The Middle Palaeolithic Leaf Points of Europe: Ecology, Knowledge and Scale 9781407300672, 9781407331379

Ecology, Knowledge and Scale

167 74 19MB

English Pages [269] Year 2007

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

The Middle Palaeolithic Leaf Points of Europe: Ecology, Knowledge and Scale
 9781407300672, 9781407331379

Table of contents :
1663 verso.pdf
John and Erica Hedges Ltd.
British Archaeological Reports
1-12.pdf
01 Intro pages I-IX Hopkinson.pdf
ABSTRACT
Pollen diagram, 1937 Mauern core
Histogram showing comparison of the lithic industries from the Sesselfelsgrotte and Mauern
Leaf point fragment from the Sesselfelsgrotte G-Komplex
Pollen diagram, Watten
Pollen diagrams, Scladina
Pollen diagram, La Grande Pile XX
Pollen diagrams, Les Echets G
Pollen diagram, Ioannina 249
Pollen diagrams, Padul 2 and 3
Pollen diagram, Gołkow
Retouched tools from Trou de l’Abime. Couvin
Foliate bifaces from Spy
Foliate bifaces from Spy
Foliate bifaces from Spy
Foliate bifaces from Goyet
Bifaces from Grotte du Docteur
Bifacially retouched tool, Trou du Diable
Bifacially retouched tool, Grottes d’Engis
Bifaces from Engihoul
Biface from Grotte du Mont Falhize
Biface from Liège-Sainte-Walburge
Bifaces from Salzgitter-Lebenstedt
Bifacial leaf point, Ranis
Points from Ranis
Leaf points from surface finds in northern Germany
Leaf point fragment from Fritzlar
Leaf point fragments from Rörshain
Handaxes from Rörshain
Leaf points from Kösten
Bifacial leaf points from Mauern, Bohmers’ excavation
Leaf points from Mauern, Bohmers’ excavation
Bifaces from Mauern, Bohmers’ excavation
Bifacial leaf points from Mauern, Layer F of Zotz’s excavation
Bifacial leaf points from Mauern, Layer G of Zotz’s excavation
Possible Altmühlian leaf points from Mörnsheim and Biesenhard
Leaf points from Kleine Ofnet
Leaf points from Haldenstein
Handaxes from the Bocksteinschmiede
Bifaces from the Bocksteinschmiede
Leaf point-like bifaces from the Bocksteinschmiede
Leaf point fragments from the Klausen Caves
Leaf points from the Obernederhöhle
Bifaces from the Obernederhöhle
Hohlen Stein, Schambach
Leaf points from Zeitlarn
Leaf points and Blattformen from Albersdorf
Bifaces from Flintsbach-Hardt
Jerzmanowice points from Buchberghöhle and Pottenstein
Leaf point-like biface from Bohunice
Bifacial leaf points from Líšeň
Jerzmanowice points from Líšeň
Small handaxes from Kůlna
Leaf point-like bifaces from Kůlna
Jerzmanowice points from Kent’s Cavern
Bifacial leaf points from Neslovice
Bifacial leaf points from Vedrovice V
Jerzmanowice points, Nietoperzowa
Bifacial leaf points, Mamutowa
Zwierzyniec
Moravany-Dlha
Bifacial leaf points, Jankovich
Bifacial leaf points, Szeleta Cave Lower Industry
Bifaces, Subalyuk
Korolevo Va
Foliate bifaces from Bouzdoujany, Rikhta and Velikij Glubotchok
Foliate bifaces, Stinka I Lower
Leaf points, Stinka I Upper
Foliate bifaces, Zhitomir
Ripiceni-Izvor
Leaf points, Ripiceni-Izvor Level IV
Foliate bifaces, Ripiceni-Izvor Level V
Leaf points, Musselievo
Leaf points, Musselievo
Leaf points, Musselievo
Leaf points, Samuilitsa
Leaf points, Kokkinopilos
Leaf points, Morfi
Foliates, Starosele
Foliate handaxes, Starosele
Leaf points, Zaskalnaya VI
Biface knives, Zaskalnaya
Bifaces, Kabazi II
Jerzmanowice points and pointes à face plane, Kostienki Tel’manskaya
Bifaces, Mezmaiskaya
02 Ch1 1-10 Hopkinson.pdf
03 Ch2 11-30 Hopkinson.pdf
04 Ch3 31-46 Hopkinson.pdf
Germany
Poland
Rozumice C
Hungary
Vértesszöllös
Ukraine
Korol’evo VI, VII, VIII
05 Ch4 47-66 Hopkinson.pdf
Oxygen Isotope Stage
Kya
REGIONAL CHRONOLOGY
PALYNOLOGICAL SUCCESSION
NW Europe
France
European Russia
Greece
Padul 2
OIS 3
OIS 4
OIS 5b
OIS 5c
OIS 5d
OIS 5e
06 Ch5 67-90 Hopkinson.pdf
07 Ch6 91-112 Hopkinson.pdf
08 Ch7 113-126 Hopkinson.pdf
09 Biblio 127-158 Hopkinson.pdf
10 App1 159-176 Hopkinson.pdf
11 App2 177-246 Hopkinson.pdf
Figure A2.3. Bifacially retouched tools from Trou de l’Abime, Couvin. Pieces shown actual size.
Illustrations by T. Hopkinson.
Figure A2.48. Bifacial leaf points from Neslovice. Pieces shown actual size. (After Allsworth-Jones 1986, Fig. 41.1, 2.)
12 App3 247-253 Hopkinson.pdf
N. FRANCE
BELGIUM
N. & C. GERMANY
S. GERMANY
S. GERMANY CONT.
BOHEMIA
MORAVIA
S. POLAND
SLOVAKIA
HUNGARY
UKRAINE
ROMANIA
CRIMEA
S.W. RUSSIA
S. GERMANY
HUNGARY
ROMANIA
SERBIA
GREECE
CRIMEA
Sary-Kaya
Prolom II
S.W. RUSSIA
SITE
BOHEMIA
MORAVIA
Pod hradem
SLOVAKIA
HUNGARY
ROMANIA
BOSNIA
SITE
GREAT BRITAIN:
BELGIUM
MORAVIA
POLAND
HUNGARY
Balla
Puskasporos
S.W. RUSSIA
Kostienki-Streletskaya
Front Cover
Title Page
Copyright
Table of Contents
Abstract
Acknowledgements
List of Tables
List of Figures
Chapter 1: An Ecological Geography of the Lower and Middle Palaeolithic. the Leaf Points of the European Middle Palaeolithic and the Nature of the Problem
Chapter 2: Hierarchies of Scale and the Ecology of Knowledge
Chapter 3: Human Ecology and Scales of Knowledge in the European Early Palaeolithic
Chapter 4: The Distribution and Dating of Middle Palaeolithic Leaf Point Industries in Europe
Chapter 5: Upper Pleistocene Environment Change in the Regions of Europe
Chapter 6: The Middle Palaeolithic of The Altmuhl Valley
Chapter 7: Conclusions: The Scalar Convergence of Structure and Agency through the European Palaeolithic
Bibliography
Appendices

Citation preview

BAR S1663 2007 HOPKINSON: THE MIDDLE PALAEOLITHIC LEAF POINTS OF EUROPE

B A R

The Middle Palaeolithic Leaf Points of Europe Ecology, Knowledge and Scale

Terry Hopkinson

BAR International Series 1663 2007

The Middle Palaeolithic Leaf Points of Europe

The Middle Palaeolithic Leaf Points of Europe Ecology, Knowledge and Scale

Terry Hopkinson

BAR International Series 1663 2007

ISBN 9781407300672 paperback ISBN 9781407331379 e-format DOI https://doi.org/10.30861/9781407300672 A catalogue record for this book is available from the British Library

BAR

PUBLISHING

TABLE OF CONTENTS

ABSTRACT

III

ACKNOWLEDGEMENTS

V

LISTS OF TABLES AND FIGURES

VII

CHAPTER 1:

AN ECOLOGICAL GEOGRAPHY OF THE LOWER AND MIDDLE PALAEOLITHIC, THE LEAF POINTS OF THE EUROPEAN MIDDLE PALAEOLITHIC AND THE NATURE OF THE PROBLEM

1

CHAPTER 2:

HIERARCHIES OF SCALE AND THE ECOLOGY OF KNOWLEDGE

11

CHAPTER 3:

HUMAN ECOLOGY AND SCALES OF KNOWLEDGE IN THE EUROPEAN EARLY PALAEOLITHIC

31

CHAPTER 4:

THE DISTRIBUTION AND DATING OF MIDDLE PALAEOLITHIC LEAF POINT INDUSTRIES IN EUROPE

47

CHAPTER 5:

UPPER PLEISTOCENE ENVIRONMENT CHANGE IN THE REGIONS OF EUROPE

67

CHAPTER 6:

THE MIDDLE PALAEOLITHIC OF THE ALTMÜHL VALLEY

91

CHAPTER 7:

CONCLUSIONS: THE SCALAR CONVERGENCE OF STRUCTURE AND AGENCY THROUGH THE EUROPEAN PALAEOLITHIC

113

BIBLIOGRAPHY

127

APPENDIX 1:

POLLEN DIAGRAMS

159

APPENDIX 2:

ILLUSTRATIONS OF STONE ARTEFACTS

177

APPENDIX 3:

TABLES OF SITES

247

I

II

ABSTRACT

This work is an investigation of the relations between heterogeneity in the material world and variations in human behaviour, particularly landscape settlement and stone tool fabrication, in the European Lower and Middle Palaeolithic. A theoretical approach termed ecological geography is developed, central to which is a consideration of the question of scale. It is suggested that human knowledgeability be understood as an ecological entity that can be resolved into two scalar domains; knowledgeable action pertains to the purposeful individual while socially-transmitted knowledge pertains to trans-generational, collective dimensions of human existence. The scale domains on which these two aspects of knowledge operated in the European Palaeolithic are investigated in a review of the Lower and Middle Palaeolithic occupation of the regions of Europe and a consideration of rates of change in stone tool fabrication practices. The conclusion is reached that, through the Lower-Middle Palaeolithic transition, the temporal reach of knowledgeable action expanded while the social transmission of knowledge became more sensitive to ecological processes operative on time scales of a few tens of millennia. This approach is then applied to the Middle Palaeolithic leaf point industries of Europe. It is shown that biface-rich industries generally attributed to the Micoquian often feature a continuum of biface morphology, one pole of which approaches leaf point form, and that they frequently occur in contexts that demonstrate an association with relatively intensive occupations of social centres. On the other hand, assemblages in which leaf points occur as a morphologically discrete type occur only in the region from the central German uplands in the northwest to the Black Sea in the southeast, and in contexts that are datable to 60-40 kya and invariably indicate ephemeral occupation in landscape peripheries. Through an examination of the Upper Pleistocene environment history of the regions of Europe, it is shown that this spatiotemporal distribution corresponds to a zone which underwent pulses of afforestation and deforestation on wavelengths of a very few millennia in this period. Attention is then focused on the Middle Palaeolithic of the Altmühl Valley, Bavaria, and on two sites – Mauern, with a leaf point rich industry, and the Sesselfelsgrotte, a deep, stratified rock shelter site with a large Micoquian industry lacking leaf points. It is shown that the relevant industries at the two sites are both datable to the millennia after 60 kya, while a comparison of their lithic industries shows that they are technologically and typologically closely similar. It is therefore suggested that the two sites were connected in a single system of landscape use in which the Sesselfelsgrotte was a centre of social life and action while Mauern was occasionally visited in the course of action in the landscape periphery. The investigation, then, indicates that leaf points as a form were coupled with wavelengths of landscape transformation to which stone tool fabrication practices, and therefore the social transmission of knapping knowledge, were apparently insensitive before around 60 kya. At the same time, leaf points are associated with patterns of site use and distribution strongly indicative of a distinction between action at the social centre and in the periphery, suggesting perpetration of acts in the landscape in which other places and other times were implicit. This in turn implies a radical expansion in the temporal reach of knowledgeable action. It is concluded that, in the course of the European Lower and Middle Palaeolithic, socially transmitted knowledge and knowledgeable action converged upon each other in scalar terms, in the process of which received bodies of knowledge as to how to act became more vulnerable to transformation in transmission and context-specific technical innovations became more readily institutionalised. Pre-modern humans in Europe were therefore not trapped in permanent behavioural stasis. Possible factors in this process of convergence are discussed and the implications for our understanding of human ‘modernity’ considered.

III

IV

ACKNOWLEDGEMENTS This volume represents the realisation of my doctoral research at St John’s College Cambridge, as informed by my subsequent experience as Lecturer in Early Prehistory in the School of Archaeology and Ancient History, University of Leicester. Its realisation leaves me indebted to many people. Some played crucial roles in enabling me, at the age of 36, to change my life by becoming an undergraduate student of archaeology in Cambridge. Others have been my friends, guides and supporters since then, and some have been, and remain, important colleagues and discussants without whose wise and friendly criticisms my ideas would never have developed as they have. Indeed, so wide is my circle of indebtedness that it is not possible here to acknowledge the contribution made by each. To the following, my hearfelt thanks: It must suffice instead that I simply list their names here, in alphabetical order: Geoff Bailey, Graeme Barker, Huw Barton, Dimitri De Loecker, David Edwards, Duncan Garrow, Peter Goddard, Colin Haselgrove, Ray Jobling, Martin Jones, Preston Miracle, Nathan Schlanger, David Van Reybrouk, Wolfgang Weiβmüller, Mark White, Todd Whitelaw and Jane Woods. To any I have overlooked, my apologies. Special thanks are due, of course, to my PhD supervisor Paul Mellars, whose clarity of thought and insistence on rigorous argument provided me with a model to which I could, and continue, to aspire. His commitment to academic freedom, and his willingness to allow me to find me own way, were inspiring. Finally I must thank my family. My father Peter and sisters Carol and Tracey, my late parents-in-law Len and Margaret, and my sister-in-law Gillian have all been rocks without whom I would probably have foundered. Without their love and support I would not today be a student of the development of the human condition in the deep past. Most of all, my love and gratitude are due to my late mother Vera, whose spirit of self-educated sceptical inquiry forms the foundation of my own, much lesser, spirit; my son Stephen, whose mischievous ribbing has ensured that I never lost a proper perspective; and my wife Carol, whose love and loyalty – which I have sometimes severely tested - have remained unflinching. Without her devoted companionship this volume, indeed anything I count an achievement in my life, would have been impossible.

V

VI

LIST OF TABLES Table 1.1 Table 3.1 Table 5.1 Table 5.2 Table 6.1 Table 6.2 Table A3.1 Table A3.2 Table A3.3 Table A3.4

Technological and typological comparisons between the Middle Palaeolithic of western Europe and central and eastern Europe Sites older than 200 kya from East of the Rhine Results by site of quantitative analysis of OIS 5c3 pollen sequences Results by site of quantitative analysis of OIS 5a pollen sequences Analysis of retouched tools at the Sesselfelsgrotte. Analysis of retouched tools at the Weinberghöhlen, Mauern Sites with Middle Palaeolithic bifaces, some of which grade toward or into leaf point form Sites with unequivocal Middle Palaeolithic leaf points Sites with more than one leaf point but which might be either Middle or Upper Palaeolithic Sites with more than one Early Upper Palaeolithic leaf point

2 39 74 74 106 107 248-250 251 252 253

LIST OF FIGURES Figure 1.1 Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 3.1 Figure 3.2 Figure 3.3 Figure 3.4 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Figure 4.6 Figure 4.7 Figure 4.8 Figure 4.9 Figure 4.10 Figure 4.11 Figure 5.1 Figure 5.2 Figure 5.3 Figure 5.4 Figure 5.5 Figure 6.1 Figure 6.2

Distribution of Middle Palaeolithic leaf points in Europe Temperature variation at resolutions of 0.3s, 30s, 2h, 1 day, 1 month, 1 year, 4 years, 50 years, 3000 years Effect of response time on perceived temperature change in plant organs Input signals produce output signals of increasingly low amplitude as input frequency increases Braudel’s hierarchical conception of historical time The oxygen isotope curve from the GRIP Greenland ice core Distribution map of sites listed in Table 3.1 Lower Palaeolithic raw material transport patterns in the valley of the Garonne Late Middle Palaeolithic raw material transport patterns in the valley of the Garonne Schematic representation of an ‘ideal’ leaf point Schematic representation of a backed biface knife Distal pointedness Schematic representation of a unifacial leaf point Schematic representation of a Breitblattspitze Schematic representation of a Jerzmanowice point Symmetrisches schmales Faustkeilblatt Blattförmige Schaber Fäustel, Moravany-Dlha point Chronological correlations for the European Upper Pleistocene before 25 Kya Map showing locations of key Middle and early Upper Palaeolithic sites Locations of palynological sites The stacked marine oxygen isotope curve for the last 300,000 years Graphical plots of data given in Tables 5.1 and 5.2 Oxygen isotope and pollen records from Tyrrenhian Sea core KET 8003 Summary pollen diagrams of OIS 3 sequences from Ioannina 1, Xinias and Tenaghi Philippon The Altmühl Valley and surrounding region Evolution of the Danube

VII

3 12 13 14 29 33 39 41 44 48 48 48 49 49 49 51 51 52 54 55 68 69 75-76 85 88 92 93

Figure 6.3 Figure 6.4 Figure 6.5 Figure 6.6 Figure 6.7 Figure 6.8 Figure 6.9 Figure A1.1 Figure A1.2 Figure A1.3 Figure A1.4 Figure A1.5 Figure A1.6 Figure A1.7 Figure A1.8 Figure A1.9 Figure A1.10 Figure A1.11 Figure A1.12 Figure A1.13 Figure A1.14 Figure A1.15 Figure A2.1 Figure A2.2 Figure A2.3 Figure A2.4 Figure A2.5 Figure A2.6 Figure A2.7 Figure A2.8 Figure A2.9 Figure A2.10 Figure A2.11 Figure A2.12 Figure A2.13 Figure A2.14 Figure A2.15 Figure A2.16 Figure A2.17 Figure A2.18 Figure A2.19 Figure A2.20 Figure A2.21 Figure A2.22 Figure A2.23 Figure A2.24 Figure A2.25 Figure A2.26 Figure A2.27 Figure A2.28

Plan of and cross-section through the Sesselfelsgrotte Schematic composite stratigraphy of the Sesselfelsgrotte Plan of the Weinberghöhlen cave system Pollen diagram, 1967 core ‘Mauern 1’ Pollen diagram, 1937 Mauern core Histogram showing comparison of the lithic industries from the Sesselfelsgrotte and Mauern Leaf point fragment from the Sesselfelsgrotte G-Komplex Pollen diagram, Watten Pollen diagrams, Scladina Pollen diagram, La Grande Pile XX Pollen diagrams, Les Echets G Pollen diagram, Ioannina 249 Pollen diagrams, Padul 2 and 3 Pollen diagram, Gołkow Pollen diagram, Zgierz-Rudunki Pollen diagram, Kąty Pollen diagrams, Samerberg ‘Kernsbohrungprofil’ Pollen diagrams, Oerel OE61 Pollen diagram, Lac du Bouchet D Pollen diagram, Machnacz MII Pollen diagram, Kittlitz Pollen Diagram, Sulzberg Bifaces from Blanzy and Germolles Bifacially retouched pieces from Frettes Bifacially retouched tools from Trou de l’Abime, Couvin Retouched tools from Trou de l’Abime. Couvin Foliate bifaces from Spy Foliate bifaces from Spy Foliate bifaces from Spy Foliate bifaces from Goyet Bifaces from Grotte du Docteur Bifacially retouched tool, Trou du Diable Bifacially retouched tool, Grottes d’Engis Bifaces from Engihoul Biface from Grotte du Mont Falhize Biface from Liège-Sainte-Walburge Bifaces from Salzgitter-Lebenstedt Bifacial leaf point, Ranis Points from Ranis Leaf points from surface finds in northern Germany Leaf point fragment from Fritzlar Leaf point fragments from Rörshain Handaxes from Rörshain Leaf points from Kösten Bifacial leaf points from Mauern, Bohmers’ excavation Leaf points from Mauern, Bohmers’ excavation Bifaces from Mauern, Bohmers’ excavation Bifacial leaf points from Mauern, Layer F of Zotz’s excavation Bifacial leaf points from Mauern, Layer G of Zotz’s excavation Possible Altmühlian leaf points from Mörnsheim and Biesenhard

VIII

95 96 101 102 102 109 111 160 160 161 162-164 165 166 167 167 167 168-170 171-172 173-174 175 175 176 178 179 180 181 182 183 184 185 186 187 187 188 189 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203

Figure A2.29 Figure A2.30 Figure A2.31 Figure A2.32 Figure A2.33 Figure A2.34 Figure A2.35 Figure A2.36 Figure A2.37 Figure A2.38 Figure A2.39 Figure A2.40 Figure A2.41 Figure A2.42 Figure A2.43 Figure A2.44 Figure A2.45 Figure A2.46 Figure A2.47 Figure A2.48 Figure A2.49 Figure A2.50 Figure A2.51 Figure A2.52 Figure A2.53 Figure A2.54 Figure A2.55 Figure A2.56 Figure A2.57 Figure A2.58 Figure A2.59 Figure A2.60 Figure A2.61 Figure A2.62 Figure A2.63 Figure A2.64 Figure A2.65 Figure A2.66 Figure A2.67 Figure A2.68 Figure A2.69 Figure A2.70 Figure A2.71 Figure A2.72 Figure A2.73 Figure A2.74 Figure A2.75 Figure A2.76 Figure A2.77

Leaf points from Kleine Ofnet Leaf points from Haldenstein Handaxes from the Bocksteinschmiede Bifaces from the Bocksteinschmiede Leaf point-like bifaces from the Bocksteinschmiede Leaf point fragments from the Klausen Caves Leaf points from the Obernederhöhle Bifaces from the Obernederhöhle Hohlen Stein, Schambach Leaf points from Zeitlarn Leaf points and Blattformen from Albersdorf Bifaces from Flintsbach-Hardt Jerzmanowice points from Buchberghöhle and Pottenstein Leaf point-like biface from Bohunice Bifacial leaf points from Líšeň Jerzmanowice points from Líšeň Small handaxes from Kůlna Leaf point-like bifaces from Kůlna Jerzmanowice points from Kent’s Cavern Bifacial leaf points from Neslovice Bifacial leaf points from Vedrovice V Jerzmanowice points, Nietoperzowa Bifacial leaf points, Mamutowa Zwierzyniec Moravany-Dlha Bifacial leaf points, Jankovich Bifacial leaf points, Szeleta Cave Lower Industry Bifaces, Subalyuk Korolevo Va Foliate bifaces from Bouzdoujany, Rikhta and Velikij Glubotchok Foliate bifaces, Stinka I Lower Leaf points, Stinka I Upper Foliate bifaces, Zhitomir Ripiceni-Izvor Leaf points, Ripiceni-Izvor Level IV Foliate bifaces, Ripiceni-Izvor Level V Leaf points, Musselievo Leaf points, Musselievo Leaf points, Musselievo Leaf points, Samuilitsa Leaf points, Kokkinopilos Leaf points, Morfi Foliates, Starosele Foliate handaxes, Starosele Leaf points, Zaskalnaya VI Biface knives, Zaskalnaya Bifaces, Kabazi II Jerzmanowice points and pointes à face plane, Kostienki Tel’manskaya Bifaces, Mezmaiskaya

IX

203 204 204 205 206 207 208 209 210 211 212 213 214 215 215 216 217 217 218 219 219 220 221 222 223 224 225 226 227 228 229 230 230 231 232 233 234 235 236 237 238 238 239 240 241 242 243 244 245

X

CHAPTER 1

AN ECOLOGICAL GEOGRAPHY OF THE LOWER AND MIDDLE PALAEOLITHIC, THE LEAF POINTS OF THE EUROPEAN MIDDLE PALAEOLITHIC AND THE NATURE OF THE PROBLEM sufficient to affect the distribution of a way of life’ (McBurney 1950, 179). In presenting palaeolithic archaeology as an environmental-geographical discipline in terms that today could be regarded as broadly ecological, McBurney displayed considerable prescience, and it is the purpose of this study to re-examine McBurney’s observations in the light of contemporary ecological and social theory.

