, D.A. & BRAINERD, G.W. 1953. Fauna! and Archaeological Researches in Yucatan Caves. Bulletin 33. Cranbrook Institute of Science, Bloomfield Hills. HEYDEN, D. 1975. The cave underneath the Pyramid of the Sun at Teotihuacan. American Antiquity 40(2): 131-147. HEYDEN, D. 1976. Los Ritos de paso en las cuevas. Boletfn de! Instituto Nacional de Antropologfa e Historia epoca II, vol. 19:17-26. HEYDEN, D. 1981. Caves, gods and myths: world-view and planning in Teotihuacan. In Mesoamerican Sites and WorldViews, edited by E. Benson, pp. 1-35. Dumbarton Oaks, Washington, D.C. KAUFMAN, T. 1976. Archaeological and linguistic correlations in mayaland and associated areas of Meso-America. World Archaeology 8(1):101-118. MacLEOD, B. & PULESTON, D. 1980. Pathways into darkness: the search for the road to Xibalba. In Third Palenque Round Table, 1978, Part I (Palenque Round Table Series IV), edited by M.G. Robertson & D.C. Jeffers, pp. 71-78. Pre-Columbian Art Research Center, Palenque. McNATT, L. 1996. Cave archaeology of Belize. Journal of Cave and Karst Studies 58(2):81-99. MARQUEZ DE GONzALEZ, L., BENA VIDES, CASTILLO, A. & SCHMIDT, P.J. 1982. Exploracfon in la Gruta de Xcan, Yucatan. Instituto Nacional de Antropologia e Historia, Centro Regional del Sureste, Merida. MERCER, H. 1896. The Hill Caves of Yucatan. Lippencott, Philadelphia. MILLER, T.E. 1984. The Karst Development and Associated Archeology of the Chiquibul, Belize. Ms. on file, the National Geographic Society, Washington, D.C. MILLER, T.E. 1996. Geologic and hydrologic controls on karst and cave development in Belize. Journal of Cave and Karst Studies 58(2):100-120. PALACIO, J.O. 1977. Excavations at Hokeb Ha, Belize. Publication 3. Institute for Social Research and Action, Belize City. PENDERGAST, D.M. 1969. The Prehistory of Actun Ba/am, British Honduras. Art and Archaeology Occasional Paper 16. Royal Ontario Museum, Toronto. PENDERGAST, D.M. 1970. AH. Anderson's Excavations at Rfo Frio Cave E, British Honduras (Belize). Art and Archaeology Occasional Paper 20. Royal Ontario Museum, Toronto.

206

PENDERGAST, D.M. 1971. Excavations at Eduardo Quiroz Cave, British Honduras (Belize). Art and Archaeology Occasional Paper 21. Royal Ontario Museum, Toronto. PENDERGAST, D.M. 1974. Excavations at Actun Polbilche, Belize. Archaeology Monograph 1. Royal Ontario Museum, Toronto. REDDELL, J. 1977. A Preliminary Survey of the Caves of the Yucatan Peninsula. The Speleo Press, Austin. REENTS-BUDET, D. & MacLEOD, B. 1986. The Archaeology of Petroglyph Cave, Belize. Unpublished Ms. SAHAGUN, B. de 1963 Florentine Codex: General History of the Things of New Spain. Book 11. Translated by A.O. Anderson & C. Dibble. School of American Research/ University of Utah, Santa Fe/New Mexico. SELER, E. 1901. Die a/ten Ansiedlungen van Chacula, im Distrikte Nenton des Departments Huehuetenango der Republik Guatemala. Verlag von Dietrich Reimer, Berlin. STONE, A.J. 1989a. Las pinturas y petroglifos de Naj Tunich, Peten: investigaciones recientes. Segundo simposio de investigaciones arqueol6gicas en Guatemala, edited by S. Villagran de Brady, pp. 182-200. Ministeria de Cultura y Deportes, Gautemala. STONE, A.J. 1989b. The painted walls of Xibalba: Maya cave painting as evidence of cave ritual. In Word and Image in Maya Culture: Explorations in Language, Writing, and Representation, edited by W. Hanks & D. Rice, pp. 319-335. University of Utah Press, Salt Lake City. STONE, A.J. 1989c. Actun Ch'on, Oxkutzcab, Yucatan: una cueva maya con pinturas del cl-sico tardio. Bolet(n de la Escue/a de Ciencias Antropol6gicas de la Universidad de Yucatan 16(99):24-35. STONE, A.J. 1992. From ritual in the landscape to capture in the urban center: the recreation of ritual environments in Mesoamerica. Journal of Ritual Studies 6(1): 109-132. STONE, A.J. 1995. Images from the Underworld: Naj Tunich and the Tradition of Maya Cave Painting. University of Texas Press, Austin. STRECKER, M. 1982. Representaciones de manos y pies en el arte rupestre de cuevas de Oxkutzcab, Yucatan. Bolet(n de la Escue/a de Ciencias Antropol6gicas de la Universidad de Yucatan 12(68):21-28. STUART, Geo. 1981. Maya art treasures discovered in cave. National Geographic 160(2):220-235. STUART, Geo. & STUART, G. 1977. The Mysterious Maya. National Geographic Society, Washington, D.C. TATE, C. 1992. Yaxchilan: The Design of an Ancient Maya Ceremonial Center. University of Texas Press, Austin. TEDLOCK, B. 1985. Time and the Highland Maya. University of New Mexico Press, Albuquerque. THOMPSON, J.E.S. 1959. The role of caves in Maya culture. Hamburg Museum fur Volkerkunde und Vorgeschichte Mitteilungen 25:122-129. THOMPSON, J.E.S. 1975. Introduction to The Hill Caves of Yucatan by H. Mercer, pp. vii-xliv. University of Oklahoma Press, Norman. THOMPSON, R.H. 1958. Modern Yucatecan Maya Pottery Making. Memoirs of the Society for American Archaeology, vol. 23, no. 4, part 2. Society for American Archaeology, Washington, D.C. VELAZQUEZ VALADEZ, R. 1980. Recent discoveries in the cave ofLoltun, Yucatan, Mexico. Mexican 2: 53-55. WERNER, L. 1992. Cultivating the secrets of Aztec gardens. Americas 44(6):6-15. ZAPATA PERAZA, R.L., BENEVIDES CASTILLO, A. & PENA, A. 1991. La gruta de Xtacumbilxunaan, Campeche. Instituto Nacional de Antropologia e Historia, Mexico.