McBurney and the Geography of the Palaeolithic Almost sixty years have elapsed since the publication of Charles McBurney’s The Geographical Study of the Older Palaeolithic Stages in Europe (McBurney 1950), in which he brought into the public domain the essential elements of his doctoral thesis (McBurney 1948). The article is in one sense simply a clarification of geographical variability in European Lower and Middle Palaeolithic chipped stone industries, particularly with respect to the distribution of handaxes. McBurney sought, however, to transcend the mere cataloguing of the distribution of tool types and reduction techniques. The Geographical Study is, in its own diffident way, a call to arms, a plea for a new understanding of the Lower and Middle Palaeolithic. McBurney acknowledged the influence of geographical archaeologies being developed in the years immediately following the Second World War both in Britain and on the Continent of Europe, and embraced what he perceived to be their central principle – ‘the deliberate attempt to use [archaeological artefact] distributions as an indication of the part played by various natural and social factors in stimulating and moulding the cultural traditions’ (McBurney 1950, 163). He also recognised that later prehistory afforded the best conditions for geographical archaeologies motivated by this principle, but rejected the idea that the Lower and Middle Palaeolithic record defied such an approach. The Geographical Study was therefore an attempt to promote, and to assess the viability of, a geographical archaeology of the European Lower and Middle Palaeolithic in which tool industrial variability was to be understood as rooted in a relationship between culture, topography and environmental variation in time and space. It contains a number of key observations which, if substantiated, would form important pillars of a geography of the European Lower and Middle Palaeolithic:

Middle Palaeolithic Leaf Points as a Geographical Problem One of McBurney’s concerns in The Geographical Study was the apparent regionalisation of lithic industries in the Middle Palaeolithic. He described an opposition between a western Mousterian restricted essentially to southwestern Europe, but with outliers in Belgium and possibly the western part of the north German plain, and an eastern Mousterian found in a huge area extending from the Ardennes and the Rhine to the Urals. This geographical division he interpreted in terms of a ‘favoured’ region centred on western France, from which innovations diffused north and northeast, and a more conservative region to the east characterised by upland regions and continental climates. The German uplands constituted a region subject to cultural influence from both east and west, but whose primary role was as a cultural boundary. Of more interest here are the gross characteristics of McBurney’s two great regional variants of the Mousterian. He recognised three facies of the western Mousterian: one with broad, faceted-platform flakes struck from discoidal cores and frequently secondarily retouched on the dorsal surface of their edges, together with a cruder flake component struck from simpler cores; the second is identical except for the presence of cordiform handaxes; the third he described as showing a higher standard of retouch finish, thinner flakes, a tendency to ventral thinning and an absolute or virtual absence of handaxes. All facies consist of moderate sized tools and are dominated by discoid reduction. The eastern Mousterian in McBurney’s scheme shares with the western variant an emphasis on discoidal techniques and the production of scrapers but is distinguished from it by a tendency to microlithic tools, the use of small pebbles as raw materials, the occurrence of thick points and the frequent intensive application of flat bifacial retouch.

1. Lower Palaeolithic sites are rare in Europe east of Germany and north of the Mediterranean region; 2. The systematic occupation of both plains and upland regions in the Early Palaeolithic is to be thought of as a quite different mode of life from the settlement of plains regions alone; 3. Climate in Europe becomes more continental, and thus more seasonal, as one moves east, and this ‘would be

1

McBurney’s student Rolland has rightly drawn attention to deficiencies in this opposition (Rolland 1990). Inadequate attention was paid to the reliability of collections from 19th and early 20th Century excavations and to the problems of stratigraphic provenance, and McBurney employed an idiosyncratic typological classification. Puzzlingly, he did not recognise the importance of Levallois reduction in many assemblages of the French cave Mousterian and in a number of eastern Mousterian assemblages. Neither did he recognise the presence of a range of biface forms in the Middle Palaeolithic of central and eastern Europe, including handaxes, bifacial knives and leaf points, an omission that can only be explained in small part by the state of knowledge in 1937-1939, when he collected his observations. Nevertheless, his identification of bifacial flat surface retouch as a characteristic of the central and eastern European Middle Palaeolithic has stood up well to 50 years of excavation and assemblage analysis, and this technological feature is highly typical of the Middle Palaeolithic leaf points of that region.

The interest of leaf points from the point of view of a geographical archaeology is not, however, limited to their distribution in space. Both Klein (1999, 424) and Mellars (1996,135) refer to the degree of apparent formal design that these pieces show and contrast them on that basis with the great majority of Middle Palaeolithic stone artefacts. They have also been central to debates over the Middle-Upper Palaeolithic transition in central and eastern Europe (e.g. Allsworth-Jones 1986, 1990; Gladilin and Demidenko 1990; Freericks 1995; Adams 1998). Leaf points therefore offer an opportunity to examine, within a wider consideration of the geography of the Lower and Middle Palaeolithic, just the kind of relationships between environment, geography and specific social-cultural behaviour that McBurney envisaged.

Rolland has outlined a series of typological and technological comparisons between the Middle Palaeolithic stone tool industries of the regions of Europe which is more robust than McBurney’s in the light of our present state of knowledge. These are shown in Table 1.1. In fact, Rolland’s own comparisons are open to challenge. Levallois reduction is more common in the ‘eastern Mousterian’ than he suggests, for example. But the strong geographical zonation of a wide range of biface forms and bifacially retouched pieces, including leaf points, is clear and widely recognised in the literature (e.g Société des Amis du Musée National de Préhistoire et de la Recherche Archéologique 1995). The same concentration of Middle Palaeolithic leaf points in the central and eastern regions of Europe and their absence from the west of the continent has also been affirmed by Bordes (1968, 1984) and by Kozlowski (1995), who recognises cultural sub-divisions within the distribution (see Fig. 1.1).

Quite the most thorough and valuable historical account of the history of the archaeological study of Middle Palaeolithic leaf points is that given by Allsworth-Jones (1986, Chapter 1) in his analysis of the Szeletian, and his work need not be repeated in detail here. In short, it is fair to say that the study of leaf points has historically been concerned with the construction or refutation of timespace industrial units or cultures and their phylogenetic relationships (e.g. Freund 1952; Valoch 1968a; Allsworth-Jones 1986). The first discovery of leaf points in Hungary was taken to show that there had been Solutrean occupation in that country (Kadić 1916), but the subsequent discovery of ‘archaic’ leaf point industries in Poland, Slovakia and Moravia challenged this cultural association. Obermaier and Wernert (1914) identified the leaf points they had excavated from the Klausen Caves in the Lower Altmühl Valley, Bavaria, as Middle Palaeolithic and derived them from Acheulean antecedents, introducing thereby the concept of an ‘archaic’ biface culture-industrial ‘phylum’

A Brief History of Leaf Points.

Western Mousterian Technology Discoidal Levallois Pebble cores Planoconvex flat surface retouch intensive unifacial retouch intensive bifacial retouch microlithisation

Eastern Mousterian

2 2 2

Typology Scrapers Limaces Prondniks Leaf points Cordiform bifaces

2 1 2

2

Other bifaces

-

-

Notched pieces

2

-

Technology Discoidal Levallois Pebble cores Planoconvex flat surface retouch intensive unifacial retouch intensive bifacial retouch microlithisation

3

Typology Scrapers Limaces Prondniks Leaf points Cordiform bifaces

3 1 2 2 -

3

Other bifaces

4

4

Notched pieces

-

3 2

3

Table 1.1: Technological and typological comparisons between the Middle Palaeolithic of western Europe (‘Western Mousterian’) and central and eastern Europe (‘Eastern Mousterian’). Modified after Rolland 1990, Tab. 3. - absent or virtually so; 1 regularly present in low frequencies; 2 variable frequency; 3 common; 4 dominant.

2

encompassing the Acheulean, the Klausen material and the Solutrean in a single evolutionary trajectory. A variation on this viewpoint continued to be articulated and can be thought of as postulating an eastern after World War II by Zotz (1955) and Valoch (1968a), ‘Mousterian of Acheulean Tradition’.

‘Altmühlian’ (Bohmers 1939, 1944, 1951); others interpreted as ‘transitional’ or early Upper Palaeolithic in character, such as the material from Ranis, have been placed in a ‘Ranisian’ culture (Hülle 1977) and tentatively linked with the Jerzmanowician and with a British ‘Lincombian’ leaf point culture (Campbell 1980; Kozlowski 1995).

On the other hand Freund (1952) rejected the idea that leaf points, despite their apparent formal design (whether real or imagined), could be utilised as ‘cultural markers’ or type fossils, arguing instead that they could occur in any cultural context. Nevertheless she subsumed all of the leaf point industries of Europe east of the Rhine in a broad ‘Prae-Solutréen’ cultural grouping. This was not widely accepted, and since 1952 the trend has been towards the construction of chronospatially discrete leaf point cultures. Most of these are thought of as ‘transitional’ between the Middle and Upper Palaeolithic, the most important being Prošek’s ‘Szeletian’ leaf point industries of central and eastern Europe other than Germany (Prošek 1953). Arguments persist as to whether this term can usefully describe ‘transitional’ leaf point industries outside the Bükk Mountains, leading to the naming of industries such as the ‘Jankovichian’ of western Hungary (Gábori-Csánk 1956, 1974), the ‘Jerzmanowician’ of southern Poland (Chmielewski 1961) and the ‘Bohunician’ of Moravia (Oliva 1979). Certain leaf point industries from Romania and Bulgaria were also given the Szeletian label (Nicolăescu-Plopşor 1957; Dzambazov 1967) although they are now regarded as unambiguously Middle Palaeolithic in character. Further west, Middle Palaeolithic leaf point industries in Germany have been assigned to the Micoquian (Bosinski 1967) and, in the Upper Danube, to Bohmers’

The evolutionary relationship between the Middle and Upper Palaeolithic, and between Neanderthals and Modern Humans, has been a recurrent theme in the culture history of leaf points, and interest in their role in those transitions has increased in recent years (e.g. Allsworth-Jones 1986, 1990; Gladilin and Demidenko 1990; Freericks 1995; Adams 1998). Increasing use has also been made of metrical analyses of leaf point assemblages for the purpose of establishing patterns of formal similarity and dissimilarity, and thus of putative evolutionary cultural affinity (e.g. Sirakov 1979; Haesaerts and Sirakova 1979; Kozlowski 1990b). Explicitly technological analyses have been rare (but see Hahn 1990 for just such an analysis of the Middle Palaeolithic leaf points from Rörshain, Germany). Concern to infer past systems of landscape use in which leaf points were implicated has not, however, been entirely absent. Hillebrand (1917) interpreted the Bükk Mountain sites with crude bifaces, such as Puskasporos, as Developed Solutrean workshop sites, a view later endorsed by Vértes (1968) who differed only in substituting ‘Szeletian’ for ‘Solutrean’. Neustupný and Neustupný (1960), on the other hand, concluded that the Bükk Mountain Szeletian was a functional-seasonal

Fig. 1.1: Distribution of Middle Palaeolithic leaf points in Europe according to Kozlowski (1995, Fig. 2). 1, Balkan LevalloisoMousterian; 2, Central European Micoquian; 3, Eastern Micoquian; 4, Altmühlian; 5, Stinka Denticulate Mousterian.

3

facies of the Aurignacian, a viewpoint recently adopted by Adams (1998). Klíma (1957) complained that no leaf point site in Moravia had been sufficiently thoroughly excavated to cast any light at all on the settlement patterns or organisation of their makers. Müller-Beck, echoing McBurney, characterised ‘Micoquoid’ industries as culturally conservative and restricted to mountainous areas (Müller-Beck 1966).

variability in the European Lower and Middle Palaeolithic, but have not primarily been concerned with understanding chipped stone technology as an aspect of human ecology and geography in the Pleistocene. This apparent disregard for McBurney’s perspective is, perhaps, unsurprising. The Geographical Study is opaquely written, does not present McBurney’s data in a succinct manner, poses questions rather than answers, presents no worked-out theory of human ecology and action in time and space, and identifies no spatial unit intermediate between the site and the continent that could function as a meaningful unit of analysis. Understanding of Quaternary chronology, climate and environment history in Europe was at a rudimentary stage of development in the 1950s and, by comparison with the masses of lithic data available in the Palaeolithic record, afforded little possibility of detailed analysis. Perhaps most important is the weakness in the post-war years of a pan-European grasp of the Palaeolithic. The geographical breadth of interest shown by figures such as Breuil, Obermaier and Garrod was lost; with few exceptions, archaeologists operated within national traditions sundered from each other by differences of approach and language and by the Iron Curtain, and this militated against the emergence of a genuinely European scale of analysis.

The most explicitly geographical (in the sense that McBurney understood the term) treatment of Middle Palaeolithic leaf points is Kozlowski’s recent synthesis (Kozlowski 1995). He argues that leaf points emerged independently in different cultural contexts, but always immediately before a pleniglacial maximum; and that leaf points thus reflect a complex of factors, including raw material availability, environmental stress, leaf point function in relation to the function of other tool types, and the social organisation of lithic production. The argument can be criticised on several grounds, not least of which is that Kozlowski advances no mechanism whereby knappers might anticipate a future pleniglacial. Nevertheless Kozlowski is clearly moving towards a behavioural interpretation in which changes in the environment, and social responses to them, are recognised as central. It is also the only work, as far as the writer is aware, in which leaf points are interpreted in terms of cyclical environmental change.

The Cambridge School of Palaeoeconomy. Although McBurney’s call for a geographical archaeology of the Lower and Middle Palaeolithic had very limited impact, the trend towards such an approach gathered speed amongst British archaeologists of the Upper Palaeolithic and Mesolithic in the years following 1950. The central figure in this development was McBurney’s Cambridge contemporary and colleague, J.G.D. Clark, whose influence McBurney grudgingly acknowledged in The Geographical Stages by mentioning, without citation, ‘some of the results suggested by recent geographical study of the Mesolithic’ (McBurney 1950, 163). Clark founded a long-lived research programme, centred on the Cambridge Department of Archaeology, which came to be known as the Cambridge School of Palaeoeconomy. This was not, and is not, a unified or homogeneous movement. It underwent considerable development through the involvement of numerous researchers, the most notable of which from a Palaeolithic perspective are Eric Higgs, Geoff Bailey and Clive Gamble. Through their work, the theoretical and methodological principles of a geographical palaeolithic archaeology have been formulated, challenged and developed. A brief examination of the ideas of the Cambridge School, and their continuing development, will therefore form the basis for constructing the broad outlines of a route towards the realisation of McBurney’s ambitions.

Geographical Archaeologies of the Palaeolithic and Mesolithic since 1950. The Geographical Study does not seem to have set McBurney’s contemporaries alight. At this time Movius was also pondering the distribution of handaxes (Movius 1948, 1949, 1955) and was drawing his ‘Line’ between an east and southeast Asian handaxe-free Early Palaeolithic and a handaxe-rich West. Despite the obvious close parallels with the problems McBurney was seeking to address, Movius does not seem to have taken McBurney’s advice; he regarded the claimed absence of handaxes in southeast Asia not primarily as an ecological phenomenon, but as a ‘monotonous and unimaginative’ consequence of cultural conservatism. More importantly, in the same year that The Geographical Study appeared François Bordes published his celebrated typological system (Bordes 1950a, 1953a) and gave new impetus to studies based on typological comparisons between Palaeolithic assemblages and aimed at the construction of industrial genealogies (e.g. Bordes 1950b; Freund 1952; Müller-Beck 1957; Bosinski 1967; Valoch 1968a). Criticisms of the Bordian method and of Bordes’ own interpretations of its results (Binford and Binford 1966; Binford 1973; Mellars 1969; Dibble 1983, 1984, 1988; Dibble and Rolland 1992) have certainly made important contributions to our understanding of industrial

4

In his influential excavations and subsequent interpretations of the Mesolithic site of Star Carr in North Yorkshire (Clark 1954, 1972) Clark showed what an economic archaeology could achieve. The excavation and analytical techniques applied by Clark and his coworkers, particularly the faunal analysis (Fraser and King 1954), underwrote an interpretation of the site as having been occupied in winter and spring as part of a mobile hunter-gatherer regional settlement system that integrated the lowlands of the Vale of Pickering and the hill country of the North Yorkshire Moors through a seasonal round of movement through the landscape. Clark therefore identified a spatial unit of analysis between the site and the continent on which a geographical archaeology could be based. The particulars of this interpretation have since been challenged (e.g. Legge and Rowley-Conwy 1989), but Clark’s approach continues to influence archaeologists of earlier prehistory throughout the Anglophone world. His realisation that archaeological sites need to be understood as traces of past activities connected to other activities carried out at other times and in other parts of the wider environment formed the glue that held - and still holds – the palaeoeconomic outlook together as a research orientation. It also introduced the critical element of scale into archaeological thinking, in that prehistoric ways of life were conceived as adapting to geographical and ecological factors at more than one level. Both long term, continental scale variations such as the glacial-post glacial transition and the gross latitudinal zonation of plant communities, and annual-seasonal variations in spatially heterogeneous environments at the local, sub-regional and regional levels, had to be taken into account by archaeologists.

J.G.D. Clark Clark’s Prehistoric Europe: The Economic Basis (Clark 1952) presented an economic history of Europe between the late glacial and the rise of Rome. His exclusion of the Palaeolithic older than the late glacial was not simply a matter of convention or convenience. He stated quite explicitly that human societies, as opposed to mere technical traditions, are unrecognisable before the Upper Palaeolithic and that the late glacial therefore represented the only workable point of departure for his project. In the first chapter, Clark laid out succinctly what he believed the fundamentals of his approach, which he termed economic rather than geographical, to be: [Economy is] an adjustment to specific physical and biological conditions of certain needs, capacities, aspirations and values. There are two sides to the equation – on the one hand the character of the habitat, itself to a greater or less degree influenced or even conditioned by culture, and on the other the kind of life regarded as appropriate by the community and the resources, in the form of knowledge, technical equipment and social organization, available for its realisation. Clark 1952, 7

McBurney’s ‘social factors’ were here fleshed out to include technology and systems of social relation and organisation. More interestingly, Clark also recognised two crucial points – the centrality of socially mediated knowledge, values and perceptions in the generation of ways of life, and the reciprocal relation between these and the inhabited environment. Clark therefore explicitly raised the question of past human social consciousness to a position of the highest importance in the construction of archaeological economic histories. He also went on to outline a conception of the terms ‘habitat’ and ‘environment’ that incorporated variability in time and space. He saw events in climate history, particularly the transition from the late glacial to the post glacial, as external to sociality but as impacting upon the reciprocal relation between society and nature through the changes they effect in the opportunities available to human beings. At the same time geographical factors such as topographical relief and latitude create ecological zones within the European continent. Clark identified the northsouth vegetational transitions from tundra, through coniferous boreal forest and temperate deciduous forest to Mediterranean evergreen woodland, as being of particular importance. Events, patterns and processes of this character, together with environmental changes caused by human action and social factors such as population growth and the establishment of long-distance contacts between cultures, set the boundary conditions for the unfolding of economic history. Clark therefore conceived economy in prehistory as consisting in the relationships between ecology, geography and sociology in space and time.

Ethnographic analogy was also an important element in Clark’s economic archaeology. In many respects, Clark's interpretation of Star Carr owed a debt to EvansPritchard’s classic ethnographic study of the Nuer (Evans-Pritchard 1940), which is cited approvingly in Prehistoric Europe. Clark drew particularly and explicitly on Evans-Pritchard's description of the Nuer as in ecological equilibrium with their environment, and asserted that European societies in prehistory would have achieved similar states of perfect adjustment to their environments. Ways of life could therefore persist for as long as the natural and social conditions on which they were predicated remained unchanged. This permitted Clark to place Star Carr in a past ‘ethnographic present’ in which history was stilled and a fixed strategy for exploitation of the environment could be described. Beyond this the intellectual structure of Evans-Pritchard’s analysis, in which the ecology of the Nuer is presented in terms of an adaptation to resource availability through a system of annual movements between dry season and wet season locations, is implicitly present in Clark's work. Indeed, it could be said that, through Clark, British structural-functionalist social anthropology has been a major element in Anglophone archaeologies of the Palaeolithic and Mesolithic since the Second World War.

5

Eric Higgs

rejection of the importance of temporal scale. In the same paper, social and ecological dimensions of human adaptation were rejected as insignificant, along with the New Archaeology and systems perspectives. Palaeoeconomy was presented as an extension of animal behaviour studies, and human adaptation as governed in a predictable manner by overriding principles or natural laws that boiled down to the time-distance factors involved in exploiting a spatiotemporally uneven resource distribution. This startling narrowness of outlook is in stark contrast to Clark’s broad concern with the inter-relations between ecology, geography, social organisation, cultural knowledge and human action, and explains Higgs’ unhelpful focus on subsistence and the food quest as the only aspect of prehistoric life worth investigating.

The economic perspective on prehistory founded by Clark was developed methodologically, but diminished theoretically, by the palaeoeconomy of Eric Higgs. In the 1960s Higgs undertook field work in the Levant, through which he sought to bring the economic approach to bear on the question of the early development of agriculture in the Natufian. He developed a conception of sites as loci in a spatially heterogeneous array of resources, and that the particular sets of resources to which occupants of any particular site would have had access was a function of distance, and therefore walking time, from the site. This concept he called site catchment, and the methodology whereby archaeologists might determine the resource base afforded by a site’s location was termed site catchment analysis (Vita-Finzi, Higgs et al 1970). The role of spatial scale, relative to the ideas of Clark, was refined in site catchment analysis in that a site was perceived to be at the centre of a series of concentric time-distance ‘circles’. Through walking the landscape, archaeologists could delineate site catchment areas defined by one hour and two hour radii. This, combined with palaeoenvironmental records and the animal and plant residues recovered archaeologically from a site, could illuminate the role of the site in the past subsistence economy. The presence of food residues or other archaeological objects demonstrably derived from outside the catchment area would indicate the articulation of the site with others in a regional annual exploitation territory (Higgs and Vita-Finzi 1972; Jarman 1972).

Geoff Bailey Some of the problems with Higgs’ methods and outlook were addressed by a number of archaeologists who had been his students, and who contributed the bulk of papers to a ‘state of the art’ volume, Hunter-Gatherer Economy in Prehistory, edited by one of their number, Geoff Bailey (Bailey 1983a). In the introductory article Bailey (1983b) reaffirmed the central role of subsistence in the adaptation of hunter-gatherer ways of life to environmental constraints, but also moved to challenge elements of the Higgsian orthodoxy. In particular, he criticised the notion that all non-agricultural modes of human adaptation could be understood through a single set of generalised overriding principles good for all places and all times. This meant that the application of the uniformitarian thinking typical of Higgs was an inadequate approach to the study of prehistoric behaviour. Behaviour in the past might, indeed must, have been different:

The need to link the Levantine late Pleistocene chronology with that of Europe led Higgs to initiate a field project in Epirus, northwestern Greece, which succeeded in locating a number of rock shelter and open air sites with Middle and Upper Palaeolithic archaeology (Dakaris et al 1964; Higgs 1965; Higgs and Vita-Finzi 1966; Higgs et al 1967). Site catchment analysis was subsequently applied to the Upper Palaeolithic of the region and an economic interpretation developed in which the Asprochaliko and Kastritsa rock shelter sites were identified as complementary home bases in a wider regional exploitation system stretching from the coast to the high Pindus Mountains (Vita-Finzi 1978). No similar interpretation of the Middle Palaeolithic settlement system in the same area was presented.

[It is] the behavioural potential of the archaeological record which forms the historical bridge between our animal ancestry and our present-day behaviour. Bailey 1983b, 4.

The metaphor of a ‘bridge’ between animal and modern human behaviour necessarily implied the existence of an evolutionary-historical span in which human behaviour and the factors governing it were different from anything observable today, but the archaeological record bridges – i.e. affords access to – that behavioural space. Bailey therefore implicitly recognised the applicability to the Lower and Middle Palaeolithic of the geographical principles of action in time and space developed by the palaeoeconomic perspective. In articulating this idea, he drew a line between himself and his generation of palaeoeconomic archaeologists on the one hand, and Clark and Higgs on the other.

The site catchment method represented a major leap in the operationalisation of economic and geographical archaeology. However, it shared with Clark’s appeal to adaptive equilibrium a reliance on timeless synchronic segments of the past in which reified exploitation systems persisted in stasis. This was not an oversight. In a theoretical exposition of the principles of palaeoeconomy (Higgs and Jarman 1975), Higgs dismissed the ‘short term trivia’ of the archaeological record and declared that long term patterning alone was worthy of archaeological study. The sense of spatial scale evident in Higgs’ outlook therefore went hand in hand with an unfortunate

The basis of Bailey’s optimism was his grasp of the integration in behaviour of multiple processes operating on different wavelengths or periodicities, and his

6

conviction that a multiscalar approach – which he termed ‘time perspectivism’ - could close not only the gap in the archaeological imagination between modern and archaic humans, but also the intellectual gap between biological and sociological approaches to human behaviour (Bailey 1983c, 2007). This latter point was made in the context of the emergence in Cambridge of what is now called postprocessual archaeology, with its distrust of functionalism and its call for subjective human consciousness to be placed at the top of the archaeological agenda (e.g. Hodder 1982a,b,c). Clark had at least recognised this dimension of human experience, as has been noted. But Higgs’ environmental determinism, in common with the American New Archaeology (Binford 1962, 1965; Watson, LeBlanc and Redman 1971) and, to a lesser extent, British social or ‘Processual’ archaeology (e.g. Renfrew 1973), had quite deliberately relegated it to a position of insignificance. Almost 20 years on, this reconciliation of the scientific and humanistic conceptions of humanity, regarded by both McBurney and Clark as essential features of a geographical archaeology, remains only dimly visible in the discipline. It is the conviction of this writer that Bailey was correct in identifying multiscalarity as the key to its achievement.

the centrality of the food quest and mobility in huntergatherer adaptation to the environment. The spatial structure of the environment was also identified as a key factor. However, unlike Higgs, Gamble embraced ecology both as a key factor in determining the character of human ways of life and as a theoretical framework for the interpretation of the Palaeolithic record. He recruited the ideas that had emerged from ethnographic and ethnoarchaeological studies of contemporary huntergatherers in the course of the 1960s and 70s, and particularly the work of American ecologically-oriented cultural anthropologists, to provide bridging arguments, while recognising the limitations of ethnographic analogy and the uniformitarian principle. At the same time, Gamble was, like Bailey, clearly dissatisfied with Higgs’ disregard for the role played by sociality in Palaeolithic adaptation to environmental constraints, and sought to resolve this problem by identifying the distinct roles played by environment and sociality. The environment was to be thought of, in his view, as determinant in that it sets out what is available for exploitation. Social relations, however, are dominant, specifying how the environment is to be exploited (Ingold 1980). The ecological and spatial character of the environment might impose upon hunter-gatherers certain behavioural strategies in terms of group size, mobility, length of residence at particular sites and resource targeting, for example; but it does so only in the context of a hunter-gatherer mode of social organisation which imposes mobility and small group size. An agricultural society would exploit the same environment quite differently by virtue of their particular mode of the social organisation of economic life.

Bailey carried forward Higgs’ work in northwestern Greece for some twenty years and has extended the network of known sites in the later Upper Palaeolithic occupation of the region. Clark’s concern with the relations between society, culture, ecology and geography form the intellectual framework of his analyses (e.g. Bailey 1997, 1999). His interpretations of the Upper Palaeolithic of Cantabria (Bailey 1983d) have also addressed the problem of change through time in settlement and subsistence behaviour, and have therefore broken Clark’s and Higgs’ reliance on the construction of past ethnographic presents. It would be fair to say that he has remained a specialist in the archaeology of Upper Palaeolithic, Mesolithic and early agricultural societies, and has not followed his own call to extend geographical archaeology into the Lower and Middle Palaeolithic. He has also eschewed the use of interpretive frameworks derived from contemporary social theory and postprocessual archaeologies, preferring to retain a fundamentally environmental-ecological perspective.

Gamble can, then, be said to have been seeking to develop Clark’s notion of the reciprocal relations between environment and society in The Palaeolithic Settlement. His observation that ‘as the environment reproduces itself on a seasonal, annual and longer term cycle, so too does the hunter-gatherer society that is dependent upon its products’ (Gamble 1986, 30) indicates a sensitivity to the multiscalar character of connections between society and environment, a position close to that of Bailey. Unfortunately, Gamble offered little clue as to which aspects or scales of human sociality might be coupled with particular scales of environmental rhythms, although he did recognise that the different levels of huntergatherer social unit – family, local group and regional population – might themselves function at different spatial scales.

Clive Gamble In the last twenty years Clive Gamble, another former Higgs student, has produced two very different synthetic geographical archaeologies of the European Palaeolithic. The first of these, The Palaeolithic Settlement of Europe (Gamble 1986) can be said to be responsible for the emergence of a geographical archaeology of the European Lower and Middle Palaeolithic as a serious research programme. In many respects, the theoretical position Gamble adopted in The Palaeolithic Settlement was squarely in the tradition of Higgs. He emphasised the unique insight the Palaeolithic record affords into the long-term processes of human behavioural history, and

Gamble’s purpose in The Palaeolithic Settlement was to trace the course of development in social responsiveness to environmental change through the European Palaeolithic. To achieve this he followed Clark in dividing the continent into gross ecological zones, although he added longitudinal divisions to Clark’s latitudinal zones to define nine geographical units or regions of analysis (Gamble 1986, Fig. 3.1). The history

7

of Palaeolithic settlement in each region through episodes of climatic and environmental change could then be used as an index of social-environmental adaptability, and, by proxy, of past social organisation. Gamble’s approach at that stage of his career was therefore explicitly social, geographical and ecological, and designed to bring the Lower and Middle Palaeolithic out of the cold and into a unified, pan-European archaeology of the Palaeolithic.

and Maryanski 1991) is based upon the idea that individuals are creative, mobile nodes who form mutual ties. In consequence individuals are always immersed in networks of relations with others. However, interpersonal ties vary in their intensity, commitment and duration. Multiplex ties, for example, are those that link individuals in multiple capacities simultaneously. In a peasant household, for example, one’s brother might also be one’s workmate, family head, marriage broker etc. Uniplex ties are unidimensional – between the shopper and the check-out assistant, for example. At the same time, people marry, leave, disagree, have children, and a host of other things that constantly terminate or change the character of existing ties and create new ones. Individuals are therefore always actively immersed in fluid networks of variable relations with others, and these networks provide the social contexts in which action, which always has a social goal, is performed.

However, his summary of the Palaeolithic record of Europe revealed no social-organisational trends at all in the Lower and Middle Palaeolithic, for two reasons. Firstly, by ‘social organisation’ Gamble meant reified, structured social systems through which energy, nutrients and information flowed like electricity through a circuit. Once one adopts this position then social organisation is archaeologically visible only insofar as past information flow was routed through artefacts as coded signs and symbols, evidence for which before the Upper Palaeolithic is vanishingly rare. Secondly, despite his comments on the multiscalarity of social reproduction in environmental context, The Palaeolithic Settlement makes no attempt to link human occupation with environmental and climatic change on any scale other than that of glacial-interglacial cycles. This is not simply a question of resolution in the record; it is a question of the theorisation of the integration in behaviour of processes operative on multiple wavelengths. Without this, The Palaeolithic Settlement becomes, in the end, a narrowly ecological work.

The varying strength and durability of interpersonal bonds underwrites a hierarchy of social networks distinguished by size, by the intensity of the relationships by which they are constituted, and by the nature of the resources brought to bear on their construction and maintenance. Intimate networks are formed through frequent, face to face interactions between small numbers of individuals – typically fewer than 10 - whose lives and routines of action are closely bound together in relations of strong and long-lasting mutual support and obligation and are constructed and maintained primarily through their frequent reaffirmation in face-to-face emotional and familiar contexts. Effective networks, which generally consist of around 20 individuals, incorporate the persons in whose presence each individual pursues their economic, reproductive and social ends in the encountered contexts of everyday life. Relations between individuals in effective networks are close and important but not emotionally unconditional, and are created and maintained largely through the use of material resources such as food sharing and cooperation in the performance of tasks. Individuals also maintain relations with acquaintances, friends-of-friends and strangers. Although there is considerable variation, an individual’s extended network might include 100-1000 others. Since the relations between individuals within extended networks are often categorical rather than personal, uniplex rather than multiplex and only rarely affirmed by face to face encounters, they are generally maintained by symbolic resources, including material culture, which allow the persons to be presenced even in their absence, thereby creating social relationships that transcend the limits of time, space and the body. Orders of sociality are therefore linked with orders of spatiotemporality.

Since The Palaeolithic Settlement much of Gamble’s work has centred on the search for a route to sociality in the Lower and Middle Palaeolithic (e.g. Gamble 1993, 1995b, 1996, 1998), culminating in his second synthesis of the European Palaeolithic, The Palaeolithic Societies of Europe (Gamble 1999). Here the central geographical concerns with time and space are still present, as is the concern with scale, but the emphasis has shifted from environment to society. The exposition of hunter-gatherer social ecology in The Palaeolithic Settlement is replaced by an account of hunter-gatherer interpersonal relations, and how these bring society into existence. Information flow through social systems is replaced by the knowledgeable social individual, and the communication of knowledge between them. At the centre of Gamble’s outlook is a new approach to the long-standing interest in human action in time and space. Rather than pursuing this from the point of view of the subsistence quest and the mapping of movement onto the spatial structure of the environment, Gamble explores how social space is created through relationships between individuals, and how the spatial and temporal extents of the actions and resources that are brought to bear by individuals on their relationships ‘stretch’ sociality across time and space.

Gamble perhaps makes too much of this. The boundaries between these orders of social network must necessarily be fuzzy, particularly if one considers the possibility of polygamous reproductive practices that prevent a coalescence of the intimate network about the nuclear family, or the possibility that, for at least certain periods

One aspect of The Palaeolithic Societies that deserves mention here is the account of social network theory and its relevance to the Palaeolithic (Gamble 1999, 42-62). Network theory (Maryanski 1993; Milardo 1992, Turner

8

at certain times, social networks might fission along other lines, for example segregation by gender or the temporary dispersal of sub-adults. Furthermore, network theory has been developed through studies in modern western societies, so its applicability to Palaeolithic Europe is questionable. For example, the identity Gamble draws between extended networks and hunter-gatherer regional populations, smaller groups within which define their categorical identity through emblemic style in material culture (Wiessner 1983) is quite speculative. However, his primary emphasis is on the universality in the Lower and Middle Palaeolithic of interpersonal social relationships constructed and negotiated without recourse to the representation of persons through material culture symbols. Only with the Upper Palaeolithic are material symbols recruited in the construction of social personhood. But that means that there are still two dimensions of resource in social relationships – intimate emotional and effective material – that might in principle leave archaeologically visible traces and thereby offer a point of entry into Lower and Middle Palaeolithic sociality. For Gamble, these traces include raw material transfer distances, the lengths of lithic reduction chaïnes opératoires and the application of specific skills to specific encountered situations in the landscape.

the others necessarily leads necessarily to a one-sided view of the Lower and Middle Palaeolithic past; 2. Human action takes place in space and time, and on a number of spatiotemporal scales from the reach of the body in action to the long–term occupation of regions by populations; 3. The environments human beings inhabit are heterogeneous in space and time on levels from the immediate and present to the glacial and continental. Linkages between human behaviour in the world and the various spatiotemporal scales of variation in it must be addressed by a geographical archaeology; 4. The more a geographical archaeology relies on specific interpretive frameworks and methodologies derived from studies of modern ways of life, the less likely it is to penetrate the Lower and Middle Palaeolithic. General principles of relations between environment, knowledgeability and sociality are more likely to produce results since they rely upon a more general order of uniformitarian principle; These cardinal points together constitute a programme in which the dynamic relations in time and space between people and the components of the environments in which they are immersed occupies a critical position. It has therefore been decided to refer to the perspective developed in this study as ecological geography. This does not imply a rejection of the ‘economic’, ‘palaeoeconomic’ or ‘ecological’ tags; it simply reflects the writer’s view as to a term that best encapsulates the processes being examined.

The Palaeolithic Societies therefore represents a major departure for a geographical archaeology in the sense that it brings to bear on the social aspect of human behaviour in the environment a body of theory essentially new to palaeolithic archaeology. Gamble’s enthusiasm for knowledgeability is refreshing, as his attempt to archaeologically operationalise sociality in the Lower and Middle Palaeolithic. But these gains come at a price. All space is now social space, of which propertied physical space is a mere adjunct. All time is now social time, to which the rhythms of the material world are subordinated. Ecology and geography on continental and glacial spatiotemporal scales are mentioned only in passing. A focus on the individual conditions a general orientation towards the micro and away from the macro. The goal of a geographical archaeology of the Lower and Middle Palaeolithic entailing a reciprocity between ecology, sociality and knowledgeability is missed, because multiscalarity is not fully embraced.

This study aims to build upon the work of the figures discussed above, and to develop a theoretical and methodological standpoint from which a general ecological geography of the European Lower and Middle Palaeolithic can be effected. Attention will then be focussed on a specific ecological-geographical problem, the Middle Palaeolithic leaf point phenomenon. The concept on which this theorisation must turn is that which runs through the work of the Cambridge School of Palaeoeconomy and which unifies the four essential pillars listed above: scale. The role of simultaneous climatic and environmental variability on multiple temporal and spatial scales, and of the complex dynamics arising from interactions between various biotic processes responding to exogenous change at different rates, must be taken into account if a productive understanding of Pleistocene European environments, and the character of their settlement by tool-making hominids, is to be achieved. Despite the sometimes explicit sensitivity of the authorities discussed earlier to the problem of scale in human ecology, only rarely (e.g. Bailey 1981, 1983c; Butzer 1982) has scale been considered as a central, rather than a peripheral, concern for archaeology. It is, then, appropriate to begin with a consideration of the importance of scale in ecology.

Summary and Aims: Towards an Ecological Geography A number of key themes have emerged from this review: 1. The recognition that ecology, sociality and knowledgeability are mutually connected factors in the generation, maintenance and development of human behaviour is essential if a geographical archaeological approach is to be successful. The privileging of any over

9

10

CHAPTER 2

HIERARCHIES OF SCALE AND THE ECOLOGY OF KNOWLEDGE perceive a temperature range of only 12ºC. At a response time of 24 hours, the output signal has an amplitude of only 4ºC (see Figure 2.2). This illustrates the attenuating or damping effect of increasing response time on input signals of a given frequency. The effect of increasing response time can also be seen on output signal phase relative to that of the input signal. Woodward and Sheehy (1983) found that, for two response times of 1 hour (equivalent to a thick, broad leaf) and 5 hours (equivalent to a stem around 30 cm in thickness), then the perceived temperature output signal not only showed reduced amplitude at the longer response time, but also displayed a significant time lag relative to the input signal; that is, increased response time led to input and output signals increasingly out of phase with each other.

Ecology, Scale and Hierarchy In an investigation of the relationship between climate and plant distribution, Woodward has shown how air temperature varies continually on temporal resolutions from 0.3 seconds to 3000 years (Woodward 1987, 29-32). In each of his time series curves, shown here at Figure. 2.1, the temperature data is averaged, either through the absence of measurement on shorter resolutions so that higher frequency variability is not recorded (0.3 s, 30s and 2h interval time series,) or explicitly by the plotting of calculated mean temperatures (daily, monthly, annual and four-yearly intervals). Curves for 50 year and 3000 year intervals are derived from oxygen isotopic measurements rather than from direct measurements of temperature. This averaging has the effect of damping down or filtering out from each time series temperature variability operative at shorter time scales. Yet each series displays significant, often cyclical variability characteristic of its own particular scale of temporal resolution. These curves clearly indicate that temperature change is not a single process, but a complex array of processes operating on multiple time scales simultaneously. There is, then, no such thing as ‘the temperature’ at a particular time and place, but rather an array of temperatures, each relating to a particular periodicity of temperature change. It is meaningless to refer to temperatures in the past unless the appropriate temporal scale is specified. This point has been reiterated, with an emphasis on climate in the Upper Pleistocene, by King (1996).