Rockshelters and Ritual Activities in the Atacama Desert of Northern Chile Penny Dransart

Abstract

Caves and rockshelters have provided sitesfor human activity since the earliest occupations at high altitude in the Andes. Two cultural areas are recognized here: the central Andes and the south central Andes. In the former, sites tend to be larger and contain more substantial cultural deposits. However, excavations in these sites have often been undertaken to address problems in understanding ways of life in the wider Andean world. In contrast, caves and rockshelters in the south central Andes usually contain more sparse occupation levels. Yet a wider range of organic material is available for study from these sites because of the prevailing arid conditions. This facilitates the interpretation of episodes of human activity inside the caves or rockshelters. Two rockshelter sites located in the Andean slopes east of the Safar de Atacama are examined in detail. Both are associated with rock art, but different activities were detected inside the shelters. Ethnographic analogy is used to interpret these activities. It is suggested that rites of passage were enacted inside Tu/an 67, and that a ritual investiture of pubic coverings and necklaces took place during the Archaic Period, or perhaps in the Early Formative Period. Site Tu/an 64, which contained a re-used standing stone, probably was used by hunters of vicuflas during the Formative Period. Since the herding of llamas and perhaps alpacas was the dominant subsistence activity in the area during the Formative, it is suggested that this site saw the enactment of rituals connected with specialized hunting activities.

Owing to the arid conditions that prevail in the area, the preservation of organic remains means that more categories of material are available for study than in other parts of the world. Here, the exploitation of the fleece of the South American camelids (vicuiia, guanaco, alpaca and llama), will be examined. The presence of organic material provides intriguing information on human activity inside the caves and rockshelters. However, before proceeding to examine the rockshelters of the Atacama in detail, it is necessary to set them in an ecological and temporal framework.

Even though humans have regularly and persistently made use of natural shelters throughout many millennia, as pointed out by Clive Bonsall and Christopher Tolan-Smith in the Preface to this volume, such natural shelters have all too often been regarded as little more than depositional environments. Instead, they suggest that we should address the question: 'what were the humans doing in the cave in the first place?'. This is a very pertinent question to ask of Andean South America, bearing in mind that the Inkas, the sophistication of whose Empire astounded the Spanish invaders in the sixteenth century, traced their origins from a cave 26km south of Cusco. This cave was known as the Pacariqtampu ('house of origin') or, alternatively, tamput'uju ('house of the windows/niches') because of its three exits from the inner world, from where the founders of the royal Inka lineage emerged into this world (Urton 1990; Dransart 1992). Even today, Andean peoples regard the inner world as a source of fertility and wealth for the inhabitants of this world. Ritual offerings are made to the earth, known as the Pachamama, that is, mother of space (the earth) and of time. The Inkas, in common with other Andean peoples, believed themselves to be the progeny of the earth itself, hence the importance of the cave site through which they entered this world. However, I wish to go back further in time and to focus upon the use of rockshelters in the Atacama Desert of northern Chile. It is important to stress that many of these sites are still used as overnight shelters by pastoralists of the Atacama in the course of their herding activities. The archaeological period of the sites considered in detail here covers the Archaic to the Formative periods. In the north of Chile, the Archaic began over ten thousand years ago, and the transition from Late Archaic to Early Formative occurred about 3500 to 3000 years ago.

Andean Landscapes The sites discussed here are located above 2500m asl, that is, in the sierra or the Andean highlands (Fig. 1). To the west of the Andean chain of mountains there is a coastal desert, bordering the Pacific. It is ribbon-like in Peru, but it broadens as it extends into the Atacama Desert of Chile. The highlands are both intersected by inter-andean valleys and dominated by ranges of high peaks. Towards the south of Peru, and along the length of the Chilean-Argentinian border, is a chain of volcanic peaks. Classifications for different ecological zones in the Andes based on observed biota have been listed by Craig (1985). These schemes demonstrate the occurrence of vegetational zones according to altitude. Craig suggests that the now relict tropical rainforest cover on the western flanks of the Andes in northern Peru was formerly more extensive along the foothills of the western Andes, becoming established at the beginning of the Holocene (Craig 1985) Postglacial desiccation, according to Craig, did not have a significant effect on the aridity of the coastal desert, which has retained its xeric character throughout most of the Late Pleistocene to the present (Craig 1985:27). At the end of the

207

Dransart '(t

/4~

~

Jl ~~

i (i I

,.. Guitarrero Cav Quishqui Pun Lauri co ch Pachamach Panaulauc ~

~.

' ).::J \~

'\,.,

''

'-~

~-:::..:: "" Toquep

,,.

.\. ARICA~""~ ~

)

'~

.Ji

La Capilla

\... ./'

'

)··t::❖:

{ .-i/.)\.

I

\\··:/~;;)

Tuina/j

. \

J

c)._.,·

Puripica San Lorenzo/.e '.c:;

i/

'·~

:-I \.

A

\

11

)

,\I

Km

0 Figure 1

200

I\

400

;'·\:···

Tula~..,..-'

SA

.

)

600

Map of the central and south central Andes showing sites mentioned in the text. Sites are represented by triangles in the central Andes and by circles in the south central Andes. Solid symbols denote caves and rockshelters, while outlines denote open-air sites.

208

Rockshelters and Ritual Activities in the Atacama Desert of Northern Chile Table 1:

Radiocarbon determinations for cave and rockshelter sites in the south central Andes. All dates are uncalibrated and cited in years BP.

Site

Level

Lab.No.

14C Age

Material

Reference

(yr BP)

Tuina 1 Tuina 1 San Lorenzo 1 San Lorenzo 1 San Lorenzo 1 Inca Inca Inca Inca

Cueva Cueva Cueva Cueva

Level V Level IV

c4 c4 c4 c4

Tulan 68

SI-3112 N-????

10,820±630 9080±130

Charcoal Charcoal

Nunez and Santoro (1988) Nunez and Santoro (1988)

Hv-299 N-3423 N-3424

10,280±120 10,400±130 9960±125

Charcoal Charcoal Charcoal

Geyh (1967) Nunez (1983) Nunez (1983)

AC-564 CSIC-4980 LP-102 LP-137

9900±200 9230±70 9650±110 10,620±140

Charcoal Stem & leaves

Aschero (1980) Aschero (1980) Figini et al. (1990) Figini et al. (1990)

Beta-25532

9290±100

Charcoal

Nunez (1988)

Tulan67 Tulan 67 Tulan67

Level VII Level V Level VII

Beta-25535 OxA-1842 OxA-1843

8190±120 5320±90 4870±65

Charcoal Camelid fibre Camelid fibre

Nunez (1988) Hedges et al. (1989) Hedges et al. (1989)

La Capilla 1 La Capilla 1 La Capilla 1

Level 4 Level4 Level 4

Gak-8778 Gak-8777 1-11,642

3670±160 2790±140 3450±90

Bone Bone ?