Woodward concludes (1987, 21) that, in broad terms, environmental signals whose frequency is at least ten times lower than the reciprocal of response time (response frequency) will induce output signal responses exhibiting scarcely any amplitude loss or phase lag. Input signals whose frequency is at least ten times higher than the response frequency will be so heavily damped or attenuated as to stimulate no output response at all; and input frequencies close to the response frequency will induce responses of varying degrees of attenuation and lag. Process-Rate Hierarchies. The applicability of these conclusions is not limited to the relations between plants and temperature change, although the importance of that for a Pleistocene ecological geography is clearly considerable. As a generalisation, the impact on any biotic system or entity of any perturbation in conditions will depend upon both the rate of the perturbation process and the characteristic frequency of the mechanism whereby the perturbed system or entity responds to the perturbation. Since the environment is constantly changing over a vast number of interconnected parameters, ecological relations can be conceived of as a network of interactions between dynamic entities, both biotic and abiotic, proceeding at many different process-rates simultaneously. Hierarchy theory, a wide-ranging body of analytical techniques and concepts based upon this understanding of system dynamics, is frequently employed in contemporary ecology (e.g. O’Neill et al 1986; Dooley and Bowers 1998; Karlson and Cornell 1998; Peterson et al 1998) and

What Woodward is seeking to understand, however, is the sensitivities of plant populations to these various scales of temperature change. That is, what are the response times of vegetation physiological and population processes to changes in temperature? Clearly, forests do not expand and contract every 24 hours in response to the daily cyclicity of daytime and night-time temperatures, nor even annually across the seasons. On the other hand, many annual summer herbs do indeed exhibit radical seasonal ebbs and flows in distribution, and produce flowers that open during the day and close at night. Woodward shows (1987, 28) how plant response times of 1 hour, 6 hours and 24 hours to changes in air temperature over 5 days - the input signal - produce quite different sensed temperature curves, or output signals. Whereas a response time of 1 hour will perceive all of a 20ºC daily range in air temperature, one of 6 hours will

11

Figure 2.1. Temperature variation at resolutions of (a) 0.3s, (b) 30s, (c) 2h, (d) 1 day, (e) 1 month, (f) 1 year, (g) 4 years, (h) 50 years and (i) 3000 years. After Woodward 1987, Fig. 2.7.

12

‘an ecosystem’ in the abstract, nor to conceive of an ecosystem as a discrete object or entity in space and time. Just as one cannot talk about ‘temperature’ without specifying the temporal resolution of temperature measurement, so one cannot talk about ‘the ecosystem’ without specifying the domain of scale to which one is referring. This involves rather more than simply identifying the area and time depth of study. Hierarchical Dynamics. In an excellent review of hierarchical concepts in ecology O’Neill et al (1986) describe the intra- and interscale dynamics of hierarchical ecological systems. They introduce a number of key terms and concepts that are of considerable value for the development of an ecological geography of the Early Palaeolithic: 1. Ecosystem. Tansley’s concept, introduced in 1935, is now virtually a universal currency within and beyond ecology, although it has no clear and universally accepted definition. Archaeologists have sometimes used the term without clarity simply to denote a collection of species of a particular age and from a particular site or region, for which a spurious, self-evident systemic unity is implied (e.g. Müller-Beck 1988). As has already been intimated, hierarchical ecology challenges notions of ecosystems simply as spatially bounded aggregates of individuals (e.g. ‘populations’, ‘species’), taxa (such as ‘the plant community’) or flows (trophic pathways, for example). Rather, ecosystem boundaries in space and time are problematic, and enclose many different classes of components that present difficulties of identification because their visibility is scale-dependent:

Figure 2.2. Effect of response time on perceived temperature change in plant organs. After Woodward 1987, Fig. 2.6.

affords a conceptual framework for the incorporation of multi-scale variability into ecological analyses.

Depending on the spatiotemporal scale or window through which one is viewing the world, a forest stand may appear (1) as a dynamic entity in its own right, (2) as a constant (i.e. non-dynamic) background within which an organism operates, or (3) as inconsequential noise in major geomorphological processes. Thus it becomes impossible to designate the components of the ecosystem. The designations will change as the spatiotemporal scale changes. (O’Neill et al 1986, 83; emphasis in original)

The hierarchical approach to ecology can be understood by analogy with a series of photographs of increasing magnification taken by an exceedingly powerful zoom camera from an orbiting spaceship (Forman 1995, 11-12). Beginning with a view of the planet and zooming in progressively on continents, regions, localities and on until one can see microscopic soil particles, eachsuccessive photograph reveals heterogeneity and clustering of discrete interacting components. At different magnifications different domains of scale or hierarchical levels are revealed. In this analogy magnification is equivalent to temporal resolution and the size of visible clusters or components represents signal frequency or process-rate. With increasing magnification, components visible at lower resolutions disappear from view and whole new complexes of components, previously invisible, become apparent. This analogy illustrates two important implications of a hierarchical approach. Firstly, the concept of hierarchy applies equally to ecological relations operating on different spatial, as well as temporal scales. Secondly, it is not possible to speak of

From a hierarchical perspective, then, there are no bounded ecosystems. We may select a spatial extent and/or time period for study, but it will always be composed of smaller-scale ecosystems in interaction, and those in turn will be composed of further subsystems on yet smaller scales. Of course, the ecosystem on the selected scale of observation and analysis will itself also be a subsystem in interaction with others as components of larger scale ecosystems. The problem facing the analyst is to define the scale of interest appropriate for the research question in hand, and to identify how far dynamics at higher and lower scales must be taken into account to explain the observed phenomena.

13

2. The ‘Black Box’. If ecosystems exhibit organisation or structure, then, according to Simon, this is a consequence precisely of differences in process rates (Simon 1962, 1969, 1973). The frequency-specific mechanisms that link input and output signals as discussed by Woodward can, for heuristic purposes, be thought of as ‘black boxes’ (O’Neill et al 1986, 77; King 1996) that connect cause and effect in ecological interactions. As discussed above, signals passing through a black box may be attenuated or amplified, as well as phase shifted, depending on the respective characteristic frequencies of the input signal and of the black box itself (see Figure 2.3). Inputs from high frequency or high process-rate components (daily variations in temperature, for example) may thus have little or no effect on lower frequency, lower process-rate components (e.g. the expansion of forest). The consequence of this in a linear system is its organisation into vertical arrays of process-rate levels. As one ‘ascends’ the hierarchy from smaller to larger spatiotemporal process scales, successive levels in the hierarchy are sealed off from each other; each level acts as a filter, so that high frequency, high process rate dynamics are restricted to lower organisational levels and their expression at higher levels in the hierarchy integrated or averaged. This has been called ‘loose vertical coupling’ (Simon 1973). Conversely, oscillations and perturbations operating at low process rates, and which are therefore characteristic of high levels in the hierarchy, will be transmitted fully to lower levels since

black boxes transmit input signals of frequencies much lower than that of their own responsiveness with little or no attenuation. Perturbations at high levels in the hierarchy will therefore have effects at all lower levels, perhaps to the extent of inducing a profound restructuring of the system and loss or replacement of many of its components, or even system collapse. The higher levels may, on the other hand, be insulated from low level perturbations by the filtration of signals as they are transmitted upwards through the hierarchy. In non-linear systems more representative of real-world dynamics there may be the potential for a much wider range of sometimes chaotic responses. Catastrophic perturbations at lower hierarchical levels, for example, may so alter the integrated signals delivered to higher levels that the tolerance limits of the higher levels are surpassed. The reverberation across the system of exceptional responses to extreme perturbations on any particular hierarchical level may therefore be a crucial structuring event in system history. 3. Process-Rate Surfaces and the Holon. Even within a hierarchical level there will be variation in the rates at which the dynamics of interaction between system components proceed. Any particular component may interact strongly (i.e. frequently and/or with high mutual sensitivity) with some components and more weakly with others. Insofar as those components that interact

1. f=n

a=x

2. f=2n

a=x/2

3. f=4n a=x/4

INPUT SIGNAL

BLACK BOX, RESPONSE FREQUENCY fr=n

OUTPUT SIGNAL

Figure 2.3. Input signals (left) produce output signals (right) of increasingly low amplitude (a) as input frequency (f) increases beyond the response frequency (fr) of the ‘Black Box’ coupling mechanism.

14

frequently and strongly are mutually conditioned, their relations with components with which they interact rarely and weakly will tend to be integrated. Systemic clusters of strongly interacting components can then be conceived as if they were a single entity from the point of view of the signals that they deliver to other components or clusters of components. Such an entity is therefore defined by a process-rate boundary or surface, and has been termed a holon by Arthur Koestler (Koestler 1969). A mole and a trout, for example, may both have similar lifespans, both reproduce with similar frequency and both be affected by seasonal variations in temperature; both are referable to the same hierarchical level in a temperate European regional rural ecosystem. Yet the direct interactions between the two are slight. The mole will interact strongly with changes in the populations of earthworms, say, and with other moles, while the trout will be closely coupled with the dynamics of aquatic insects; one can conceive of a terrestrial undersoil holon and a stream holon of which the mole and the trout are respective components. The two do, of course, ultimately impinge upon each other, but the connection operates at much slower process rates.

of horizontal coupling dynamics as of vertical linkages through frequency-specific ‘black boxes’ (O’Neill et al.1986, 80). It is important to stress that holon boundary surfaces need not in any way be identical with tangible boundaries, such as membrane, skin or bark, that delimit a material organic entity such as a cell, an individual mammal or a tree, nor even with edges that define the extent of vegetation patches in a landscape. Such boundaries and edges can certainly be thought of as holonic, but intangible holon surfaces can also be identified in process rate fields (Allen et al 1984). Metabolic cycles and species gene pools may be regarded as holonic entities even though, as cycles and pools, they lack discrete tangible form entirely. It is the frequency of mutual interaction between its components that defines the holon. Such interaction may be ‘cooperative’, as is the case with intra-cellular metabolic processes or symbiosis; alternatively it may not, as in the cases of parasitism, predation and niche competition. Koestler’s holon concept illustrates how a hierarchical approach in ecology is non-reductionist in character. One cannot privilege any spatiotemporal scale of ecological or biological relations as the source of all higher scales, or regard all scales of form and process as epiphenomenal but the lowest (contra Dawkins 1976, 1982, 1988), nor even see the hierarchy of structure as emerging sequentially from level to level, lowest to highest. Rose, Lewontin and Kaman summarise the point:

At the same time as a particular holon is composed of frequently-interacting components, the holon itself interacts at various rates with other holons, and so can be thought of as a component of a larger scale holon higher in the hierarchy. Similarly the low frequency inter-holon dynamics of one level become the high frequency intercomponent, intra-holon dynamics of the next higher level. What appears as a system of components in dynamic mutual relation at one scale of observation will appear as a single component at a higher scale, and as a subcomponent at still higher scales. A hierarchical level that can be decomposed analytically into holons may be said to display ‘loose horizontal coupling’ (Simon 1973). Chaoimh has shown, for example, that spring and summer temperatures in Europe north of the Mediterranean are significantly negatively correlated with the extent of January-February maximum snow cover (Chaoimh 1998). Clearly, at a monthly scale of observation maximum winter snowfall and springsummer temperatures are discrete phenomena which cannot interact directly by virtue of their temporal spacing; it can be said that they constitute two holons. At the annual scale, however, the two are strongly linked and form aspects or components of a single systemic entity or holon, namely inter-seasonal temperature variation in non-Mediterranean Europe.

…the material universe is organized into structures that are capable of analysis at many different levels. A living organism - a human, say - is an assemblage of subatomic particles, an assemblage of atoms, an assemblage of molecules, an assemblage of tissues and organs. But it is not first a set of atoms, then molecules, then cells; it is all of these at the same time. This is what is meant by saying that the atoms, etc., are not ontologically prior to the larger wholes that they compose. (Rose et al 1984, 277-8)

Equally, holism is an inappropriate way of thinking about ecosystems since one can never comprehend the whole through a single scale-specific ‘observation window’, nor regard those high level ‘macrostructures’ (individual organism, population, community, ecosystem, biosphere, etc., depending on one’s disciplinary perspective) that are visible as the real ‘end point’ to which lower levels are merely subordinated contributors.

The dominant time constants or process rates become slower as one ascends the hierarchy, and output signals from components are averaged as they become part of a holon’s aggregated output to other holons, and thus to higher levels of system organisation. The greater the process-rate distance between two scales of interest, the less will be the recognisable influence of the lower on the higher. The stratification of hierarchical systems can therefore be understood as being as much a consequence

4. Asymmetry and Constraint. Interactions between ecosystem components need not proceed equally in both directions; X may influence Y without Y exerting an equal, or even any, influence on X. The coats of many northern mammals become white in response to the onset of winter, but this change in coat colour does not affect the weather. The relation between seasonal climate

15

change and mammal coat colour is asymmetrical. Asymmetry is a fundamental feature of ecosystems. In particular, the asymmetry of signal transmission between higher and lower hierarchical levels discussed above is critical to ecosystem structure. Grene (1969) has argued that hierarchical organisation involves a double asymmetry in that lower level dynamics are essential to the functioning and persistence of higher level structures which in turn constrain flexibility and response range at lower levels. Constraint is therefore of fundamental importance in understanding system or component persistence in time and extent in space.

1982; Foley 1984). However, it is obvious from this discussion that the use of stasis concepts such as equilibrium can lead to confusion and imprecision unless the appropriate spatiotemporal scales of observation and analysis are established. This danger derives in part from the problems inherent in applying cybernetic concepts to ecosystem analysis. Cybernetics regards systems as selfcontrolling through the operation of positive and negative feedback mechanisms and as maintaining themselves in homeostasis (in a stable steady state) or homeorhesis (in a stable trajectory of change). They respond to perturbation from the outside by self-organised return to preperturbation homeostasis or homeorhesis, or by collapse. This emphasis on equilibrium as internally-maintained stable response has produced important insights in ecology (e.g. Patten and Odum 1981); ecosystems are indeed structured by interactions between their constituent components. However, as O’Neill et al point out (1986, 46), cybernetic approaches in ecology are founded on an understanding of the community as a superorganism. This depends in turn upon the notion of a self-organising system of aggregated individuals as a discrete real-world entity distinct from its surrounding environment and bounded in space and time. In fact this idea of a community or ecosystem as a superorganism is no more than a metaphor. It is useful in certain circumstances but limited in its ability to make sense of systems which can only be treated as real bounded entities within a dynamic scale range effective at specified scales of analysis and observation. Cybernetic approaches are also of limited value in understanding systemic responses to perturbation other than homeostasis, homeorhesis or collapse.

For example, woodland insects with annual or sub-annual life cycles play important roles in pollination, seed dispersal and the recycling of the nutrients in fallen trees, and are therefore essential components of forest ecosystems. At the same time they are constrained by forest population dynamics, which operate on time scales of centuries or even millennia; if deforestation occurs in response to, say, global or continental scale climate dynamics referable to yet longer time scales, the woodland insects disappear with the trees. This means that, if one were to study the life cycle and population dynamics of a tree-dependent beetle in an extensive forest over a period of twenty years, the woodland could be disregarded as a constant, a non-dynamic overarching constraint or boundary condition. Conversely, a study of forest spread in the Holocene of Europe could safely disregard annual and sub-annual fluctuations in the beetle population as background noise, since the signal these fluctuations deliver to the scale domain of Holocene tree population dynamics is so heavily attenuated and integrated with signals from other components that its effect is unmeasurable.

Connell and Sousa (1983), in an extensive review of the literature on equilibrium in ecology, concluded that the demonstration of equilibrium in ecosystems is fraught with difficulty, and that the very idea derives from a 19th Century Romantic ‘balance of nature’ paradigm. Lloyd and Stupfel make the same point with regard to metabolic and physiological rhythms with wavelengths of less than 24 hours. They point out that organisms are systems open to the exchange of matter and energy, that they perform far from any equilibrium state (which is attained only on death) and that in such systems oscillations can develop and be maintained. In their view the emphasis on equilibrium and steady state is a product of investigator intellectual bias traceable to the work of the medical physiologist Bernard (Lloyd and Stupfel 1991; Bernard 1856). It may, then, be advisable to treat the equilibrium concept as a useful way of describing a set of dynamic relations that, at a particular scale of observation, display significant temporal persistence within a range of variability. This variability will be perceived as ephemeral at broader scales of observation, and at smaller scales will appear as a non-dynamic constant. Equilibrium is therefore a problematic concept where ecosystems exhibit unstable or chaotic behaviour or where the research problem requires that the ecosystem be treated as distributed across a range of spatiotemporal

Any particular level in a hierarchy can be regarded as a non-dynamic constraining constant in relation to much lower levels and, at the same time, as stochastic noise by comparison with dynamics taking place at much higher levels. It is suggested here that the hierarchical ‘distance’ one must travel in both directions from a specified level in a hierarchy before one reaches an upper ‘ceiling’ of non-dynamic constraint in one direction and a lower ‘floor’ of stochastic noise in the other be referred to as the dynamic scale range characteristic of the ecosystem, component, process or phenomenon being investigated. This range will depend not only on the horizontal and vertical organisation of the system in the hierarchical vicinity of the relevant level, but also on the spatiotemporal scale of the ecologist’s ‘observation window’. In general, the shorter the time depth and the smaller the spatial extent of a study, the more the dynamic scale range is likely to be skewed towards lower levels in the hierarchy. 5. Equilibrium. A considerable body of technical terms is employed by ecologists and by ecologically minded archaeologists and palaeoanthropologists to describe patterns of ecosystem stasis and movement (see Butzer

16

scales. The latter at least certainly pertains to any investigation of Pleistocene human ecological geography.

may constitute an external perturbation for prey species, but the increase is itself coupled with the size of the prey species population. There are strong negative feedbacks between the predator and prey populations that are internal to the ecosystem at a higher level; the relation displays an important degree of symmetry at an appropriate scale of analysis. This is exemplified by Norrdahl’s investigation of regular cyclicities in the populations of northern small mammals of 8-11 years in boreal forests and 3-5 years in arctic tundra (Norrdahl 1995). On the other hand, plant life cycles may incorporate seasonal climate change but they can have no effect on the astronomical factors that generate this variable. Insofar as life cycles are horizontally coupled with seasonal climate change within a rate-process surface the system has incorporated the fluctuation, but the relation is strongly asymmetrical. This asymmetry means that seasonal climate change cannot be considered as a perturbation for ecosystems considered on annual and longer time scales, but it nevertheless emanates from outside ecosystems considered on virtually any spatiotemporal scale.

The greatest difficulty with the equilibrium concept, however, is its assumption that an ecosystem can be described as being either in equilibrium or out of it. This implies that there is a single ecosystem state coupled with one, or at most a few, realms of external perturbation. In fact, in a hierarchically-organised ecosystem composed of elements interacting at different rates, perturbations take place and induce responses on multiple spatiotemporal scales simultaneously. The problem arises of establishing which segments of the ecosystem are in equilibrium with which scale of perturbation. Furthermore, signals that are perceived as perturbations in external constraining conditions at lower levels in the hierarchy are stable internal dynamics within or between holons at a higher level, and ineffective background noise at yet higher ones. There cannot be any single ecosystem state nor any hard and fast boundary between the internal and the external. Different elements of the ecosystem will be coupled horizontally and/or vertically with different scales of input signal that will vary independently of each other to a greater or lesser degree. Relative to any particular scale and category of input signal, then, the various vertical and horizontal components of the ecosystem will at any time be in varying states of equilibrium or disequilibrium. Equally, any particular ecosystem component will, at any time, be in varying states of equilibrium and disequilibrium relative to the various scales and kinds of input signals it perceives. Relative to scales of perturbation that lie outside the dynamic scale range, no interaction will be visible, so the equilibrium concept is, in that circumstance, redundant.

7. Multiplicity of Hierarchies. Unidimensional hierarchical schemes abound in biology and ecology. Examples are the Linnaean taxonomic hierarchy, the individual-population-community-ecosystem hierarchy of population ecology, and functional-ecological trophic hierarchies ascending from primary fixer to top predator. In fact, real ecosystems at any spatiotemporal scale are implicated in multiple hierarchies. An individual organism may be decomposable into organs, tissues, cells etc. and itself a component of a population, a species, a genus and of higher taxa. It is also a phenotype emanating from a genotype and contributing to a gene pool; it is an active behaving entity whose actions can be regarded as the products of sensory and motor physiology, and also as elements of, for example, a life cycle or a seasonal migration pattern; it is a complex of directed nutrient and energy fluxes, and a component of larger scale fluxes; it may be a product of and a contributor to patterns of social interaction; and all these things at once. Any ecosystem element or component of any class is therefore multivalent, immersed in many cross-cutting and interwoven hierarchical dimensions simultaneously. Abiotic components of the environment such as nutrient minerals, trace elements and soil are implicated in hierarchies and cannot, as a consequence, be marginalised as external to biotic systems (O’Neill et al 1986, 62-3). As is argued below, perception and knowledge may also be admitted as hierarchically-organised ecosystem components, in which case there can be no hard and fast boundaries between an ‘inner’ realm of consciousness and an ‘outer’ physical environment. The decision as to which classes of hierarchy are important depends upon the nature of the study being undertaken.

6. Incorporation. A perturbation that is external and capricious at one level is internal to, that is, it is controlled by, a higher hierarchical level. For an individual annual herb, the onset of winter is an external and fatal catastrophe. But for the reproducing population of the plant, with an annual life cycle of germination, growth, flowering, pollination, seed setting, seed dispersal and finally winter seed dormancy, annual seasonal alternations are structurally internalised. Seasonality is thus incorporated at the level of the herb’s life cycle. Similarly, precipitation, wind speed and air temperature are unpredictable and capricious external variables to an individual tree, but large forests create their own rainfall, reduce air flow and dampen air temperature fluctuations; these variables are not external to the forest, but internal to it, incorporated by it (O’Neill et al 1986, 164). Incorporation, then, illustrates the ambivalence of the concept of perturbation. By definition, perturbation is external and derives from outside the system, but a perturbation at one level may be an internal dynamic higher in the hierarchy. It is important in this context to consider the degree of symmetry or asymmetry in the interaction. An increase in the predator population, say,

8. Catastrophe. Zedler et al (1983) have shown that the fire-adapted California chapparal recovers well from small fires providing that their frequency does not exceed

17

one per year, but that more frequent conflagration exceeds recovery capacity and drastically alters the vegetation. The implication is that vegetation communities considered on landscape or regional scales may incorporate smaller scale catastrophic disturbances in this case fire - providing that the catastrophes are limited both in spatial extent and in frequency (O’Neill et al 1986, 167). Under such a regime patches within the landscape may experience severe disturbance or even destruction, creating localised conditions for colonisation by opportunistic low-succession species. At the landscape or regional scale the vegetation is a mosaic of such patches, with all successional stages of re-establishment present. Each patch furnishes a seed source for the successional development of nearby patches. The vegetation ecosystem at regional or landscape scales maintains an equilibrium dependent upon the amplitude and frequency of catastrophically-induced disequilibrium patch disturbance or destruction. This is the historically important concept of the shifting mosaic steady state, introduced by Bormann and Likens (1979a) and inspired by A. S. Watt’s pattern-process hypothesis of cyclical upgrade-downgrade dynamics in vegetation community history (Watt 1947). Bormann and Likens (1979b) concluded from a study of catastrophic disturbances such as fire and storm in Swedish boreal forests and balsam firs in northeastern USA that steady states are often established in these northern forests. Their analysis was of a broad geographical sweep, but Henry and Swan found comparable upgrade-downgrade dynamics of species representation in a small (0.04 hectare) area of mixed coniferous-deciduous forest in New Hampshire with a known history of catastrophic fire and storm (Henry and Swan 1974). However, Romme (1982) could detect no such steady state in Yellowstone Park on his 73 square kilometre scale of analysis. The identification of a shifting mosaic steady state is, then, clearly contingent upon analysis at the appropriate scale, which may or may not be relevant to the system dynamics of the problem in hand.

easily be understood in terms of incorporation as discussed above. Woodward’s key insight, however, is that catastrophic amplitudinal excursions beyond the normal range in a signal of a particular frequency can have an immediate and far-reaching impact higher up the process-rate hierarchy, and that this is a crucial exception to the general rule that high frequency signals are damped as they are transmitted up the hierarchy. Woodward and Sheehy (1983) cite the winter of 1981-82 in Cambridge, UK, when minimum air temperatures fell to –16.1ºC. Exotic trees of warm-winter species such as Cistus and Hebe were killed outright. The evergreen oak Quercus suber, despite escaping death, suffered a complete defoliation that, if repeated annually, would have placed the species at a serious competitive disadvantage relative to the native deciduous oak Q. robur, which exhibits faster photosynthetic metabolism over a shorter season and which invests less energy in its relatively thin leaves. The central point here is that the catastrophe concept is ambivalent. A catastrophe at one scale is a structurallyincorporated variable at a higher scale, and this incorporation may underwrite a persistent equilibrium at larger spatial and temporal scales even when the history of catastrophe and succession at local and sub-local spatial scales and shorter time scales is random and chaotic. At the same time, rare and extreme climatic events at any time scale may so exceed the tolerance of some species that they disappear from all or part of the landscape or region, or at least are placed at a competitive disadvantage; a radical, non-linear restructuring of the higher-level ecosystem may ensue. Plant species presence, abundance and distribution will therefore be an expression of complex histories of both regular and catastrophic climate variation, of other catastrophes like fire and storm, and of succession, all operating at multiple spatiotemporal scales. Yet it remains possible for such ecosystems to display a persistent equilibrium in terms of species representation and distribution, providing the appropriate spatiotemporal scales are identified.

Fire, storm, flood, drought, overgrazing and pest or parasite epidemics are all natural catastrophes that may be incorporated at higher system levels and thus become essential structural features of the landscape or regional ecosystem and indispensible to its persistence. Woodward and Sheehy also consider the catastrophic impact of climatic events. Woodward emphasises the crucial importance of amplitude in regular climatic cycles, such as the annual alternation between summer and winter at temperate latitudes, in determining plant species geographical distribution. An annual temperature range of -30ºC to +30ºC would be a catastrophe for frost and chilling-sensitive tropical evergreen trees, but not for temperate deciduous or boreal coniferous trees, primarily because of life cycle parameters. For Woodward the gross latitudinal forest zones of the Holocene are products of competitive exclusion operating through selection for life cycles most cost-effective in particular annual temperature ranges (Woodward 1987, 24-6). This can

9. Area, Grain, Grain Responsiveness. The distribution of plant and animal species and communities in space is never perfectly homogeneous. There are always discontinuities and clustering (Butzer 1982, 8), and this is true at whichever magnification of zoom photograph one chooses. This clustering confers upon land viewed at any spatial scale, i.e. over any defined area, a texture or granularity that can, in an extension of the photographic metaphor, be glossed as grain (MacArthur and Levins 1964; Godron 1982; Guthrie 1984). “Thus, scale refers to the spatial proportion of a mapped area, and grain refers to the coarseness of elements within the area” (Forman 1995, 10). A coarse-grained area is composed predominantly of large clusters or patches, and a finegrained area of small patches. Of course, the precise meanings of ‘large’ and ‘small’ are context-dependent.

18

dogmatic ontology. Ecosystems and their manifold classes of components are not necessarily real, bounded entities in the world so much as scale-specific phenomena whose reality is dependent upon a particular scale of observation and analysis. From this starting point one can outline a general theory of ecological relations which recognises the multiplicity of dimensions and of spatial and temporal scales on which the world varies, and which conceptualises the dynamic processes that link, or otherwise, components within and between scale domains or process-rate levels. The nine interlinked key concepts discussed here furnish a language in which these dynamic relations can be expressed, and a repertoire of insights from which ecological analysis can proceed.

‘Grain’ is almost invariably used in ecology with reference to the size of stands of vegetation in humanscale landscapes, and that is the sense in which it will generally be used in this thesis. However, it can be applied at any scale, and Forman draws from this an important principle: The resolution of an aerial photograph thus depends not only on the physical grain and contrast present, but also on the ability of the observer to distinguish spatial elements. Similarly animals in the landscape have different grain responses, i.e. perceive or are sensitive to different grain sizes. The deer…responds positively to the clearings provided by wildlife biologists, but does not perceive small spider holes in the ground or the pattern of regions in Texas. The mammoth…responded to the mosaic of wetlands and ridges, whereas its fleas responded to individual hairs and hair densities on the mammoth’s knee… This perception of the grain size of the environment is a key to landscape functioning and change. Animals and humans perceive and respond to only a fraction of the multi-scale heterogeneity present. The spatial elements causing a response determine the direction, route and rate of movement.

The critical importance of these insights for an ecological human geography of the European Pleistocene can hardly be overstated. Human beings are and were organisms of a particular size with characteristic sensory and conceptualising capacities and propensities, rates of movement through the landscape and life cycle parameters in time. In addition, Pleistocene human survival depended on the consumption and exploitation of organisms of various sizes, mobility, generation times, life spans and ecological tolerances, and all distributed unevenly in time and space. One hardly need add that the European Pleistocene was a period of high-amplitude cyclical climate variability on wavelengths from the daily and annual, through sub-Milankovitch scales of centuries and a few millennia, to Milankovitch wavelengths of tens of millennia (King 1996). These scales of variation operated simultaneously and irregularly, and both reinforced and countered each other; and all were played out on spatial scales from the individual stand, through landscapes and regions to that of a topographically, climatically and hydrologically diverse continent. And finally, human behaviour (or, if one dislikes the ethological tone of that term when applied to people, practice) was linked to the world by scale-limited modes of perception and conception which, insofar as they were learned or acquired, inhered as much in the social aggregate and the individual’s social experience as in biological-genetic propensities and predispositions.

(Forman 1995, 10-11; emphasis in the original)

Thus far in this consideration of scale in ecology, ecosystem components have been described, following Woodward, as ‘perceiving’ input signals in a metaphorical sense. If a flow of nutrients between a woodland stand and surrounding grassland is disrupted or transformed by, say, a change in the prevailing wind direction, then the nutrient flow as an ecosystem component can be said to have ‘perceived’ the change via some linking mechanism or black box which translates the input signal into an output signal. But although one may be able to conceive of a nutrient flow as a holonic entity defined by a process-rate boundary and with reality at its own scale domain, it is not a superorganism and lacks sentience. Forman’s point illustrates that sensory perception by sentient, or at least sensing, beings is itself a kind of black box linking mechanism that couples a heterogeneous world in flux with organisms’ action in it, or behaviour. Indeed, animal sensory perception as the interface between the world and appropriate behaviour is the paradigm of ecological cause-effect linkage, otherwise Woodward’s metaphorical use of the term ‘perception’ would be meaningless. Perceptions and sensitivities must therefore be regarded as ecosystem components in their own right, and as having phenomenal reality no less than that of any other component visible at any particular scale. From this perspective, grain responsiveness is the equivalent in the realm of cognition of the dynamic scale range.

There are clearly major obstacles in the way of achieving an ecological human geography founded on these insights. The archaeological and palaeoenvironmental records are incomplete and riddled with problems of resolution, correlation and chronology. They are, in any case, only proxies for past practices and systems of behaviour in changing environments, not practices or environments in their own right, except, perhaps, for archaeologists and Quaternary palaeoecologists. Neither can one afford to make the uniformitarian assumption that the scalar character of historically and ethnographically-known human relations with the world pertained also among Pleistocene human populations. On the contrary, palaeolithic archaeology is nothing if it is not, ultimately, a comparative undertaking seeking to make sense of difference in a deep human past. It is the

Towards a hierarchical human palaeoecology An emphasis on scale in ecology, like the adoption of a hierarchical perspective, is not to be mistaken for a

19

contention of this paper that no understanding of the ecological character and environmental contexts of European human settlement in the Pleistocene can adequately address this past if it fails to account for the multi-scalar nature of both people and the world in which they were immersed.

mental states entirely in terms of component physiological and biochemical processes, and for historical disciplines whose time depth disposes their practitioners to adopt supra-individual systemic-holistic perspectives. Glynn has pointed out that these difficulties were made explicit 120 years ago by William James (Glynn 1993; James 1879). James was dissatisfied with the dominant Cartesian-interactionist view that physical events in the brain caused mental states but that mental states could cause neither physical events in the brain nor subsequent mental states; that is, that consciousness is an epiphenomenon. He argued that if consciousness is epiphenomenal and lacks causative power, then it can have no influence on behaviour, and will have no survival value. Only real behaviour, untouched by consciousness, would be relevant to the struggle for survival. Why, then, has consciousness evolved? Glynn concludes that this problem remains unresolved, but that epiphenomenalism can offer no possible solution and that real causative power must be accorded to mental states.

The Ecology of Knowledge All organisms are aware and knowledgeable. An organism that does not react to stimuli is dead. Even the simplest bacterium is able to sense the pH, say, of its surroundings and will mount a response. This may involve biochemical-physiological pathways that effect a selective secretion or absorption of ions across the cell membrane. It may be behavioural in character, for example movement in space that relocates the organism in a pH gradient. It may be both. That the organism senses things that are important to it and acts accordingly indicates that it has awareness and possesses a means whereby the information it acquires provides a cue for action; that is, the organism knows what to do. It must, in turn, have a memory. Nicolis has put the point well:

James’ point has also been made, from a completely different starting point, by Lukács, the founder of western Marxism, in his refutation of what he saw as Engels’ scientific epiphenomenalism: …only in labour, in the positing of a goal and its means, consciousness rises with a self-governed act, the teleological positing, above mere adaptation to the environment – a stage retained by those animal activities that alter nature objectively but not deliberately – and begins to effect changes in nature itself…Since realization thus becomes a transforming and new-forming principle of nature, consciousness, which has provided the impulse and direction for this, can no longer be simply an ontological epiphenomenon.

Living beings are undoubtedly the most complex and organized objects found in nature…They are literally historical structures, since they have the ability to preserve memory of forms and functions acquired in the past, during long periods of biological evolution. (Nicolis 1989; quoted in Lloyd and Rossi 1993)

There is little controversial in attributing to the sensorimotor and sensoriphysiological apparatuses of ‘lower’ organisms ecological reality. It is a different matter when awareness and knowledge is held to inhere in consciousness. It is true that Butzer has no difficulty; for him, human ecosystems (perhaps ‘ecosystem interactions between humans and the components of their environments’ would be a more accurate way of putting it) are distinguished by the roles played by information, technology and social organisation, but, more importantly, by the capacities of human individuals and groups for purposive behaviour and conscious goalorientation. Together with the significance of “value systems” and “group attitudes”, these factors render archaeology as human ecology impossible unless “the pivotal role of human cognition” is recognised (Butzer 1982, 32). Butzer’s assertion that these factors, which are all aspects of human conscious awareness and knowledgeability, are unique to human ecology is open to question (see Ingold, 1986a, 1988b; McGrew, 1992; Boesch and Tomasello 1998). The point here, however, is that Butzer regards human consciousness as internal to ecological relations, not external to them.

(Lukács 1980, 22-3)

Lukács’ conviction that consciousness as a real phenomenon in the world is uniquely human is, of course, as subject to challenge as Butzer’s. It is, in any case, inadvisable to presume either ecologicalpsychological identity or difference between modern peoples, with which both Butzer and Lukács are concerned, and the non-modern inhabitants of Pleistocene Europe, since a specific understanding of the latter’s interaction with the world is the aim of this investigation, not its starting point. Nevertheless, the case for consciousness, knowledge and perception being accorded ecological systemic reality is unanswerable once the validity of the domain scale on which it operates is admitted, as a hierarchical approach demands. The fallacy of epiphenomenalism derives from classical science’s drive for objectivity, which compelled – and still compels – unreflective scientists to privilege infra- and supraindividual levels of the world over that of the aware human agent, at which level the centrality of subjective consciousness to human action is inescapable. From this perspective it can be said that the epiphenomenalist antipathy of the American New Archaeology, as

Butzer does not, however, seem to be stating the obvious. Consciousness poses particular difficulties for both reductionist sciences that seek to explain behaviour and

20

exemplified by Binford (1962, 1965) and Watson, LeBlanc and Redman (1971), to individual consciousness as a dynamic element in the human past is a consequence of its quasi-scientific insistence upon the absolute primacy of dynamics at supra-individual system levels over dynamics at the level of individuals.

intention here is to consider the character of perception and knowledge as linkages between the world and action in it from a hierarchical point of view. This will begin with an examination of affordance. Perception and Affordance

This ontological privileging of the supra-individual is also a feature of structuralist thought which, with delicious aptness, inverts the New Archaeology’s ‘material:ideal::real:epiphenomenal’ opposition. For structuralists only meaning and the ideal realm have reality, since structuralism is founded on a pervasive linguistic metaphor in which human signification relates arbitrarily to concepts of the world, not to the material world itself, which is therefore dismissed as unknowable. Hence materiality is seen as epiphenomenal and lacking in causative power, and the transforming power of the idea relates only to a self-referential field of relational meaning. Neither the material world nor practical human action in it play any role; cultural phenomena are products of interplays in structured oppositions of meaning that transcend the individual subject, which is subsumed and determined (Gosden 1994, 45-51). This is why Lévi-Strauss could acknowledge geology, Marxism and psychoanalysis as his three inspirations, since all of these derive superficial surface appearances from deeper underlying realities of structure (Lévi-Strauss 1955), and why Bourdieu could dismiss structuralism as ‘objectivism’ (Bourdieu 1990). Structuralism therefore shares with the New Archaeology an epiphenomenal conception of knowledgeable individual human action or agency. Taken together with its disregard for practical reality in a material world, this renders structuralism a particularly unsuitable mode of thought for the development of a geographical human palaeoecology, despite its commitment to the reality of ideas.

The late American psychologist J.J. Gibson, whose work has been introduced to British anthropological thought by Ingold (Ingold 1986b), devoted much of his career to investigating the nature of animal visual perception of the environment. His central insight was that, although the world may well be composed of atoms, molecules and energy fluxes for physicists and chemists, these are not what animals see. Gibson developed an elaborate scheme that sought to describe and explain how, by detecting ecological information, animals perceive the opportunities and hindrances afforded by the things around them. In the major exposition of his ideas, The Ecological Approach to Visual Perception (Gibson 1979) he contends that animal visual perception is an internalisation of ecological values or affordances present in the world in which the animal is immersed – its environment (Reed 1988a). What is more, perception is not identical with sensation. The latter arises from the passive stimulation of the senses and does not gather ecological information. Only when animals actively seek information by looking purposefully at the environment is ecological information gathered, and only this constitutes perception. The perceived environment is, then, not wholly external to the animal but is also a product of the animal’s own purposeful activity, so that the animal and its environment exist in mutual relation mediated by perception. There is no organism without a surrounding environment, and no environment without a surrounded organism (Gibson 1979, 8). The Earth before life evolved may have been a world but it was not an environment (Ingold 1986a).

Post-structuralist thought as exemplified by the work of Derrida and Foucault has sought to reassert human agency by emphasising that the relational field of signifiers from which meaning derives exists only insofar as it is produced and reproduced in action, and this view has exerted considerable influence on post-processual archaeology over the last ten years or so (e.g. Shanks and Tilley 1987, 1992; Tilley 1991; papers in Bapty and Yates 1990). Nevertheless, post-structuralist archaeologies remain wedded to the ontological primacy of the ideal over the material, despite Hodder’s cautionary words (Hodder 1989), and effectively conceive of action as coterminous with interpretation. Post-structuralism offers no way of understanding the dynamics of past human action in a changing Pleistocene world.

Gibson referred to his ideas as ecological psychology. From the perspective adopted here this may appear an overly narrow view of ecology, which has been defined as multi-scalar and not confined solely to those scales of observation visible to a visually-perceiving animal. This would be an unfair characterisation. Gibson’s work, although he does not explicitly recognise it, deals with the world as seen with a human grain responsiveness. His failure to appreciate the varying levels of grain responsiveness between species is related to important flaws in his work that will be discussed further below, but the emphasis on human scales of perception means that his ideas are of especial relevance to human ecology. In The Ecological Approach Gibson gives his account of the elements or values of the ecological environment. These are not the elements of the periodic table but a specific realm of ecological elements that animals perceive directly. The elements include:

It is not, however, the purpose of this paper to advance a critique of neo-evolutionary functionalism in archaeology over the last 35 years, nor to review the vast literature of polemic that has been spawned by the so-called ‘processualism vs. post-processualism’ debate. The

21

1. Media; more-or-less transparent and fluid, media - air and water - do not resist the movement of objects through them and thus afford locomotion. They also transmit, and are filled with, vibratory waves emanating from events, and so afford vision, hearing and smell. Media therefore afford a matrix of possible observation points, linked by possible locomotory pathways, from which an animal can acquire ecological information about its environment.

hindrance are objects. This is to be contrasted with the general or universal affordances such as breathing afforded by air as a medium, or the danger of falling to one’s death afforded by a brink as a layout feature. These general affordances tend to inhere in the more persistent components of the environment. Gibson identifies two special categories of object, tools and other animals, that are discussed below.

2. Substances; more-or-less solid and opaque, substances resist deformation and the movement of objects through them. Substances vary widely in their properties, for example elasticity, plasticity, hardness, density etc., and thus in their affordances. They are generally heterogeneous. For a terrestrial animal, water has an ambiguous quality as at once a medium that offers some resistance to movement and affords not breathing but the danger of death by drowning, and also a non-rigid substance, confined to bounded volumes in space, that critically affords the quenching of thirst.

5. Paths afford locomotion (or perhaps transit would be a more useful term; locomotion per se is afforded by the support and friction of the solid ground surface, and air’s low resistance to movement) between terrain surface features that hinder or prevent locomotion, such as obstacles, barriers, water margins and brinks, and constitute or connect openings in cluttered or wrinkled terrain. These are all layout properties of the earth’s surface and, as such, are liable to change. 6. Destruction is not an ecological element as such but an event, and one that it is important for an animal to perceive. It involves disappearance following events such as burning, death or consumption but is quite different from disappearance following mere occlusion. The destruction of environmental objects, particularly those whose size and persistence contribute to ground surface layout and those which afford important opportunities or hindrances, may alter the spatial relation of resources, dangers, paths and viewpoints. At the same time destruction is counterbalanced by restitution whereby objects of the same or similar affordance to those destroyed are generated. The environment therefore exists for Gibson in a kind of steady state or equilibrium in which ephemerality and persistence coexist.