Munoz and Chacama (1982) Munoz and Chacama (1982) Munoz and Chacama (1982)

Beta-25530

1540±80

Charcoal

Nunez (1988)

Tulan 64

last glaciation, it would seem that the steppe-like grasslands, which constitute the puna vegetation at high altitudes, became more extensive with the upward shrinking of the permanent snow cover on high peaks. The Andean area has been divided into a series of geographically located cultural areas that do not conform to modem political boundaries (Lumbreras 1981:42). According to this division, the Atacama rockshelters which will be considered here fall within the south central Andes. This unit includes southern Peru, northern Chile, part of Bolivia, and northwestern Argentina. The central Andes is the cultural area immediately to the north of the south central Andes, and it consists of the greater part of Peru. If the highland sierra of the central Andes has been characterized as a harsh and unpredictable habitat with pronounced wet and dry seasons interspersed with occasional lengthy droughts (Wheeler 1984b:196), conditions may be said to be considerably worse at high altitudes in the south central Andes. The weather conditions are characterized by dry, extremely windy winters and summer by precipitation that often falls as snow above 3000m asl (November to March). The low humidity in the highlands and in the Atacama Desert to the west produces diurnal temperature changes with a daily variation reaching up to 35 °C between a daytime high and nighttime low (Popper 1977:9).

occupations seem to have involved the use of caves and rockshelters, and authors have pointed to the similarities that have been observed in the stone tools in both areas. This section will review archaeological interpretations on the basis of work conducted in the central Andes. It will then be possible to evaluate more clearly the situation in the south central Andes. The very earliest occupations, predating 12,000 BP, have been detected at Pikimachay (Ayacucho Complex) near Ayacucho (MacNeish et al. 1981); Pachamachay (Rick 1980) and Uchkumachay (Wheeler et al. 1976), both situated on the Puna de Junin; and Guitarrero Cave, above the Rio Santa in the Callej6n de Huaylas (Lynch 1980). These are all caves or rockshelters which continued to be used in later periods. The earliest levels tend to contain slight amounts of archaeological material, and it is not until 11,000 to 10,000 BP onwards that abundant cultural deposits are present at such sites. From this time, the Lauricocha caves, Huanuco (Cardich 1964) and Telarmachay, on the Puna de Junin (Lavallee et al. 1985), should be added to the list. Given the high density of chipped stone tools at these sites, one of the main concerns in the published literature has been that of examining the lithic industries represented and to establish stone tool traditions which emerged in the area. For example, Lynch has proposed a 'Central Andean Preceramic Tradition' for the central and north central sierra of Peru between 11,000 and 3800 BP (Lynch 1980:301), but Rick has expanded this concept to include the Ayacucho area as well (Rick 1988:17). These findings have been developed to argue for a sedentary lifestyle, relying on the hunting of vicufias as the basis of the subsistence economy at Pachamachay (Rick 1980), or for a way of

The Archaeological Background i) The Central Andes

Although, in time, the central Andes and the south central Andes were both incorporated into the expansive Inka Empire, their cultural dynamics did not follow the same trajectories. However, in both areas the earliest detected

209

Dransart N

t 0

5

10

15

20

25

30km

Figure 2

Map of the Tulan Quebrada in the south central Andes showing sites mentioned in the text.

lates in the highlands (vicufia, guanaco and huemul deer) to the specialized hunting of the vicufia and guanaco. This constitutes early evidence for the long and close association that was to emerge between human and camelid populations in the Andes. The considerable decline in percentages of cervids accompanied by a corresponding increase of camelids in faunal assemblages from highland sites in the central Andes has been observed by Wing (1972), Wheeler (1984a) and Moore (1988). However, it is the site of Telarmachay at 4420m asl, where more than 400,000 animal bones were excavated, that has provided the most detailed accounts to be published to date (Wheeler 1984a; Lavallee et al. 1985). Changes in dentition in the camelids are seen to provide evidence for the domestication of vicufias in the form of alpacas from 6000 to 5500 BP (Wheeler 1984a:402). This interpretation is supported by the high proportions of new-born animals at Telarmachay which may be due to mortality caused by disease (such as enterotoxaemia) related to the herding of animals in enclosed spaces (Wheeler 1984a:403). These high numbers of camelid foetus/neonate remains also imply that llamas were being domesticated (Lavallee et al. 1985:72). Thus the site of Telarmachay heralded the coexistence of humans and herded animals that was to become so characteristic of life at high altitudes in the Andes.

life which involved seasonal mobility or nomadic movement between various sites, not all located in the same ecological niche (Lynch 1980; Lavallee et al. 1985; Lavallee 1990). Evidence for plant remains from the cave sites provides another theme for investigation. Early plant domesticates have been claimed for Guitarrero Cave, at an altitude of 2580m asl. Two· types of bean, bottle gourd and a single chili pepper were found in a level the beginning of which dates from 10,600 BP (Smith 1980; but see also comments by Vescelius 1981b). At a much higher altitude (4300m asl), the cave of Pachamachay contained, among other plant remains, Chenopodium seeds. Those recovered from preceramic levels compared well in size to the wild species of Chenopodium, but notably larger seeds occurred in the ceramic levels after 3500 BP, suggesting that this resource was either tended in some way, or that cultivation had started to take place (Pearsell 1980:196-199). Yet another category that has received attention is animal bone remains, of which substantial amounts have been recovered from the aforementioned sites, in addition to the faunal assemblage from Panaulauca, located on the Puna de Junin (Moore 1988). Faunal analyses have revealed that a development occurred from hunting all the Andean ungu-

210

Rockshelters and Ritual Activities in the Atacama Desert of Northern Chile Hence it can be seen that excavations in caves and rockshelters in the central Andes have been undertaken to address problems in understanding ways of life in the wider Andean world. As Smith and Bonsall have suggested, these sites have been regarded as little more than depositional environments (an exception to this generalization being Lavallee et al. 1985). A further point remains to be made regarding the central Andes. Virtually all the known sites of the preceramic period in the highlands are caves and rockshelters. The apparent absence of open-air sites, with only two notable exceptions (Quishqui Puncu in Ancash and Ambo in Huanuco) has been noted by Rick (1988:24) and Lavallee (1990:110).

that are geographically distant. Archaeologists have been tempted to fill in the gaps by calling upon sites situated beyond the boundaries of the south central Andes, as does Nunez in citing Intihuasi in San Luis, Argentina, when faced with an absence of inland sites between 9000 and 6000 BP (Nunez 1983:195). Sites in this area have not been the subject of major monographs as in the central Andes, with the result that detailed specialist studies are often lacking. In any case, it would seem that bone remains in particular are more scarcely represented in the deposits of caves and rockshelters in the south central Andes. Brian Hesse examined 359 and 108 bone remains from Tuina 1 and San Lorenzo 1, respectively. In the Tuina sample he recognized camelid and rodent bones, and in the San Lorenzo sample, camelid, rodent and bird bones (Hesse 1984:54). These samples are far too small for the purposes of faunal analyses. In fact, it is an open-air Late Archaic site, Puripica 1, near San Pedro de Atacama, at which the domestication of camelids was recognized. The analysis undertaken by Hesse indicated that the guanaco was brought uµder domestication, and that this change took place between levels ill and II, during the fifth millennium BP (Hesse 1982a:210, 1982b:ll). The occupation of this site during the Late Archaic was limited in duration, and Baied & Wheeler comment that the lack of time depth represented makes it difficult to determine whether herds were introduced into the area or if domestication took place at Puripica (Baied & Wheeler 1993:148-149). It is true that cultural material is more sparsely represented in the south central Andean cave and rockshelter sites, and there is a concern shown by archaeologists for using that material to interpret the adaptive strategies adopted by the inhabitants of those sites in the wider world. Yet a wider range of cultural material is present in the deposits, due to the arid conditions that favour the preservation of organic remains. Thus it is possible to interpret more clearly episodes of human activity within these sites than is the case with the central Andean sites. The double site of Toquepala, consisting of a cave and rockshelter, the walls of both of which display mainly painted rock art, provides a case in point. Excavations in the cave brought to light two thin sticks, each of which had one end wrapped in fleece, stained with red pigment, in the manner of a paintbrush (Muelle 1963:191). A different activity was detected at Inca Cueva cave 4, where stone tools were reported as retaining animal hairs on the edges, suggesting that the preparation of hides took place in the cave (Yacobaccio 1984:74-75).