3. Surfaces are regions where a bounded volume of substance abuts against another or against a medium. The ground is the most crucial surface, and affords support and, together with air as a medium, pedestrian locomotion. Surfaces have layouts, of which the topography of the earth’s surface is the most fundamental example because it is the most persistent. Various classes of object may be distributed on a surface, and these objects may be more-or-less persistent and more-or-less mobile. The layout of the ground as a surface may afford visibility or occlusion, depending on relief and the height of opaque substantial objects attached to it, and on the particular location of the perceiving animal on the surface. Open ground surfaces (Gibson confusingly calls them ‘environments’ although they do not exhaust the surroundings of the organism, the definition of ‘environment’ he uses elsewhere) are relatively ‘unwrinkled’ and ‘uncluttered’, and afford high visibility and locomotion in many directions. Cluttered ground layouts, by contrast, hinder visibility and afford locomotion only at openings. The layout will change through time, but some aspects will persist more than others so that the layout of the earth’s surface as a perceived environment will exhibit variance and invariance simultaneously.

7. Tools are a special class of object that afford the application of force to other components of the environment. When in use they are not only attached to the body but are a functional extension of it, so that the boundary surfaces that separate the body and tool, and therefore the body and the environment, must be regarded as permeable or mobile. 8. Other animals are objects that differ from inanimate objects by virtue of their affordance of interaction, as opposed to merely action. Central to animacy and interaction is movement; Gibson treats plants as indistinguishable from non-living objects and shows, on the basis of much experimental work, that animacy is itself an ecological value readily perceived by animals. A further category of object of particular relevance to human beings is the human conspecific, which affords not only interaction but proper interaction. These distinctions are important in that different categories of ‘other individual’ afford quite different things and must be treated quite differently.

4. Objects are regions of space occupied by a substance and bounded by a surface. Objects must exist on perceptible scales; neither atoms nor continents are ecological objects in Gibson’s terms because their size does not afford being seen. Objects may be attached to other objects, such as a tree to the ground or a limb to the axial skeleton, or detached, such as a pebble on a beach. Attached objects may become detached by the application of force (although, it must be said, this distinction is essentially one of degree rather than of kind and depends upon the nature and amount of force required to detach the object). Most (but by no means all) of the entities in the environment that afford specific opportunities and

9. Distance and duration. Ecological elements or values do not passively ‘fill’ empty abstract time and Euclidian space. There is no prior empty time and space to be filled

22

and animals cannot perceive them. Instead, animals perceive distance and duration as properties of ecological values and events.

kaleidoscope, the specific shape of the environment may change while its components and their texture or grain remain within narrow limits of variability. The shifting distribution of resources in the mosaic may then be accommodated simply by acting in the here-and-now on the ecological information acquired through perception, the linking mechanism or black box whereby behaviour accommodates or adapts to environmental change.

Gibson explicitly intended that his ideas should illuminate the active ecological immersion of ‘human ancestors’ in their environments, and emphasised the importance of direct or technologically unaided perception as opposed to indirect perception mediated by, for example, microscopes and telescopes that make visible scales of the world other than those that he regarded as ecological (Gibson 1979, 10). The value of Gibson’s achievement in this regard is considerable. Ecological psychology provides a framework for understanding how spatiotemporal variation in the world on the scales of visual perception impacts upon human ecology through perceptive activity.

Adaptation and Affordance Perception The adaptation concept requires some qualification. The term is widely used in ecology, evolutionary biology and archaeology, and a commitment to the universality of adaptation by natural selection is a defining feature of the neo-Darwinian New Synthesis in evolutionary biology (e.g. Dobzhansky et al 1977). So strong is this commitment that a dogmatic doctrine of adaptation has been identified and termed ‘adaptationism’ (Lewontin 1978, 1993; Gould and Lewontin 1979; Gould and Vrba 1982; Rose et al 1984). This dogma lies at the heart of neo-evolutionist anthropologies (e.g. L. A. White 1949, 1959; Steward 1953, 1955; Harris 1979) and archaeologies (e.g. Binford 1962, 1965, 1968; Flannery 1967, 1968; Watson et al 1971) in which human social and ecological relations are conceived as adaptive culture systems. Adaptationism in this sense remains a significant force in palaeolithic archaeology (e.g. Gamble 1986; Binford 1989; Clark 1992), but adaptation is an ambivalent term with multiple meanings even for those sympathetic to neo-Darwinism (Foley 1987, 56-7).

Affordance and Adaptation If human beings are immersed in a surrounding environment characterised by a particular degree and distribution of ‘wrinkling’ or ‘clutteredness’, whether this derives from topography or the abundance, spatial extent and distribution of occluding and obstructive ecological entities such as forest stands, then this will have a profound impact upon the locomotory paths that are perceived to exist, and thus on patterns of mobility. The relationships between these paths and the spatiotemporal occurrence of important affordances such as plants and small animals for eating, game ungulates for eating and skins, wood for artefacts and fuel, leaves for bedding, stone for sharp edged tools, water for drinking, caves for shelter and other people for cooperative action, sex and reproduction, for example, will be critical factors in shaping the patterns of individual and group behaviour. The simple presence or absence of key affordances like food and water, at least for critical periods, may determine the habitability of an environment, although this may also be influenced by factors such as the duration and distanciation of a range of key affordances.

For the purposes of this discussion, adaptation will be defined in a very general sense as a directional response by a behaving component entity to external perturbation. This incorporates both adaptation as process, the dynamics of a black box mechanism linking perturbation as cause with response as effect, and adaptation as outcome of the adaptive process, the particular form or dynamic state of an entity whereby the experienced perturbation is accommodated. By understanding adaptation as a directional accommodation, this definition excludes chaotic responses to change and responses that culminate in the demise of the component or entity. However, it does not imply that there can be one and only one adaptive response to a perturbation, nor does it insist upon the universality of any one mechanism whereby responses are generated.

Perhaps the most valuable of Gibson’s insights, however, are those concerning environment transformation in time. His emphasis is on the co-occurrence of persistence and change. In highlighting the tendency of environments to exist in a steady state or equilibrium in which some ecological values are more persistent than others and destruction is countered by restitution, Gibson is effectively adopting Watt’s ‘cyclical upgrade-downgrade’ hypothesis of vegetation dynamics and Bormann and Likens’ ‘shifting mosaic’. At this level of ecosystem change, the persistence of landforms and of species abundance and fragmentation within an area may be the scalar aspect of the environment with which habitation is coupled. Visibility and resource abundance and spatiotemporal distanciation integrated over a sufficiently large area may change very little from the standpoint of human perception. Like the changing patterns in a

Adaptation is then, from this perspective, closely linked to equilibrium in that it deals with the maintenance of sustainable relations between system components and their ambient, and is subject to the same caveats as attach to the use of the equilibrium concept. Adaptation is scalespecific, and this is true whether one is considering the process of adaptation or its specific outcome. An organ, organism, community or ecosystem (for example; any other level or class of component could be cited) may be persistently adapted to high-level constraints, responding

23

stably to low process-rate change in such constraints (i.e. ‘tracking’ change), or maintained at an optimum steady state relative to an oscillating input signal operative at a hierarchical level only marginally higher than its own. It may equally be responding to catastrophic change at lower levels. Of course, all of these may apply simultaneously, and different linking processes may be involved. Lewontin, for example, has argued that only cyclical environmental change operating on similar wavelengths to an organism’s generation time will impact upon the dynamics of its population genetics and thus induce evolution by natural selection. Higher frequency change may be accommodated physiologically or behaviourally (Lewontin 1966).

close to a lower limit of sensibility measured in seconds. The dynamics of a shifting mosaic are referable to rather longer time scales but are driven by the frequency of catastrophes such as fire, storm and exceptionally severe weather, and by the process rate of vegetation patch regeneration from nearby seed sources. Both of these perturbations occur at wavelengths which are at least in part commensurate with human biographical and subbiographical time scales, even given the apparently short human lifespans of the Pleistocene (Trinkaus and Thompson 1987). High amplitude environment change on long time scales of glacial-interglacial cycles is quite another matter, and here the limitations of ecological psychology as conceived by Gibson become apparent. The individual-centredness of Gibson’s ecological psychology limits its utility as a framework for understanding human ecology and adaptation at the level of high amplitude, long wavelength environmental variation. The solution to this problem is to be found in the solution to the second key weakness in Gibson’s ideas, namely the fallacy of direct perception and the shared environment.

One cannot, then, talk of adaptation as a unitary process or state or without specifying the scale and class of perturbation or input signal to which the entity or component is adapting. An adaptive component response to change at one hierarchical level may be thought of as an internal dynamic at a higher level, and vice versa. The designation of what is ‘external’ will therefore depend on the scale of observation. Equally, it cannot be assumed that the mechanism whereby human populations living at northern high latitudes have tended to lose skin pigmentation necessarily operates according to the same principles as those that link tree species representation in woodland communities to climate. Descent with modification through differential reproductive success – natural selection – cannot be expected to explain all change through time in all aspects of all classes of biotic system.

‘Projects for Living’ and Received Knowledge. In one sense Gibson’s emphasis on direct perception is obviously justified. Pleistocene human palaeoecology cannot proceed on the basis that Palaeolithic people were aware of scales of existence that can be perceived only indirectly through the mediation of devices such as microscopes, telescopes and aerial photographs. But this is not the fundamental meaning of direct perception for Gibson. For him, ecological perception is direct because it is not a subjective construct in which a sensory stimulus is rendered meaningful by the mediation of any mental process. Rather it is an undivided functional activity of animals as observers, the product not of a mind operating on a stimulus, but of a unified ‘perceptual system’ whereby animals see the world immediately and as it really is (Reed 1988b). It is no accident that Gibson describes the elements of the environment as ecological values; the term is not intended as a metaphor. Value, meaning and significance for Gibson inhere in the environment and the objects in it and are grasped directly by the perceiving animal. They are not imposed on the world by historically specific projects (Reed 1988a). Animals, even animals of different species, therefore perceive the same environment, which is shared and public, not private. Because animals are mobile, they occupy myriad points of observation in the course of their lives so their view of the environment is not specific to a perspective that only a single individual adopts. Action is motivated by values which originate outside the animal but which are internalised by perception and action. Both Gibson and Reed are clear, following this argument, that affordances as ecological values are properties of matter, and their existence is not dependent upon their discovery by any animal, including human beings. Despite Gibson’s recognition of the ecological reality of perceived

People’s behavioural responses to environment change, then, must be considered from a scale point of view if an understanding of their ecological character is to be understood. If it begins to rain, I may adapt by sheltering under a nearby tree. The process involved in this is individual perception of affordances – rain affords wetness and therefore coldness and discomfort, the leaves of trees are objects which, when clustered together in the crown of a tree, afford shelter – and of their ecologicalspatial relationship to each other and the soon-to-besoaked observer – the rain is ambient in the air as medium, but that tree over there is close enough to afford me shelter before the rain gives me a thorough soaking. I am therefore making judgements as to the affordances surrounding me, how close the tree is, how heavy is its foliage, how hard it is raining, how quickly I can traverse the terrain between myself and the tree. I am informed of the perturbation (i.e. the onset of rain) and of an object that affords accommodation to it (the tree) by perception, and perception guides my motor response. Perception enables adaptation by bodily action. Both fine-grained change through time in a shifting mosaic and the onset of rain are scale-limited. In terms of amplitude, neither effects a radical transformation in the affordances of the environment. In terms of temporality, the onset of rain is, in perceptual terms, a rapid event

24

information, it has no necessary connection with human consciousness and knowledge which are, for him, every bit as epiphenomenal as they are for neo-evolutionists. Gibson’s theory of perception requires no theory of mind.

argue that water affords, and has always afforded, cooling for internal combustion engines, but Neanderthals failed to spot this. In truth, pointed sticks afford piercing for planting only for medium and large bodied animals which possess prehensile-manipulative capacities and which cultivate plants – i.e. human agriculturalists. Only in the light of the agricultural project does a pointed stick afford this function.

There is, then, a contradiction at the heart of ecological psychology. On the one hand, perception comes about not passively, but only through an active search for ecological information. The perceived environment therefore finds its source as much in the perceiving organism as in the world outside it, so the organism and its environment exist in mutual relation. On the other hand, perception is objective in a Baconian sense, and directly furnishes the observing animal with unmediated information about an environment whose existence and value are quite independent of any particular perspective or scheme. The environment does not rely for its existence on the perceiving animal, and affordances as values pre-exist their own discovery. Ecological psychology cannot have it both ways.

We return, then, to the knowledgeability of all organisms. The appropriate course of action for an organism to take in response to a perceived affordance depends upon its project, which is historical in character. Heritage confers upon an organism a way of perceiving the world, of appropriating it into its own scheme and acting accordingly. A frog is no more equipped to fly than a thrush is to absorb oxygen transcutaneously underwater, but each knows perfectly well how to ‘do its own thing’. The organism therefore brings the world into functional relationship, with itself at the centre, by virtue of its historical project, and thereby creates its environment (Ingold 1986a). The project for living is literally a body of received knowledge in that the organism’s physiological, perceptive and motor organisation that resides in the chemical and morphological architecture of its body, it receives from its predecessors. At the same time the organism’s specific behaviour is a matrix of enacted appropriate expressions of this received knowledge in experienced, i.e. perceived, environmental contexts. ‘Knowledge’ is, of course, used metaphorically, and the metaphor is an ambivalent one that subsumes two quite distinct, but linked, sets of processes operating at two quite different scales. Knowledge as a received orientation to the world can be thought of as a supraindividual ecological entity that is transmitted through the generations and which is strongly horizontally coupled with perturbations in the world at characteristically evolutionary time scales. Knowledge as the appropriate enactment of responses to perturbations perceived through the senses, on the other hand, is a property of the individual and operates at event timescales. The vertical relation between the two is doubly asymmetrical sensu Grene (1969) in that individual action is constrained by the received project (as with the frog and the thrush) but the project persists only insofar as individuals enact it and reproduce it.

Gibson implicitly concedes that the standpoint of the organism determines ecological values as much as do the properties of matter when he asserts that water is a substance for people but a medium for a fish. A similar argument could be made for, say, a leaf. For a deer it is an object that affords eating, for a fly it is a surface that affords support and for a caterpillar it is both. The ecological value of any entity in the environment depends not only on its inherent material properties, but also on the terms upon which each animal engages with it. Affordances, then cannot be eternal and immutable properties of matter in itself. Even gravity, which may be regarded as the very paradigm of a universal, objective physical law of nature by which all organisms are bound, turns out to be nothing of the sort. For a vast number of microorganisms gravity has no relevance at all; their tiny size, together with their immersion in a fluid medium, renders them weightless. They are, however, constantly buffeted by another ‘universal natural force’, Brownian motion, to which mammals are entirely insensible by virtue of their comparatively great bulk (Rose et al 1984, 275). None of this, of course, is intended to imply that environmental affordances are the invented products of unconstrained organismic creativity. The point is that the affordances of the environment are embedded in the mutuality of organism and environment. The opportunities or dangers that environmental entities and forces present to any organism are indeed functions of their objective material properties, but what those properties afford, if anything, depends on an organism’s grain responsiveness and upon its project for living and cannot be prior to the organism’s own subjective orientation to the world. This is also true of human technicity. It is absurd for Reed to argue that pointed sticks afford, and have always afforded, piercing the ground for planting but that this affordance was only recently discovered (Reed 1988a, 113). One may as well

This vertical asymmetry has a tendency, however, to be only partial. Processes and events at the individual scale may impact upon the received project, amending it in some way. This is, in a general sense, what the modern concept of evolution conveys. To put it another way, sensory perception is a property of the individual, and although the acts that it informs operate on biographical and sub-biographical time scales they will, in hierarchical terms, deliver signals to higher scale domains of behaviour. Nevertheless the latter are not reducible to the former. Gibson’s affordance concept retains validity on long time scales insofar as long wavelength ecological change is identical with long wavelength transformations

25

in the material properties of the world, from which the affordance environments defined by organisms are not independent. Knowledge as a historically received project is indeed a linking mechanism between animal behaviour and the affording world on these time scales. Sensory perception is not.

of an organism defines ‘aptness’; the species is therefore as much the ‘cause’ of natural selection and its own evolution as is the environment (Rose et al 1984). By the same token, behavioural plasticity at the individual level may reconstitute aptness, and organs or behaviours that arose in evolutionary time for one purpose may be coopted for another. Evolution may therefore be exaptive rather than adaptive in Gould and Vrba’s terminology (Gould and Vrba 1982). Of course, evolutionary-genetic change may come about through mechanisms other than selection, such as genetic drift and the founder effect. Finally, an individual’s knowledge of how to behave in the world is not located solely in its genotype and thus in its phylogenetic heritage, but also in its own development in the context of experience, that is, in ontogeny (Gould 1977). These objections may be sufficient to undermine reductionist adaptationism but they do not alter the fact that genetically-physiologically mediated knowledge is transmitted only by sexual reproduction and the passing on to progeny of genes that are immutable in the face of experience, if not of chance.

Domains of Knowledge: Structure and Agency Thus far the discussion has studiously avoided drawing any distinction between human and animal knowledge. Perhaps this should be put another way; the distinction between instinct and thought has been avoided. In part, this is because to oppose the one to the other is fraught with danger in terms of definition and of empirical recognition (A. Leroi-Gourhan 1993[1964], 219-27). In any case, it is not clear that thought is restricted to humans (e.g. de Waal 1982, 1991; Baker 1978), while it is obvious that many important aspects of the human behavioural repertoire (smiling, narrowing one’s eyes when dazzled, grasping objects with the fingers or placing objects in the mouth, for example) are in large measure innate. The question here is not so much ‘instinct vs. thought’ as biological-genetic ‘knowledge’ vs. learned knowledge. For an insect, presumably, its project is the bequest of its phylogenetic heritage, and it is reasonable to assume that its knowledge of ‘how to go on’ in the world, together with its application of this knowledge, resides in and is perpetuated through the genetic realm. For a human one could say the same about the conscious realm. From a hierarchical point of view, these two forms of knowledge repository and guide to action are in some ways comparable. Both constrain and enable comprehension of the world, and the repertoire of appropriate action in it, as projects for living that precede the individual. Both structure and inform the individual’s actions in the world. The central difference from the point of view of hierarchical ecology is their mechanisms of transmission in time.

Learned knowledge is quite different. It is acquired by experience, and for people this essentially means social experience. We learn from other people how to go about the business of life and how to make sense of the world. This knowledge may come to us by myriad means. We may watch other people and seek to copy their actions. They may show us how to do, recognise or name something; that is, we may be explicitly taught. They may explain to us or tell us what to do. They may reward us for some actions and punish us for others. We may rebel, and others may join us. We may deploy our bodies in a stylised manner for social purposes unthinkingly and repeatedly so that the knowledge becomes habituated or intuitive rather than conscious (Bourdieu 1977). What begins as action guided by conscious knowledge may also, through repeated and routinised application or expression, become habituated (Gosden 1994). We may acquire knowledge not from personal contact with people at all, but in engagement with their material sublimates such as artefacts, monuments and texts. We may have access to narratives (A. Leroi-Gourhan 1993[1964]). Most likely all of these mechanisms will contribute to our acquisition of knowledge. Crucially, a human individual’s repertoire of knowledge, unlike his or her genes, is liable to change in the course of a lifetime as a result of experience. What is more, experience is not formless or undirected. What we experience is goaldirected engagement with the world, Lukács’ ‘teleological positing’ (Lukács 1980).

Genetic transmission or heredity, at least for vertebrates living outside modern cloning laboratories, is achieved by sexual reproduction. Genes which, according to an axiom of modern genetics, remain entirely immune to directional change induced by the life experience of the individuals that carry them, are thereby passed on to progeny. The most widely recognised mechanism whereby processes at the level of the individual impact upon the genetic repertoire of a species in time is, of course, natural selection; genotypic variation between individuals underwrites phenotypic variation, and the phenotypes that are most apt for living in the encountered environment tend to produce the most progeny. Genes that underwrite apt phenotypes proliferate in the gene pool at the expense of those that underwrite less apt phenotypes. The species’ store of genetic knowledge, and therefore its project for living, is transformed. Evolution has taken place. One may object that the organismenvironment relation is mutualistic, and so the way of life

People, then, are immersed from birth in a social milieu, and it is in this context that we come to know the world in which we are immersed and with which we engage. The properties of the world are not irrelevant; the materiality of the world limits what we can know of it. But our inexpressible sense of immersion in the world and of our proper relations with it, the way we see and understand it, our perception of the affordances our

26

environment contains, all flow from practical engagement with the world carried out in terms defined by the social knowledge we acquire in the course of life. Human projects for living, then, are received just as surely as those that are genetically-physiologically mediated and just as surely predate us. Ingold’s Sartrean commitment to each individual’s potential to recast the world anew (Ingold 1986a) simply cannot accommodate the individual’s irreducible sociality and the impossibility of escaping society’s conceptual shackles by standing outside it. In hierarchical terms, socially transmitted knowledge is a high level structuring constraint on human action.

to be understood broadly so as to include intuition and acquired habit as well as calculation and reflection. From the point of view of the project for a geographical palaeoecology of the European Pleistocene, the character of the coupling between these two knowledge domains, and between each of them and a world changing on multiple time scales, is of critical importance. Hierarchy theory provides a set of questions to ask. To what extent does variability at the scale of agency – behavioural flexibility – deliver a damped signal to the domain of socially transmitted knowledge? That is, to what extent is their vertical coupling asymmetrical? Does the vulnerability of different aspects of socially transmitted knowledge to perturbation from below vary with the particular mechanism of their transgenerational transmission? What may the links between each scale domain of knowledge and the negotiation and construction of social relationships be? These questions address the problem of the relation between agency and structure, which has been a central and intractable problem in Western thought since the Enlightenment (Shanks and Tilley 1992, 123) and which may be conceived of as a modern formulation of the Christian theological problem of free will versus predestination. Nevertheless, answers to some of these questions are suggested in some contemporary social and political theory.