ii) South Central Andes Evidence for human occupation in this area has been detected from about 11,000 BP. Like the central Andes, early sites are reported as being located in small caves or rockshelters. These include Inca Cueva cave 4 in the Department of Humahuaca, northwestern Argentina (Aschero 1984), Tuina 1, east of Calama, and San Lorenzo 1, east of Toconao, both in the north of Chile (Nunez 1983) (Fig. l; Table 1). An open-air site, Acha-2, Arica, has been published by Munoz Ovalle et al. (1993). Many of the published references to these sites have to do with the lithic industries they contain, for example, similarities have been seen in the bifacially-worked triangular points encountered at Tuina and Inca Cueva cave 4 (Nunez & Santoro 1988:26). However, the sites known to date are somewhat scattered and comparisons have been made between sites

~

0

oo

The Tulan Quebrada

1m

L__J

The two sites to be discussed in detail here are Tulan 67 (TU 67) and Tulan 64 (TU 64). They are situated above the Tulan Quebrada, one of a number of valleys containing intermittent streams that cut through an arid rhyolitic plateau from the high altitude terrain known as the puna, with its chain of volcanoes and highland lakes. The valleys descend westward to the Salar de Atacama, a salt basin with a few freshwater lakes seasonally occupied by flamingos.

...... c:::::,.,. .. •·

Figure 3

Plan of site Tulan 67, showing the location of grid square 4. The dotted line represents the overhang of the rock, and the hatched areas contemporary corral walling. Drawn by the author.

211

Dransart

Over 100 archaeological sites have been identified in the vicinity of the Tulan Quebrada (Nufiez 1988). These sites are often situated next to or underlie corrals and rockshelters which are currently used by herders of llamas, sheep and goats in a seasonal cycle of exploitation of the different pasture grounds (Dransart 1991a:280-283). When moving along the Tolan Quebrada, the herders spend the night in rockshelters surrounded by drystone walling which divides the shelter for the humans from the adjacent corrals occupied by their animals. The earliest human inhabitants of the Tolan Quebrada were Archaic hunters and gatherers. Evidence for the earliest occupations is slight, and there are large gaps in the available radiocarbon dates (Table 1; see also Dransart 199lb:307). After about 5000 BP more abundant evidence is available and open-air sites are added to the archaeological inventory (Fig. 2). To date, evidence for the domestication of camelids has not been forthcoming from the Tolan Quebrada itself, but it has been detected at the site of Puripica 1, as mentioned above. However, the open-air site of TU 54, which is dated between 2840±60 BP and 3080±70 BP, does contain evidence for the presence of domesticated camelids from the initial occupation of the site, on the basis of an examination of the fibre remains (Dransart 199lb:312-313). Yet despite the presence of herded animals, the people continued to practise hunting and gathering activities (Hesse 1984; Holden 1991; Dransart 1991a:399-400). In sum, it can be said that at the beginning of the Archaic Period, subsistence activities involved hunting and gathering, but towards the end of the Late Archaic herded animals were incorporated into the economy. Thereafter, hunting and gathering were to diminish in importance; today they constitute only residual activities. The use of sites TU 67 and TU 64 will be considered against this backdrop of the activities practised by people in the wider world.

Figure4a

Stylized human figure wearing what appears to be a fringed pubic covering. Painted in red on the rear wall of site Tulan 67. Not to scale. Drawn by the author.

~ /~-·

Tulan 67 This rockshelter is also known as Tchulin (Nufiez 1988:205). It is situated beneath an enormous overhanging rock and contemporary drystone walling has been built under and around this rock (Fig. 3). Beyond the shelter of the overhang, and on the surface of the ground, is an arc of small upright stones. In this respect, the plan of TU 67 resembles a number of rockshelters in the area, such as TU 68, a preceramic site (Table 1). At TU 67, a large scatter of surface material, mostly lithic, covers the ground on both sides of the arc of stones. Inside the shelter, the rock wall bears linear designs painted in red. Included are outlines of camelids, geometric designs and a human figure, apparently wearing some form of pubic covering (Fig. 4a). Previous excavations in the rockshelter of a trench 3m long and lm wide have not been published. The account given here is based on the excavation in September 1987 of grid square 4, actually a rectangle measuring lm by 1.5m (Nunez 1988; Dransart 1991a:297-303). Items recovered from grid square 4 include lithic material, animal bones and vegetal material from all seven levels; bird beaks were

Figure 4b

Human figures wearing fringed pubic coverings painted on a rock near the confluence of the Salado and Caspana rivers in northern Chile. Not to scale. After Le Paige (1965:plate 17b).

present in most levels. Two pieces of shaped wood, painted red, were found in level IV, and small sticks painted red or pink were encountered in levels III, IV, V and VII. Level VII also contained animal bones painted red. A piece of

212

Rockshelters and Ritual Activities in the Atacama Desert of Northern Chile tabular stone in level III was used to grind red pigment, and a small piece of red mineral came from level IV. An undiagnostic pottery sherd was found in level I, and another in level III. Small disc shell beads, similar to those from Late Archaic/Early Formative sites in the area, appeared in levels III and IV. Oval shell beads and a broken piece of worked and pierced mussel shell were located in level V. To this list of items should be added camelid dung in all levels apart from IV, where a small amount of rodent droppings was observed. From an examination of the stone tool types present in the different levels, Nunez (1988:207) proposed that levels V-VII are precerarnic. Yet level VII also contained a fragmentary textile. It most probably dropped down from an upper level, since the radiocarbon dates obtained for level VII are surely far too early for a loom product, perhaps woven with the aid of heddles, in the north of Chile (Dransart 1991a:374-375). In fact, there is considerable evidence which demonstrates that the entire stratigraphy of TU 67 is disturbed, as suggested by the inversion of the AMS dates for levels V and VII (Table ]). Not only were rodent droppings present in level IV, but the deposits from this level were found to contain a large beetle that was still alive. Rodent and insect activity would allow items to drop to lower levels. Snailshell was present in level III, and an alfalfa seed (a Spanish introduction) was observed in level II (Holden, pers. comm.). Hence no level may be assumed to be intact at TU 67. Apart from the textile fragment, no yarns were found in level VII, nor in level VI. However, yarns and animal fibres were present in levels I to V. Fifty-five items were inventoried as yarns, and they are all of camelid fibre, except for one strip of vegetal material that had been wound and knotted round some now missing object. The yarns, the textile fragment and a piece of basketry from level VII have been analyzed by Dransart (1991a:493-496). Given the emphasis on the use of red at this site (the items of painted wood in levels III, IV, V and VII; the palette for grinding red pigment and red mineral in levels III and IV; and animal bones painted red in level VII), it is perhaps surprising that only four of the camelid yarns should show signs of having been dyed red. This is a similar proportion to the 57 dyed red yarns out of a total sample of 702 yarns from grid squares 1 to 6 at the open-air site of TU 54 (4 out of 55 yarns at TU 67, compared with 57 out of 702 at TU 54). However, there are differences that may be observed between the yarn assemblages at the two sites. For example, the TU 67 yarns tend to be somewhat finer in diameter than the TU 54 yarns. Approximately half of the grid square 4 yarns at TU 67 (27 out of 55) have a diameter of 1mm or less, but only 100 out of 702 from grid square 1 to 6 at TU 54 have similar fine diameters. It is probable that yarn making activities carried out at the rockshelter served different, or more restricted purposes than at the open-air site. Analysis of the TU 54 assemblage indicated that yarns served many different purposes, including quickly made and soon discarded pieces that were used to tie things together, snares of vegetal material for catching birds, wellmade yarns with a diameter of 2mm for making into fabrics,

and finer yarns (Dransart 1991a:380-401). Given the abundant presence of small beads of shell, bone and stone at Archaic and Early Formative sites in the area, it is possible that yarns with a diameter of 1mm or less were used for stringing beads together. Human remains from the Tolan Quebrada, and more notably from the Arica area of northern Chile, show that shell beads were used as necklaces and as pendants hanging from turbans consisting of large amounts of yarn wrapped round the head. It is possible that some of the activities undertaken at TU 67 had to do with the adornment of the human body. In this respect, the stylized human figure which appears to wear a fringed pubic covering painted on the rear wall of the rockshelter is of great interest (Fig. 4a). Unfortunately, it is not possible to correlate the painting with any of the archaeological levels. However, it is worth drawing a parallel with the coastal site of La Capilla 1, which is a small cave immediately south of Arica, where 23 fringed pubic coverings were found, folded and tied, in the archaeological deposits (Munoz & Chacama 1982:plate 8). Radiocarbon determinations for La Capilla indicate an Archaic Period occupation (Table I). Two panels of geometric, zoomorphic and solar designs are painted on the walls of La Capilla, and the excavators considered that, in the absence of hearths, the site served as a cult centre where rites of passage were enacted, during which youths probably put on pubic coverings (Munoz & Chacama 1982:42-45). Although the La Capilla garments were made of vegetal fibre, it is more likely that camelid fibre would have been used for such garments in the highlands, if indeed they were worn as suggested by the TU 67 rock painting. Three human figures wearing fringed pubic coverings are painted on rock near the confluence of the Salado and Caspana rivers (Fig. 4b; Le Paige 1965:plate 17b), so the Tulan example is not an isolated one. In the Formative Period, camelid fibre fringed pubic coverings are known from cemeteries at Tarapaca 40 (Nunez 1970:86) and Azapa 71 (Azapa phase) (Santoro 1981:41). At TU 67 it is not possible to distinguish between the yarns of the Archaic and Formative periods owing to the disturbed character of the deposits. Unlike the securely dated Early Formative site of TU 54 with its densely packed midden deposits, the cultural material is slight. The timescale represented at the rockshelter is a long one, but within that stretch of time it was perhaps only used for a few brief events. Since charcoal is present in the TU 67 deposits, hearths may be present, unlike La Capilla, located in an area where the weather conditions are much more benign than in the Tolan Quebrada. However, TU 67 rockshelter may also have been a location where rites of passage were conducted, rather than serve as a dwelling place. Whether the humans who practised those rituals were hunters and gatherers, or had incorporated the herding of llamas into their hunting and gathering way of life, is not yet known.

Tulan 64 The site of TU 64 is one of a number of archaeological sites (rockshelters and open-air sites consisting of clustered

213

Dransart

buildings with a circular ground plan) situated near Tolan at approximately 2900m asl on the south side of the Tolan Quebrada (Fig. 2). It is positioned near where a footpath begins a steep descent from the high ground into the valley, down to the Tolan spring which feeds the Tolan river. TU 64 is, in fact, an indent at the bottom of an expanse of rock face which has been used as the back wall of a modem rectangular corral. The site itself is among large blocks of fallen rock, the various faces of which have been incised with drawings. The rock art was evidently done at different times and periods, for it includes presumably pre-hispanic panels of camelids (Fig. 5) and also a post-hispanic human figure on a quadruped with reins, as well as incised crosses.

,:···'

···· ..,\

0

arrows. A late eighteenth century technique for hunting vicuiia was reported by the governor of Potosi. He described how hunters laid in wait with dogs near a source of water where vicufias went down to drink during the



..

10 cm

Figure 5

Incised camelids on a large fallen block immediately east of the excavation at site Tulan 64. Drawn by the author.

Excavation of one square metre took place in October 1987 in an area free of fallen rock beneath the overhang of the rockshelter. Ceramic sherds were found in nearly all levels, including fine polished wares and coarser burnished pottery, but camelid dung was concentrated in upper levels. Also in the upper levels were five barbed-and-tanged arrowheads, all of which were broken (Dransart 1991a: plate 7.8). A scant amount of lithic waste was present in the deposits. As the excavation progressed, a standing stone (the top of which had protruded above the surface of the ground before excavation began) was uncovered. This stone was inserted into a socket dug into sterile soil and it was held in place by a mound of small stones on one side of the socket. The standing stone had been re-used; when it was removed an incised geometric design was discovered at the end which had been placed downward in the socket (Fig. 6). Although only one radiocarbon determination is available for TU 64 (Table ]), the occupation of this site is clearly more recent than TU 67. The presence of the ceramics confirms this, and perhaps also the perspective view of the camelids in the rock art (Fig. 5). Late Archaic drawings of camelids at Puri pica, dated to the fifth millennium BP, present animals in profile, showing only two legs (Dransart 1991b:315). However, this distinction is not made by Berenguer (1995) in his study of camelids in rock art of the Atacama. Even though TU 64 was occupied when the herding of camelids had become the dominant economic activity in the Tulan Quebrada, activities at this rockshelter evidently had to do with the hunting of wild camelids using bows and

0

10 cm

c:::::E3

Figure 6

An incised and re-used standing stone from the excavation at Tulan 64. Drawn by the author.

midday heat. When the vicuiias finished drinking and started to return uphill, the hunters released the dogs, which easily brought down the prey since vicuiias are less agile with the stomach weighted down with water (Canete y Dominguez 1974:249). Site TU 64 is suitably located for the ambush of animals returning from the spring below. Given the presence of a standing stone and rock art, the hunting of vicuiia may have been an activity which implicated ritual observances of some kind.