Socially transmitted knowledge as a field of values, attitudes, rules and concepts that precedes, transcends and directs individual action clearly shows close parallels with structure as understood by structuralism and discussed above. However, it differs from a LéviStraussian mentalism in that it does not construe structure as an interplay of binary opposites or take language to be a universal metaphor valid for the interpretation of all dimensions of human experience. More critically, classic structuralism is interested only in the ideal realm and ignores practical human action in the material world. Socially transmitted knowledge is understood here, by contrast, to be a coupling mechanism which connects humanity with the material world on super-biographical time scales, defines the affording environment, structures practical action and which possesses reality at its own scale domain. It can, then, be understood as tradition, or, more precisely, as an expression of an ecological aspect of culture as classically defined by Tylor:

Symmetry and Asymmetry: Structuration A concise solution to the problem of vertical coupling between agency and structure is to be found in Marx:

[Culture is]…that complex whole which includes knowledge, belief, art, morals, law, custom and other capabilities acquired by man as a member of society.

Men make their own history, but they do not make it just as they please; they do not make it under circumstances chosen by themselves, but under circumstances directly encountered, given and transmitted from the past. The tradition of all the dead generations weighs like a nightmare on the brain of the living. (Marx 1969, 398)

(Tylor 1871, 1)

Social transmission of knowledge as understood here is intended to be an explicitly ecological and hierarchical concept which throws into sharp relief the confusion over scale that dogs cognitive ecologies in anthropology. Butzer, for example, recognises the structuring reality of socially transmitted knowledge when he refers to the ecological significance of ‘group attitudes’ and ‘value systems’; but he draws no distinction between these and ‘purposive behaviour’ and ‘goal orientation’ (Butzer 1982, 32), which refer not to overarching knowledge systems but to their contextual application on biographical and sub-biographical time scales.

‘Circumstances…given and transmitted from the past’ (i.e. structure), then, constrain the scope of human freedom to act (agency, knowledgeable action), but do not still it. Men (sic) can still ‘make their own history’ (i.e. transform structure), but they begin from a starting point that precedes them. The relation between the two domains is incompletely asymmetrical. And of course, Marx is forever associated with the conviction that, at critical historical junctures, action can bring about the revolutionary transformation of structure.

We are left, then, with two domains of knowledge by which people are connected to the world; a domain of individual and collective action, of agency, of knowledgeable action informed by affordance perception and constrained by a hierarchically higher domain of socially transmitted knowledge that confers affordance values on the material world. ‘Thought’ in this context is

The most influential recent treatment of this question, at least in British sociology, is that made by Giddens in The Constitution of Society: Outline of the Theory of Structuration (Giddens 1984). Structuration theory conceives of the world of social practice as a system, ‘the networking of relations between individuals and groups in a field of existence embracing the categories of the

27

economic, the political the ideological and the symbolic which together constitute conditions of existence for the social strategies of individuals in time-space’ (Shanks and Tilley 1992, 127). The system is structured in that practices (the term ‘behaviour’ is avoided as excessively ethological in character by thinkers of this school) are oriented or fixed at junctures in space and time by principles which transcend the spatiotemporally bounded actions they inform and underwrite. Structuring principles, however, are themselves constituted in social action and are present and reproduced only in the practices enacted by human agents. Individuals are not automata, motivated and determined by a structure that exists wholly outside themselves, but conscious, knowledgeable and purposive agents that monitor their actions and their consequences in the light of their goals, which are always social goals pursued by means of social strategies. Practical action may also have unintended consequences. There is thus a duality of structure in operation in that structure is simultaneously the medium and outcome of practice; structuring principles inform and orient practices, but are also structured principles in that they are produced, reproduced and transformed in enacted practices. Bourdieu makes the same point with his concept of habitus, which is closely related to Giddens’ structuration, and which he defines as ‘systems of durable transposable dispositions, structured structures predisposed to function as structuring structures, that is, as principles of the generation and structuring of practices’ (Bourdieu 1977, 72).

scales far beyond those that sociology considers. This is directly related to the problem of reification (Giddens 1984, 179-80), the attribution of real-world object status to supra-individual levels of organisation. From a hierarchical standpoint, the reification of social structures as distinct entities external to individuals is simply another fallacy that derives from an insensitivity to the problem of scale. As discussed at length earlier with reference to the concepts of equilibrium, adaptation and the holon, what is internal and external to a system, together with the visibility of discrete bounded entities, depends upon the scale of observation. That socially transmitted knowledge is reproduced by and made materially manifest in individual action alone cannot be taken to refute either its supra-individual systemic reality or its discreteness when considered as a component linking humanity and the world at the appropriate scale of analysis. Two more caveats remain to be dealt with, both concerning society. The chronic predisposition of practice and structure to mutual transformation propounded by Giddens and Bourdieu is predicated on practices being coopted as instruments in strategies for the pursuit of social goals. Of course there is a sense in which all human knowledgeable action, perhaps even all chimpanzee action (McGrew 1992) is irreducibly social in that it is informed by structuring principles that are socially transmitted, it is directed towards ends that have meaning only in terms of the character of the mutual relation between the agent and the world – affordance perception – that is, at least in part, socially given, and it may be cooperative, competitive or conflictual in intent. It does not follow from this that practices are necessarily coopted to play meaningful and psychological roles in the construction and negotiation of social identity. It is perfectly conceivable that social practices as knowledgeable action remain directed exclusively to utilitarian material ends. In particular, an instrumental status for technical practices as resources for social strategies in which the social subjectivity of the artefactas-person is implicit, certainly cannot be assumed to be a feature of social life in the Palaeolithic. Indeed, the apparent virtual absence of what may be termed ‘symbolic’ artefacts before the Upper Palaeolithic can be interpreted precisely as evidence that lithic artefact fabrication played little or no part in Early Palaeolithic social relations (Gamble 1998).

The point of this examination of structuration and habitus, neither of whose originators show more than a cursory interest in either ecology or palaeolithic archaeology, is to demonstrate that the two domains of social knowledge in question are seen as operating in mutual symmetrical relation in contemporary social theory. There are certain problems, however, that arise from an attempt to apply structuration theory to a hierarchical ecology of knowledge, especially if the Palaeolithic is the object of investigation. Foremost among these is the relatively vast expanses of Palaeolithic time, which are far longer than those considered by sociologists. This narrower window of sociological observation contributes to Giddens’ antipathy towards methodological distinctions between a ‘microsociology’ directed to immediate interpersonal relations, and a ‘macrosociology’ concerned with a reified supraindividual social structure (Giddens 1984, 139-44). For Giddens, the ‘micro’ and the ‘macro’ in social life are completely interpenetrating and thus analytically inseparable. His refutation of microand macrosociologies, however, rests upon the ill-advised attempts by each to claim ontological and systemic primacy for its own preferred scale of reference. A hierarchical ecology of knowledge as proposed here explicitly aims to escape from this scale-chauvinism that Giddens rightly deplores, while also insisting that, for example, the persistence of the Acheulean for 1.5 million years demands its recognition as a systemic entity on time

Secondly, there is good reason to imagine that the temporalities of social life on historical time scales can indeed be conceived in terms of domains defined by process rate, and that the vertical relations between them are less than perfectly symmetrical. The historian Braudel, in an epic history of the Mediterranean (Braudel 1949), introduced his view of historical time as decomposable to at least three classes of process, all operating simultaneously but on different wavelengths. Short term history is assigned to the realm of Événements, corresponding to events and individuals.

28

History in the medium term relates to processes termed Conjonctures by Braudel and corresponds to, for example, social and economic history, world views and ideologies (Mentalités). Long term history, Longue Durée, relates to geohistory, constraints, world views. Both Conjonctures and the Longue Durée are dimensions of structural history, dealing with constraining and enabling frameworks of life that lie beyond the perception of individuals (Bintliff 1991). If one accepts this view of history then the relation between socially transmitted knowledge as a structural constraint and knowledgeable action informed by perception is not completely symmetrical in the way that Giddens implies. A more complete account is given at Figure 2.4.

biotic landscapes presents more intractable difficulties of reconstruction, but the diversity of floral and faunal species represented in the fossil record, and the often widely distanced ranges of their descendants in the Holocene, suggests that shifting mosaic ecosystems characterised much of Europe through large segments of Pleistocene time (Guthrie 1984). Theorisation of the spatiotemporal dynamics of mosaic landscapes may, then, if integrated with the record of palaeoenvironmental change with an explicit sensitivity to scale, cast light on the grain of past landscapes, the horizontal and vertical coupling of ecosystem components and the impact of climatic catastrophes on their structure and history. The central task of a geographical human palaeoecology of the Pleistocene in general, and of the late Middle Palaeolithic in particular, is the articulation of patterns of human settlement and technicity with the ecological history of the continent. Topographical, altitudinal, and climatic factors must be taken into account in any attempt to understand the dynamics of Pleistocene human ecology and its development, but a view in which human environments are conceived as static landscape types must be transcended. It is the relationship between persistence and change in the world, and persistence and change in patterns of human behaviour, which is the point at issue. This relation affords the possibility of identifying the operative scales at which knowledgeable action and social transmission of knowledge were coupled with the world. Since the human organism and the environment are in mutual relation, and the environment’s affordances derive from a dialectic between the world’s material properties and the socially transmitted projects for living that orient people to it, the changing ecological contexts of human occupation provide what is almost a laboratory test of the scalar character of human ecology in the Pleistocene. Environmental change that is not associated with changes in settlement or technical practices in time and space may be regarded as within the range of tolerance of the received project and as accommodated by knowledgeable action and perception on biographical and sub-

A Hierarchical Ecology and the Palaeolithic There are, then, persuasive grounds for adopting a hierarchical approach to systems of human knowledge operating on different scales. To apply it productively to the ecology of human settlement in the European Pleistocene, it is necessary to specify the kinds of questions one needs to ask and the classes of archaeological and palaeoenvironmental information that may illuminate them. The starting point for this enterprise must be an investigation of the spatiotemporal variability of Pleistocene European biotic landscapes. Climate oscillations of varying amplitude are known to have occurred throughout the Pleistocene on wavelengths of tens of millennia, as recorded in, for instance, river terrace sediments, deep sea foraminifera, detrital carbonate and oxygen isotope sequences, and terrestrial pollen and faunal fossil records; and on wavelengths of centuries or a few millennia as evidenced in Greenland, Antarctic and Tibetan ice core oxygen isotope, dust and atmospheric gas records, and in marine sediments, for example. Seasonal/intra-annual change is harder to resolve but may be indicated in certain malacofaunal and aeolian sediment sequences. The spatial variability of

HISTORY OF EVENTS STRUCTURAL HISTORY

SHORT TERM – EVENEMENTS Narrative, Political History. Events, Individuals. MEDIUM TERM – CONJONCTURES Social, Economic History. Economic, Demographic Cycles. History of Eras, Regions, Societies World Views, Ideologies. LONG TERM – THE ‘LONGUE DUREE’ Geohistory. Enabling, Constraining. History of Civilisations, Peoples. Stable Technologies. World Views.

Figure 2.4. Braudel’s hierarchical conception of historical time. After Bintliff 1991.

29

biographical time scales without any consequent upward delivery of a perturbing signal to the domain of socially transmitted knowledge.

directs action towards the world, including knapping, in particular ways. Lithic industrial variability in time and space, assessed at multiple resolutions and placed in the context of multiple rates of ecological change, may then be regarded as an index of symmetry or asymmetry in the dynamic between knowledgeable action and socially transmitted knowledge in the European Palaeolithic.

By focusing on seasonality and grain it may also be possible to delimit the spatiotemporal reach of human knowledgeable action, and any transformations in this reach, through the course of the European Palaeolithic. If an identification of the upper and lower scalar limits of human action and knowledge in the Pleistocene can be approached, then characterisations of human dynamic scale ranges may be deployed as a comparative methodology for the investigation of settlement and lithic-industrial practices in the Late Middle Palaeolithic.

A summary of an agenda for a geographical human palaeoecology of this brevity is necessarily imprecise. On the other hand, a properly theorised treatment of the problem on the scale of the whole of the Palaeolithic in Europe, taking into account all visible scales of observation and perturbation in both human action and the environment, and seeking to understand their myriad interactions, could not be completed in a career. The intention of Chapter 3 is to paint a broad brush picture of the pulsating ecological world of the Pleistocene and of human relations with it; and to draw preliminary conclusions as to the scalar scope of knowledgeable action and its relations with both socially transmitted knowledge and a changing world in the Palaeolithic.

The palaeolithic record of chipped stone techniques may also be analysed from a hierarchical standpoint. Stone tools are made by individuals as means to a perceived end; they are, as functional objects, the practical products of knowledgeable action. But stone does not demand that it be knapped. That humans did it, and how they did it, are expressions of socially transmitted knowledge that

30

CHAPTER 3

HUMAN ECOLOGY AND SCALES OF KNOWLEDGE IN THE EUROPEAN EARLY PALAEOLITHIC (Ruddiman and Raymo 1988; Prentice and Denton 1988; Shackleton et al 1988; Raposo and Santonja, 1995); this resulted in the persistence of a mosaic, rather than a zoned, pattern of species distribution and a high species diversity (Guthrie 1984);

The Pleistocene Mosaic Gamble and the Pleistocene Occupation of Europe The agenda for a geographical human palaeoecology outlined in Chapter 2 has been most directly addressed in archaeology by Clive Gamble. In an important recent article (Gamble 1995a) he mounts an explicitly ecological and implicitly hierarchical investigation of the earliest settlement of Europe in environmental context. This dissertation is not directly concerned with the short chronology-long chronology debate which is the focus of Gamble’s attention (see papers in Roebroeks and van Kolfschoten 1995a), but some of Gamble’s observations and claims are nevertheless of considerable relevance here.

ii. The western oceanic effect which, although variable, renders the climate of oceanic western Europe persistently less seasonally differentiated than that of continental eastern Europe (von Koenigswald 1992); iii. The longer growth seasons for both plants and animals in the low-seasonality oceanic west of Europe, which attenuated competitive exclusion and promoted a fine grained mosaic, in contrast to the east where growth seasons were shorter, competitive exclusion more intense and the mosaic coarser grained (Guthrie 1984). These are highly pertinent points which demarcate the outlines of a specific geographical human palaeoecology and which lend themselves readily to a hierarchical analysis. Still, the internal dynamics of a temporally persistent mosaic in conditions of very long term, high amplitude climatic fluctuation have scarcely been addressed in palaeoecology, as Gamble himself observes. An examination of Guthrie’s concept of the mammoth steppe (Guthrie 1984, 1990), on which Gamble’s interpretation relies, is a useful point at which to begin the task.

The coarsest-grained, most inclusive index of human coupling with the world is the presence or absence of settlement, and Gamble points out that Middle Pleistocene human occupation in Europe appears to be virtually independent both of Milankovitch-scale glacialinterglacial climatic cyclicity and of statically-conceived vegetation landscape types. Traces of occupation occur in local palaeoenvironmental contexts ranging from open grassland to closed forest, and in palaeoclimatic contexts ranging from interglacial to glacial (Gamble 1995a, 281). At some sites local occupation persists through perturbations that shift local environments from one extreme to the other, for example at Cagny-la Garenne and Cagny-Cimetière in northern France (Tuffreau and Antoine 1995) and Boxgrove, southern England (Roberts et al 1995). The presence of humans in Pleistocene closed canopy climax forests remains uncertain (Gamble 1986 versus Roebroeks et al 1992). On the other hand, in east central and eastern Europe Gamble argues that there is a persistent absence of Middle Pleistocene human occupation through glacial-interglacial cycles. This apparent restriction of Pleistocene settlement to the west of the continent will be considered in some detail later; the point of interest here is the ecosystem scale components whose operation and persistence Gamble identifies as correlating with the presence or absence of human occupation in the Middle Pleistocene. He discerns three interlinked factors:

The Mammoth Steppe A striking feature of Pleistocene floral and faunal species associations at northern temperate and sub-arctic latitudes, as known from the fossil record, is their dissimilarity from Holocene associations. This is not a matter of new species having recently evolved; the Pleistocene witnessed many extinctions, but virtually no speciation, certainly as far as plants are concerned (Tzedakis 1993). Rather, it is a case of species whose modern ranges are widely separated in space on continental and sub-continental scales occurring sympatrically in the Pleistocene. Guthrie (1984) cites numerous examples, including: the Pleistocene Alaskan tundra was home to species that occur in Alaska today (e.g. moose, reindeer/caribou, musk ox) but also to species that today have southerly (badger, ferret) Eurasian (saiga antelope, horses, camels) and grassland (bison, horse) distributions (Guthrie 1968); beetles found

i. The migration of flora and fauna within and into Middle Pleistocene Europe in response to controlling glacial-interglacial climatic cyclicity and its amplitude

31

today in tundra, forest, mountains, wetlands and cold steppe existed sympatrically in Late Pleistocene northern Siberia (Vereshchagin 1977); during the last glaciation beetles that occur today only in Tibet, and ground squirrels now restricted to central Asia, occurred in England (Coope 1974; Mayhew 1977); in central Europe the arctic collared lemming coexisted with the southern hamster (Kowalski 1967); and plant assemblages similarly lacking modern analogues are also common (Iversen 1954; Sparks and West 1972). One could add to the list. At the Grotte Scladina, Belgium, putatively cold steppic pollen spectra from the Würmian interpleniglacial are contemporaneous with a temperate microfauna (Bastin 1992; Cordy 1992; Cordy and Bastin 1992) and similar ‘contradictory’ associations occur at Piekary, Poland (Svoboda et al 1996).

weave’. This local proximity, combined with the diversity of plant resources on sub-annual time scales, underwrote an ungulate optimum that, Guthrie claims, reached its peak in the Middle Pleistocene but then collapsed with the increased seasonality of the late Pleistocene, leading to widespread regional and total extinctions. The fine-grained mosaic character of vegetation and animal distributions in the Pleistocene seems hard to refute. The strange plant and animal communities referred to above testify to persistently high levels of local diversity and landscape fragmentation. However, Guthrie’s account is also important in that it draws our attention to the inadequacy of reconstructing Quaternary environments according to preconceived static types based on Holocene stereotypes. The crucial ecological issue is the multiple scales of spatiotemporal dynamics in Pleistocene mosaic ecosystems. Clearly, Guthrie’s plaidmosaic mammoth steppe is a long time-scale version of Bormann and Likens’ (1979) ‘shifting mosaic steady state’ model of vegetation dynamics, in that its extraordinary persistence suggests an equilibrium at spatial scales larger than the local, and maintained on Pleistocene time scales.

The inescapable conclusion is that Pleistocene plant and animal associations have no modern analogues (Huntley 1996) and that the ecological dynamics of Pleistocene Europe were radically different from those of the Holocene. Plant and, to a lesser degree, animal communities in the Holocene are characteristically distributed in well-defined climatic zones determined by latitude, altitude and precipitation. The species diversity apparent in the Pleistocene record, however, suggests strongly that, at northern latitudes, the terrestrial ecosystem was a grassland mosaic, an open ‘mammoth steppe’ with variable tree cover, that extended from to Britain to Canada (Guthrie 1984). Although Gamble attributes this mosaic to species migrations climatically forced on glacial-interglacial time scales, Guthrie identifies as the constraining condition an increase in seasonality that had begun in the early Tertiary and culminated in the Holocene. In conditions of reduced seasonality relative to the Holocene, longer growth seasons limited competitive exclusion among plants since species could schedule their life cycle events (germination, production of first leaves and shoots, flowering, pollination etc) apart in time. Species could therefore persist in climates much closer to their physiological tolerance limits than is the case in the Holocene, in which plant distribution is limited by marginal competitive advantage and disadvantage in the context of a shorter growth season. The consequent diversity of plant resources in time afforded ungulates access to a wide range of dietary supplements, and this persistent feature of Pleistocene plant ecology enabled levels of local ungulate diversity encountered today only in Africa.