Discussion and conclusions The two rockshelters considered here are small in size and the cultural material contained therein is far less abundant compared with central Andean sites. There is a need to establish a much more rigorous chronology, especially for sites such as TU 67 which were perhaps used only for brief episodes during a very long timespan. To be fair, this caveat does not only apply to the south central Andes. Caution in interpreting the stratigraphy should also be applied to the apparently more densely occupied central Andean sites, a

214

Rockshelters and Ritual Activities in the Atacama Desert of Northern Chile

case in point being the much debated Guitarrero Cave (Vescelius 1981a,1981b; Lynch et al. 1985). Yet the brief episodes of human activity left in the two rockshelters discussed here raise some interesting issues. Therefore, it would be misguided to turn one's attention wholly to openair sites such as Puripica 1 and TU 54, with their architectural remains and dense midden deposits accumulated over a shorter period of time. Activities proposed as having taken place in TU 67 were elucidated by evidence brought to light at the coastal site of La Capilla 1, with a more restricted timescale limited to the Late Archaic (Mufi.oz & Chacama 1982:9-10). It was suggested that rites of passage took place at these sites, with initiates undergoing a ritual investiture of garments (pubic coverings, shell necklaces and perhaps yarn turbans). In more recent times, it was recorded that the Inkas and other ethnic peoples in the Andes believed that their ancestors emerged from specific points of origin in the landscape, which were venerated. The Inkas considered themselves to have come from a cave site, the Pacariqtampu. We know from the chronicles written following the Spanish invasion of 1532 that noble Inka boys were ritually invested with a breechcloth at the same time as they had their ears pierced during initiation ceremonies. Cristobal de Molina, a priest born in Cusco, recorded that the boys' mothers spun the yarn for the garments, helped by other family members, during a festive occasion accompanied by drinking (Molina 1947:92). It is interesting to note that the ritual investiture itself took place at Huanacauri, a sacred site of the Inkas which marked the point where one of the male Inka ancestors who had emerged from Pacariqtampu was transformed into stone (Molina 1947:94). It is possible that the activities at La Capilla 1 and TU 67 may be earlier instances of such ceremonies which reached a greater elaboration in Inka times. With the arrival of Christian missionaries in the wake of the Spanish invasion, such religious observances were suppressed. However, this has not been the case with herded animals. Herders today in the Andes observe elaborate and complex festivities which constitute a rite of passage for their herds. The ears of animals are cut and pierced with tassels which make the ears bleed, and they are ritually 'dressed' with dyed yarns and fleece. Spahni (1962) published an account of such ceremonies in the Atacama. In contrast, the hunting of wild vicuiias was suggested as having taken place at TU 64, if the arrows were used for hunting quadrupeds as implied by the rock art. The presence of a standing stone is somewhat enigmatic, but evidently at some stage the site was 'exorcised' by the addition of incised crosses to the rock art, even though by that time much of the standing stone would have been lost in the deposits. Although both TU 67 and TU 64 display rock art, the activities which are suggested as having taken place are markedly different. In northwestern Argentina, Yacobaccio (1984:80) suggested that cave and rockshelter sites may be differentiated by the main activity represented at each site. We should therefore expect to find traces left by a variety of activities. What the ethnographic evidence makes clear is that the users of caves did and do not merely regard them as

places in which to discard or leave items. Instead these sites were located at significant focal points of the landscape and provide entrances into the earth itself. Acknowledgements: I wish to thank Lautaro Nunez for including me in his team working for a Chilean Fondecyt funded project, based at the Instituto de Investigaciones Arqueol6gicas R.P. Gustavo Le Paige, in San Pedro de Atacama. I would like to extend my thanks to the staff of the Institute for their help. Finally, I gratefully acknowledge financial support from the Pirie-Reid Scholarship Fund, the Emslie Hornirnan Scholarship Fund and the British Council in Santiago.

References ASCHERO, C. 1980. Comentarios acerca de un fechado radiocarb6nico del sitio Inca Cueva-4, Departamento Humahuaca, Jujuy, Argentina. Relaciones 14(1):165-168. ASCHERO, C.A. 1984. El sitio ICC-4: un asentamiento preceramico en la Quebrada de Inca Cueva (Jujuy, Argentina). £studios Atacamefios 7:62-72. BAIED, C.A. & WHEELER, J.C. 1993. Evolution of high Andean puna ecosystems: environment, climate and culture change over the last 12,000 years in the Central Andes. Mountain Research and Development 13(2):145-156. BERENGUER, R.J. 1995. El arte rupestre de Taira dentro de los problemas de la arqueologia atacameiia. Chungarti 27(1): 7-43. CANETE Y DOMINGUEZ, P.V. 1974(1791]. Documentos Numeros 1 y 2. Norte Grande 1(2):233-251. CARDICH, A. 1964. Lauricocha. Fundamentos para una prehistoria de los Andes Centrales. Studia Praehistorica 3. CRAIG, A.K. 1985. Cis-Andean environmental transects: late Quaternary ecology of northern and southern Peru. In Andean Ecology and Civilization, edited by S. Masuda, I. Shimada & C. Morris, pp. 23-44. University Press, Tokyo. DRANSART, P. 1991a. Fibre to Fabric: the Role of Fibre in Came/id Economies in Prehispanic and Contemporary Chile. Unpublished DPhil dissertation, University of Oxford. DRANSART, P. 1991b. Llamas, herders and the exploitation of raw materials in the Atacama Desert. World Archaeology 22(3):304-319. DRANSART, P. 1992. Pachamama: the Inka Earth Mother of the Long Sweeping Garment. In Dress and Gender: Making and Meaning in Cultural Contexts, edited by R. Barnes & J.B. Eicher, pp. 145-163. Berg, New York and Oxford. FIGINI, A.J., CARBONARI, J.E. & HUARTE, R.A. 1990. Museo de La Plata radiocarbon measurements II. Radiocarbon 32(2):197-208. GEYH, M.A. 1967. Hannover radiocarbon measurements IV. Radiocarbon 9:198-217. HEDGES, R.E.M., HOUSLEY, R.A., LAW, I.A. & BRONK, C.R. 1989. Radiocarbon dates from the Oxford AMS system: Archaeometry datelist 9. Archaeometry 31(2):225-234. HESSE, B. 1982a. Archaeological evidence for camelid exploitation in the Chilean Andes. Saugetierkande Mitteilungen 30(3):201-211. HESSE, B. 1982b. Animal domestication and oscillating climates. Journal of Ethnobiology 2(1):1-15. HESSE, B. 1984. Archaic exploitation of small mammals and birds in northern Chile. Estudios Atacamefios 7:42-61. HOLDEN, T.G. 1991. Evidence of prehistoric diet from northern Chile: coprolites, gut contents and flotation samples from the Tulan Quebrada. World Archaeology 22(3):320-331. LAV ALLEE, D., JULIEN, M., WHEELER, J. & KARLIN, C. 1985. Telarmachay: Chasseurs et Pasteurs Prehistoriques des Andes. Editions Recherche sur les Civilisations and Institut Fran9ais d'Etudes Andines, Paris.

215

Dransart LAVALLEE, D. 1990. Le preceramique de la sierra peruvienne. In Inca - Peru 3000 Ans d' Histoire, edited by S. Purin, pp. 106-117. Musees Royam: d' Art et d'Histoire, Brussels. LUMBRERAS, L.G. 1981. Arqueologia de la America Andina. Editorial Milla Batres, Lima. LYNCH, T.F. (ed.) 1980. Guitarrero Cave: Early Man in the Andes. Academic Press, New York. LYNCH, T.F., GILLESPIE, R., GOWLETT, J.A.J. & HEDGES, R.E.M. 1985. Chronology of Guitarrero Cave, Peru. Science

PEARSALL, D.M. 1980. Pachamachay ethnobotanical report: plant utilization at a hunting base camp. In Prehistoric Hunters of the High Andes, by J.W. Rick, pp. 191-231. Academic Press, New York. POPPER, V.S. 1977. Prehistoric Cultivation in the Puna de Atacama, Chile. Unpublished undergraduate dissertation, Harvard University. RICK, J.W. 1980. Prehistoric Hunters of the High Andes. Academic Press, New York. RICK, J.W. 1988. The character and context of highland preceramic society. In Peruvian Prehistory, edited by R.W. Keatinge, pp. 3-40. University Press, Cambridge. SANTORO VARGAS, C. 1981. Formative temprano en el extreme norte de Chile. Chungara 8:33-77. SMITH, C.E. Jr 1980. Plant remains from Guitarrero Cave. In Guitarrero Cave: Early Man in the Andes, edited by T.F. Lynch, pp. 87-119. Academic Press, New York. SPAHN!, J.C. 1962. L' 'enfloramiento' ou le culte du lama chez les indiens du Desert d' Atacama (Chili). Bulletin de la Societe Suisse des Americanistes 24:26-36. URTON G. 1990. The History of a Myth: Pacariqtambo and the Origin of the Inkas. University of Texas Press, Austin.

229:864-867.

MACNEISH, R.S., GARCIA COOK, A., LUMBRERAS, L., VIERRA, R.K. & NELKEN-TERNER, A. 1981. Prehistory of the Ayacucho Basin, Peru: II. Excavations and Chronology. University of Michigan Press, Ann Arbor. MOLINA, C. de ('El Cuzqueiio') 1947(1574]. Ritos y Fabulas de los Incas. Editorial Futuro, Buenos Aires. MOORE, K.M. 1988. Hunting and herding economies on the Junin Puna. In Economic Prehistory of the Central Andes, edited by E.S. Wing & J.C. Wheeler, pp. 154-166. BAR International Series 427. British Archaeological reports, Oxford. MUELLE, J.C. 1969. Las cuevas y pinturas de Toquepala. In Mesa Redonda de Ciencias Prehist6ricas y Antropol6gicas, vol. II, Seminario de Antropologfa del Institute Riva-Aguero, pp. 186-196. Pontificia Universidad Cat6lica del Peru, Lima. MUNOZ OVALLE, I. & CHACAMA RODRiGUEZ, J. 1982. Investigaciones arqueol6gicas en las poblaciones precerarnicas de la costa de Arica. In Documentos de Trabajo no. 2, edited by I. Munoz & J. Cordova, pp. 3-93. Universidad de Tarapaca, Arica. MUNOZ OVALLE, I., ARRIAZA TORRES, B. & AUFDERHEIDE, A. (eds). 1993. Acha-2 y los Origines de[ Poblamiento Humano en Arica. Ediciones Universidad de Tarapaca, Arica. NUNEZ A. L. 1970. Algunos problemas del estudio del complejo arqueol6gico Faldas del Morro del Norte de Chile. Abhandlung und Berichte des Staatlichen Museums fur Volkerkunde Dresden 31:79-109. NUNEZ, L. 1983. Paleoindian and Archaic cultural periods in the arid and semiarid regions of northern Chile. Advances in World Archaeology, vol. 2, edited by F. Wendorf & A.E. Close, pp. 161-203. Academic Press, New York. NUNEZ, L. 1988. Analisis multidisciplinario de domesticacion y crianza inicial de camelidos en los Andes de! norte de Chile. Unpublished report for Fondecyt Project 1017-86. NUNEZ, L. & SANTORO, C.M. 1988. Cazadores de la puna seca y salada del area centro-sur andina (Norte de Chile). Estudios Atacamefios 9:11-60.

VESCELIUS, G.S. 1981a. Early and/or not-so-early man in Peru: the case of Guitarrero Cave, part 1. The Quarterly Review of Archaeology 2(1):11-15. VESCELIUS, G.S. 1981b. Early and/or not-so-early man in Peru: Guitarrero Cave revisited. The Quarterly Review of Archaeology 2(2):8-20. WHEELER, J.C. 1984a. On the origin and early development of camelid pastoralism in the Andes. In Animals and Archaeology. 3: Early Herders and their Flocks, edited by J. CluttonBrock & C. Grigson, pp. 395-410. BAR International Series 202. British Archaeological reports, Oxford. WHEELER, J.C. 1984b. Review of Prehistoric Hunters of the High Andes, by J. Rick. American Antiquity 49: 196-198. WHEELER PIRES-FERREIRA, J., PIRES-FERREIRA, E. & KAULICKE, P. 1976. Preceramic animal utilization in the central Peruvian Andes, Science 194:483-490. WING, E.S. 1972. Appendix IV: Utilization of animal resources in the Peruvian Andes. In Excavations at Kotosh, Peru: a Report on the Third and Fourth Expeditions, edited by S. Izumi & K. Terada, pp. 327-351. University Press, Tokyo. YACOBACCIO, H.D. 1984. Aproximaci6n a la funci6n de los asentamientos preceramicos en la puna y su borde oriental (Jujuy, Argentina). Estudios Atacamefios 7:73-84.

216

The Human Use of Caves Christopher Tolan-Smith & Clive Bonsall

There are few aspects of the human experience that are such worldwide phenomena as the use of caves and few natural features in the landscape that have so excited the imagination. The papers published in this volume, document that experience from five continents. It is also an activity that spans the full temporal range of Homo sapiens, extending from the present back over half a million years. Some finds hint at cave use by other members of the genus Homo, though the context and biological status of the early humans represented in some caves by their fossils or artifacts is rarely unambiguous. It is interesting to note that while cave use is a behavioural trait of many animals it does not appear to be one widely shared by humans' closest relatives, the other higher primates, though this may be because caves are not usually a feature of their preferred habitats. If we wish to compare human cave use with similar behaviour by other mammals we have to turn to the social carnivores, the wolves and hyaenas, both of which regularly use caves. Bears can provide another point of comparison though their use of caves is mainly as secure quarters for hibernation, which is not a habit shared by many humans. If the use of caves is not a uniquely human trait some aspects of that use certainly are. The papers presented in this volume describe a wide range of different activities undertaken in caves by humans over several hundred thousand years which fall broadly into two categories. Some activities may be described, rather loosely, as economic. These include residence, either short or long term, the acquisition of raw materials such as workable stone, minerals, water and chemicals, storage, and the disposal of waste. It is behaviour of this kind that humans have in common with many other animals. Wolves and hyaenas live in caves and dispose of their food debris there, elephants and antelopes have been reported entering caves in search of minerals and water, and one of the authors has excavated a squirrel's store of nuts in a cave in Scotland. It is the second broad aspect of cave use that is uniquely human, the use of caves for ritual purposes. Two activities can be described under this heading, the use of caves as theatres for ritual, usually evidenced by cave art and/or the presence of votive deposits, and the use of caves as burial vaults. Not all human remains found in caves are the result of formal burial and the victims of rockfalls can probably be classified along with the remains of bears that died in hibernation, while some human fragments should probably

be treated as the remains of another cave-using carnivore's meal. Formal burial is, of course, a form of ritual but not all rituals carried out in caves involved human remains. The distinction between economic and ritual behaviour may be partly false as it arises from the application of a twentieth century rationalist perspective which may not be entirely appropriate in other contexts. We know from ethnography, ethnohistory and everyday experience that many aspects of economic behaviour have a ritual dimension, while ritual behaviour can often have an economic aspect. However, as the papers published here show, most researchers concerned with the human use of caves find it convenient to summarize their data in terms of economic or ritual behaviour. The use to which a cave can be put depends on a number of factors of which two of the most important are its size and shape. A single burial can be squeezed into a small, dark crevice but residential activity requires space and light, while the more arcane of rituals seek obscurity and concealment. Several contributors draw attention to the way it is possible to classify caves according to their physical characteristics. Broadly speaking, on the one hand we have open, or daylit, caves with which we may include rockshelters, while on the other there are deep caves. The former can be, and frequently are, used for both economic and ritual activities, whereas deep caves are rarely used at all and then only for ritual purposes. Although open caves exhibit evidence for both economic and ritual use, this rarely appears to have been simultaneous and many papers document a pattern of change from one to the other. Virtually without exception the change is from economic to ritual use. It appears to be a worldwide phenomenon that open caves, initially used for economic purposes, eventually become the scenes of ritual activity that may or may not include human burial. The timing of this change is, of course, not synchronous in an absolute sense but does seem to coincide with other, fundamental changes of an economic and social kind. The most common example is the change from hunting and gathering to farming. The view that formerly mobile hunters and gatherers, on adopting a more sedentary food-producing mode of subsistence, nevertheless retain proprietorial, emotional and ideological ties with the more tangible aspects of their former existence is persuasive. In cases where people regarded as farmers are found engaged in the economic use of caves they are nearly always pastoralists

217

Tolan-Smith & Bonsall who are often nomadic. The gypsies, travellers and tinkers who have habitually used caves throughout Europe in recent centuries might be advanced as an exception to this rule, but these groups usually have more in common with huntergatherers or pastoralists than sedentary farmers, and they often maintain a degree of mobility in their lives in which natural shelters could be seen as a convenient alternative to frail temporary structures. The common factor that the economic use of caves by humans has is mobility. When people decide to adopt a sedentary mode of existence they usually come out of their caves and the use of these natural shelters is given over to ritual. Deep caves, without natural light, difficult of access, damp and often poorly ventilated, have rarely been the scene of economic activities, except in extreme circumstances of social unrest when they could offer a measure of security. The use of such deep caves is otherwise confined to ritual activity. The penetration of truly deep caves before the development of modem speleology, is reported from only three areas, Central America, Southwest Europe and the Urals. However, if we accept any cave without natural light into the 'deep' category, penetration and ritual use is as widely documented as economic use. In practice many caves are comprised of both 'open' and 'deep' components, and several authors document economic activity in cave entrances and ritual activity in deeper zones within the same complex. Observations of the latter kind emphasise one of the most important themes to emerge from a consideration of the papers presented here, the necessity of adopting a contextual approach to the study of human cave use. This applies at many scales of analysis, not just in the case of the juxtaposition of different activities. For example, our appreciation of the wonders of Upper Palaeolithic cave art is greatly enhanced by attempts to reconstruct the circumstances under which these images may have been viewed originally, taking account of light, shade and view point and acknowledging a potential role for other sensory stimuli such as sound. Taking a somewhat broader view, several contributors have emphasized that in the realm of ritual behaviour caves are not the only natural features imbued with special significance and many societies attribute ideological meaning to a range of natural features such as rocks, trees, streams, waterfalls and chasms. Ritual cave use

needs to be considered within this wider context if it is to be understood. Similarly, the economic use of caves did not take place within a vacuum and it is unlikely that there has ever existed a society of exclusively cave dwellers. On the contrary, where we have any evidence, cave use often seems to have been an exception and many cave dwellers appear to have spent at least part of their time using other kinds of shelters, which they built themselves. As more than one of our contributors point out, in economic terms, caves are simply 'convenient cavities', and their use must have articulated with activities in the wider world in general. It would also be inappropriate to ignore the social context of cave use in favour of the topographical and ecological. This dimension of human behaviour is notoriously difficult to appreciate in periods before the emergence of documented history and recorded ethnographic observation, though periods for which such sources do exist illustrate the range of possibilities. In economic terms, caves are usually regarded as fairly low status facilities, but exceptions to this rule can be found such as the cave dwelling Guanche 'noble women' of the Canary Islands. Study of votive deposits in Latvian caves during the post-Medieval period documents a clear correlation between increasing socio-economic stress and intensification of ritual behaviour. These examples offer provocative insights into human cave use regardless of the period under review. The papers brought together in this volume might at first sight be thought to run contrary to the principle of viewing human behaviour in context, in that they focus on one aspect, the use of caves. Individually, many of the contributors redress this apparent imbalance and review the results of their cave researches within a wider context. But the grouping of these papers here is deliberately intended to underline the commonality of one aspect of human experience. The fact that we can still visit caves used by our forbears over hundreds of thousands of years offers us the opportunity of sharing in some aspects of their experience, be it a sense of awe, a sense of safety or a sense of fear. We all know that caves are something special and we know of no society that does not have a special relationship with the caves in its area. There are few ways of feeling more human than to find oneself in a deep cave when the lights go out!

218