At the same time, there are a number of serious flaws and omissions in Guthrie’s model. Foremost among these is the boundary constraint he proposes, namely low levels of seasonality relative to the Holocene. There is evidence to support the idea that persistently high seasonality enhances the role of competitive exclusion as a process linking plant distribution with climate on time scales above the annual. An example already cited is the complete defoliation of evergreen oak in the exceptionally cold Cambridge winter of 1982. The tree survived the shock, but if such severe winter temperatures were repeated annually and the species were forced into obligate winter leaf loss, it would be at a serious reproductive disadvantage relative to deciduous oak, which has a shorter carbon fixing season but a higher photosynthetic rate (Woodward and Sheehy 1983). The problem with the seasonality explanation for the mammoth steppe is that it is virtually impossible to demonstrate low but increasing seasonality through the Pleistocene. There is no proxy measure of seasonality with even remotely sufficient time depth to corroborate Guthrie’s claims directly, and evidence derived from, for example, the palynological and malacofaunal records depends for its interpretation on the use of the very same modern analogues whose value in Quaternary palaeoecology Guthrie rejects. In any case, Rousseau et al (1990) have reconstructed the history of climatic oceanicity and continentality in Alsace on the basis of the 350,000 year malacofaunal record in the Achenheim loess. Their analysis found no trend towards increasing continentality in Alsace through the Pleistocene, but an alternation between oceanic and continental climatic regimes culminating in a period of oceanicity, and thus moderated seasonality, that persisted throughout the Würm. Meyer and Kottmeyer (1989) also found that

According to this view, the zoned climax communities of Clementsian succession theory are artefacts of extremely high levels of Holocene seasonality. Pleistocene vegetation distribution in Europe, according to Guthrie, was not zoned or ‘striped’, but chequered, like the patterns of a tartan or ‘plaid’ textile. Vegetation occurred in patches much smaller than Holocene zones such as ‘Boreal’ and ‘Atlantic’ forest, so diverse resources were always in close proximity; the plaid exhibited a ‘fine

32

aeolian sediments as proxies of wind regimes in Europe, contrary to the predictions of their model, indicate the incorporation of central Europe at the last glacial maximum into a western province predominantly influenced by westerly winds blowing from the Atlantic rather than by easterly and north easterly winds blowing from the Arctic and Siberia. It seems pertinent to add that several of the examples of diverse communities cited by Guthrie as evidence of a low seasonality mosaic actually date to the final Pleistocene, i.e. at just the period when his model would predict a decline in diversity.

McManus et al 2003) and the Lower Pleistocene (Raymo et al 1998). Crucially, the Holocene climate signal recognised by Bond and his co-workers in foram and icerafted debris curves indicates a powerful amplitude damping of climate periodicity on D-O wavelengths relative to the Pleistocene signals. The presence of this signal in the north Atlantic indicates that climate systems over the Greenland ice and the northern ocean were strongly horizontally coupled on DO time scales, but there is also evidence that this coupling was indeed global. Upper and late Middle Pleistocene events of a comparable scale have been recognised in magnetic susceptibility curves from the French Massif Central (Thouveny et al 1994), in marine pollen cores from the Iberian Atlantic (Roucoux et al 2001), in sediments from the Arabian (Schultz et al 1998) and Japanese (Ikahara and Itaki 2007) Seas and the Atlantic offshore from Namibia (Little et al 1997), in Chinese loess sequences (Chen et al 1997) and in the structure of dunes in the Kalahari (Stokes et al 1998). This evidence for D-O scale temperature fluctuations from all over the world and throughout the Pleistocene powerfully suggests

If low Pleistocene seasonality is a poor candidate for the high level enabling condition, then the 41kyr and 100kyr Milankovitch climatic cycles that, according to Gamble, controlled the gross movement of species within Pleistocene Europe, are not convincing alternatives. There is no reason to believe that these cycles ceased at the beginning of the Holocene, when the mammoth steppe disappeared. Another explanation needs to be advanced if the scalar dynamics of the mammoth steppe and its human occupation is to be found. The probable answer is the high amplitude climatic fluctuations operating on wavelengths of centuries or a few millennia revealed in the oxygen isotope curves derived from the Greenland GRIP and GISP2 ice cores (Dansgaard et al 1993; Grootes et al 1993). These so-called DansgaardOeschger (D-O) events were episodes in which mean annual temperatures above the ice sheet seem to have risen by as much as 5-8ºC – around half the rise in global temperature from the last glacial maximum to the Holocene - in the space of 50 years or less. Each was followed by a rather slower decline in temperature. At least 12 such events can be identified in the 35 kyr duration of OIS 3, most of which lasted between 500 and 2000 years (Figure 3.1). Temperature shifts of this rapidity, frequency and magnitude would clearly have exerted considerable influence on patterns of plant and animal species distribution on regional and landscape scales. There have, however, been doubts as to the relevance of the Greenland record to climate in terrestrial Europe. D-O scale temperature changes above the Greenland ice sheet cannot necessarily be extrapolated to Europe. In addition, the events are most clearly represented in the ice core segments relating Oxygen Isotope Stage (OIS) 3, between 60 and 25 kya. Because of ice compression at depths of more than 2900 metres or so, the cores do not afford sufficient resolution to determine whether these fluctuations continue back into the Middle Pleistocene. Questions over the generality of this frequency of climatic perturbation are heightened by the disagreement between the GRIP and GISP2 cores as to whether it is visible in the last interglacial, OIS 5e (Grootes et al 1993). Significant climatic perturbations of comparable wavelength have, however been detected in north Atlantic marine sediments referable to the Holocene (Bond et al 1997), the Upper Pleistocene (Rasmussen et al 1997), the Middle Pleistocene (Oppo et al 1998;

Figure 3.1. Oxygen isotope curves from the GRIP ice core, Greenland. Curve A, Holocene; curve B, Upper and later Middle Pleistocene. Note the rapid, high amplitude fluctuations in curve B. After Dansgaard et al 1993.

33

that temperature fluctuations on wavelengths of centuries to a very few millennia were a pervasive feature of climate over terrestrial Europe throughout the duration of the mammoth steppe. Furthermore, the close correlation in time between the demise of the mammoth steppe and the dramatic loss of amplitude in the D-O scale climate signal is suggestive of a causal link.

contribute to the maintenance of the mosaic by consuming seedlings and saplings, but the high level asymmetric constraining-enabling factor with which mosaic persistence is coupled is periodic temperature and climate shifts of D-O wavelength and amplitude. This would indicate that a high level constraint need not be thought of as a signal value that remains persistently stable when viewed from lower in the hierarchy, but can also be a persistent frequency and amplitude of cyclical signal perturbation; the constraint resides not in constancy, but in a constant rate of variability.

The probable impact of these sudden, high amplitude temperature changes on ecosystem dynamics can properly be described as catastrophic, in the specific sense in which the term is used in Chapter 2. Consider once again the unfortunate Cambridge evergreen oak. Very cold winters would place it at a competitive disadvantage relative to deciduous oak, which would proliferate at the evergreen form’s expense. Evergreen individuals would not necessarily be killed outright but would be unlikely to be replaced by conspecifics after eventual death. In favoured localities within the landscape such as sheltered valleys, where the extremes of winter cold may be ameliorated, patches of seedproducing evergreens may persist. The incorporation of temperature and precipitation perturbations in the centre of large forest stands would also dampen the amplitude of the intra-annual climatic cycle and afford a less hostile environment for the evergreens. Nevertheless, if the regime of extreme winters endured for a sufficient length of time, larger scale regeneration dynamics would overwhelm the evergreens on a regional scale, leading to a deciduous forest zone of limited species diversity. If, on the other hand, climate improved radically and relatively rapidly within centuries then the competitive boot would be on the other foot before sufficient time had elapsed to effect a zonation. The deciduous species would find itself at a competitive disadvantage and would persist best in localities such as stand edges, where exposure to the elements would expose evergreens to the full, unattenuated force of occasional severe winters. The consequence would be a fragmentation and overlapping of the distribution of both deciduous and evergreen forms, with some individuals of each surviving in the landscape at all stages in the D-O scale cycle and providing seed sources for local expansion in conditions of species-specific climatic optimality.

Of course, climate and temperature oscillations on other time scales were in operation in the Pleistocene simultaneously with D-O scale events. Extreme seasonal temperature excursions such as that in Cambridge in 1982 were surely also a feature of Pleistocene climate, along with cycles akin to the 15, 23, 40, 164 and 711 year wavelengths of Holocene temperature oscillation identified by Woodward’s spectral analysis (Woodward 1987, Figure 2.8). There is as yet no published study of northern latitude Pleistocene climate fluctuations on these scales, although it is worthy of note that the D-O events visible in the GRIP oxygen isotope record exhibit a degree of subordinate noise that may conceivably refer to high amplitude cyclicity on some of these wavelengths. The presence, between 15 and 33 kya, of clearly defined 200 year oscillations in the oxygen isotope record from the Guliya ice core, Tibet (Thompson et al 1997), may eventually prove to be of significance in this regard. Of course, Pleistocene ecosystems were strongly constrained by the climatic consequences of Milankovitch solar insolation oscillations on the 23, 41 and 106 kyr wavelengths visible in the deep sea oxygen isotope record (Imbrie and Imbrie 1979; Shackleton and Opdyke 1973). The gross transformation of vegetation across Middle and Upper Pleistocene glacial cycles apparent in long pollen sequences such as at Tenaghi Philippon, northern Greece (Wijmstra 1969) and La Grande Pile, France (Woillard, 1978), and the well established interglacial vegetational successions such as that in the Eemian (Zagwijn 1961, 1989, 1996) appear to be the consequence of ecosystem coupling with those wavelengths of climatic cycle. However, it would appear likely that, for so long as climate cyclicity on D-O wavelengths and amplitudes was in operation, vegetational landscapes retained their mosaic character and Milankovitch scale climate signals were accommodated by major shifts in the relative proportion of vegetation patch types in the landscape. It is for this reason that resolution of the debate over whether D-O events occurred in the Eemian interglacial would be a crucial contribution to an understanding of Pleistocene human settlement dynamics.

If this dynamic is extended beyond evergreen and deciduous oaks to include all the species occurring locally and regionally then the result would be fragmented landscapes with patchy distributions of vegetation types. Patches may succumb to unfavourable climatic shifts and the patch area be made available to colonisation by other species, but no regime of mean annual temperatures, or amplitude of intra-annual seasonal climate fluctuation, persists sufficiently to induce a narrow community of optimally-adapted species. Patches would not necessarily be fixed in space; the shifting mosaic concept predicts that, on longer time scales, they would ‘move’ around the landscape. Forces internal to the mosaic on somewhat shorter time scales, such as grazing and browsing by ungulates, may

In conclusion, the fine grained Pleistocene mammoth steppe appears to have been maintained by repeated climatic catastrophes, occurring predominantly on centennial-millennial scales but possibly also on shorter

34

time scales. These seem better candidates for the role of catastrophe than those usually advanced, such as fire, storm and pest outbreaks. To describe the mosaic as ‘in equilibrium’ or ‘adapted’ is therefore to miss the point. The multi-scale low frequency controlling signals, combined with non-linear effects induced by the persistent occurrence of climatic catastrophes, produced ecosystems in which different components within and between scale domains were coupled with different scales of perturbation. Only on Milankovitch time scales could the steppe be described as in equilibrium, and even then this had a homeorhetic rather than a homeostatic character. At lower scale domains the mammoth steppe was chaotic, pulsating, and it was precisely this that, on the one hand, afforded persistent resource diversity for human populations and, on the other, constantly posed them adaptive problems of resource transformation and relocation in time and space.

climate, vegetation communities on slopes break up the surface and direct run-off, thereby impacting on erosive degradation of landform features, and patches may be persistently distributed in plains mosaics in accordance with spatial variation in soil and bedrock. The significance of the mosaic concept for a geographical human palaeoecology lies in the precise scalar character of resource distribution within it in time and space. It follows that altitudinally-controlled heterogeneity in resource distribution may have been equally favourable for human and ungulate occupation providing that the zones, considered as patches, were sufficiently small and sufficiently closely packed to place all resources in close proximity from the point of view of a medium-large bodied mammal. Whether landscapes of this character were indeed occupied by Pleistocene human populations is addressed directly below. 2. Landscapes, Regions, Continents. Recent archaeologies and anthropologies of landscape have conceived of the landscape not as an object or array of objects external to human inhabitants, but as a social and cultural construct, a socially mediated way of perceiving and acting in the world and attributing significance to it (e.g. Tilley 1994; Gosden 1994; Ingold 1986a). Gamble’s concept of the ‘social landscape’ (Gamble 1993, 1996) is a version of this approach, and corresponds essentially to Gibson’s ‘environment’ and is an aspect of a project for living. Forman, however, defines a landscape as a scale domain of mosaic, characterised by a mix of ecosystems ‘repeated in similar form over a kilometres wide area’ (Forman 1995, 13), and extending ‘in any direction until the recurring cluster of ecosystems or site types significantly changes’ (Ibid, 21). From this point of view, plateau and valley in broken uplands are two distinct landscapes. The key is human grain responsiveness and scale of observation, which dictate Forman’s choice of kilometres as the appropriate unit of measurement. At larger scales, landscapes cluster into regions, and regions into continents.

Mosaic Patch Dynamics Although the mammoth steppe concept advanced by Guthrie is of great value, he fails to consider its internal spatial dynamics beyond making unamplified reference to the ‘plaid’. As a consequence a number of important spatial processes and properties characteristic of land mosaics go unremarked. The consideration of these that follows is based on a major recent review and synthesis of land mosaic ecology by Forman (1995). 1. The hierarchy of mosaics. Guthrie contrasts ecologically ‘zonated’ land ecosystems with the mosaic. This is, to a degree, a fallacy. The term ‘mosaic’ refers only to a heterogeneity, clustering or distancing of ecosystem entities in space and is neutral with respect to the processes maintaining this heterogeneity, the persistence or ephemerality of specific spatial distributions of ecosystem entities, and the scale of observation. It is for the investigator to make or derive these essential specifications in the light of the particular question or investigation in hand, without which the mosaic concept has no meaning. After all, the whole world is a mosaic. Major deserts, grasslands, boreal and tropical rain forests are all patches in a spatially heterogeneous distribution of vegetation on the planet’s surface. The concept of the ‘zone’ is equally context specific. Depending on one’s spatiotemporal scale of observation, it is possible to see persistent low-diversity communities everywhere; for example, the nettle patch at the bottom of the author’s garden. By exactly the same token, even the densest climax deciduous forest is far from perfectly homogeneous. Each tree is a patch surrounded by other entities. The important aspect of the mammoth steppe, then, is not that it was a mosaic, rather than a zonated, ecosystem; one could call it, or elements within it, one or the other, depending on how one chooses to look at it. To restrict ‘zonation’ to clustering related to external abiotic constraints such as climate and altitude does not help; forests exert some controlling influence on

3. Patch, Matrix, Corridor. In Guthrie’s scheme the mosaic consists solely of patches, clusters of vegetation types, mutually contiguous and with undefined but unproblematical boundaries. This is an incomplete account of the clustering of entities in a mosaic. Forman identifies three categories of mosaic spatial element. The patch is a spatially bounded entity with a high degree of internal homogeneity. Patches, however, do not often cluster tightly together like cells in a honeycomb, at least as far as non-agricultural landscapes are concerned. More frequently they are scattered, like a leopard’s spots, in an expanse of ‘background’ vegetation, the matrix. The nature of the matrix is an important factor in delimiting the boundary of the landscape; a change in matrix can be thought of as a transition from one landscape to another. The defining matrix of the mammoth steppe is grass. It is not necessary that the matrix constitute a greater proportion of the land surface cover than the patches or

35

than any particular patch type; it is defined by its continuity.

wider landscape, particularly since many species depend upon the exploitation of disturbed patches.

Mosaics may also contain corridors, which can be defined as generally elongated patches frequently connected with geographical features such as ridges, steep-sided slopes and watercourses. Corridors are important in that they are simultaneously habitats in themselves, but also conduits when flows of organisms, nutrients or energy move through them. This renders them important factors in migration, recolonisation and the connecting of distanced patches. They may also be barriers if they inhibit movement across them, and sources if they harbour creatures that exploit the surrounding matrix or patches.

Patch size alone, however, does not determine the landscape’s capacity to support a diverse community of species, nor its resilience under perturbation. Landscapes with a high number of small patches are resilient since the risk of catastrophe is spread in space, and because the high incidence of contact edges in such landscapes promotes multivalent ecological links which enable localised disturbance or catastrophe to be bypassed and landscape integrity preserved. Pleistocene human populations can be thought of as a matrix species that moved from patch to patch in a fine grained patch scatter. The role of patch size in their ecology is problematical. The buffer against extinction that large patches provide may have been an important structural condition for mammoth steppe persistence and resilience and therefore an essential landscape feature for human groups, even if large woodland patches were in themselves difficult for people to exploit. Large patches may also have been indispensible as core habitats for species such as preferential browsers which could be hunted or scavenged in the matrix or in more open patches.

It is important to emphasise that all of these spatial elements of mosaic landscapes also constitute ecosystems in themselves, with their own internal patterns of spatial elements. Since the landscape is defined in part by its scalar match with human perception, a mosaic distribution of ecosystems may or may not be a landscape, depending on human land use. The spatial arrangement of these mosaic elements can also be understood in Gibson’s terms as a network of objects, openings and paths that, together with landform features, confer upon an environment its shape.

5. Patch Shape. The configuration of patches is of considerable ecological importance. A patch may be large in terms of its area, but if it is very elongate, or if it is convoluted with lobes projecting into the matrix or adjacent patches, it will nevertheless have a high edge:interior ratio. It may therefore be unsuitable as a core habitat and support few interior species, even if the dominant vegetation within the patch is otherwise favourable for them. The organisms within elongate and convoluted patches, by virtue of the patch’s high edge:interior ratio, tend to be in a state of close proximity to the surrounding landscape. Such patches therefore promote interaction and flow between the patch and other elements in the mosaic. This applies not only to organisms, but also to nutrient and energy flows. Dendritic or labyrinthine patch shapes are often associated with river systems and share those properties of elongated and convoluted patches that derive from a high edge:interior ratio. In addition they frequently afford transit or direct movement through the landscape and may be regarded as corridors.

4. Patch Size and Number. There are complex relationships between the size and number of patches in a mosaic landscape, and the diversity and ecological character of the species they support. Large patches possess a greater extent of interior relative to edge; the smaller the patch, the more important the role of the edge becomes. Small patches do not support species that are restricted to the interior, although they will support species that are confined to edges; therefore they generally do not contain a high diversity of species. They can, however, have an important role as ‘stepping stones’ for the movement across the landscape of species that target particular vegetation patch types. The probable origin of hominid bipedalism as an adaptation to moving across open country between woodland stands would suggest that humans have a long ecological history as a stepping stone species. Large patches generally support larger populations and exhibit greater species diversity, and may be core habitats for wide-ranging vertebrates. They also tend to favour a greater diversity of edge species than small patches, especially of those species which range from edges to exploit the patch interior and adjacent patches and matrix. Patch persistence in the face of perturbation and catastrophe also increases with patch size through incorporation, and this confers upon large patches an important role in maintaining species diversity in the

Compact patches have a length:width ratio close to unity, and few convolutions. This minimises the edge:interior ratio and confers upon compact patches some of the properties of large patches, for example high species diversity, limited rate of interaction with the surroundings and some resistance to disturbance. Although convoluted and dendritic patches certainly occur in flatlands whether one is considering a lowland or an extensive

36

high plateau – such landscapes tend towards compact patches whereas the topography of broken and steeply-cut valley landscapes promotes elongated and convoluted patch shape. This has important implications for the ecological contrast between lowlands and broken uplands in the Pleistocene. The tendency towards high edge:interior, high interaction rate patch shapes in steeply-cut valleys and broken terrain means that, in these landscapes, limited within-patch species diversity and vulnerability to catastrophe and disturbance combines with high levels of interaction to reduce ecosystem resilience. Perturbations are less well buffered than is the case with compact patches, and reduced species diversity limits the potential for the re-routing of interactions following disturbance or perturbation. With some notable exceptions, such as deciduous trees, seed and regeneration sources may be less likely to persist in broken uplands and the conditions for mammoth steppe resilience and persistence may not be met. This may be reversed if the patches, although elongate and convoluted, are large, as is the case with gross altitudinal zonation, but this in turn coarsens mosaic grain and increases resource distanciation.

Patterns of Settlement in Pleistocene Europe The Geography of Occupation in Time and Space 1: Earlier than 200 Kya. The powerful bias of Lower Palaeolithic occupation towards the west of the European continent is well known. The following review of the evidence reveals just how strong this bias is. For the purposes of this review, the term ‘Lower Palaeolithic’ has not been used. Rather, the occupation of Europe prior to 200 kya is taken as a convenient unit of analysis. It has become clear that classificatory schemes applied to lithic industrial typology and technology offer no clear basis for the definition of a Lower-Middle Palaeolithic boundary. ‘Middle Palaeolithic’ industries such as those at High Lodge, England (Ashton and McNabb 1992), La Cotte de St. Brelade, Jersey (Callow 1986) and Combe Capelle Bas in the Périgord (Dibble and Lenoir 1995) predate the Eemian interglacial, in the case of High Lodge by more than a quarter of a million years. Equally, handaxe and biface industries dating from the Upper Pleistocene are common in central and eastern Europe (Bosinski 1967; Valoch 1968a). One may add, on the basis of the author’s personal examination of the material, that the final Middle Palaeolithic assemblage from Level 1A at Scladina Cave, Belgium, would not look out of place in a central European Lower Palaeolithic biface-free small flake industry. The choice of 200 kya as a boundary for this review is to an extent arbitrary; it might perhaps have been placed at the Middle-Upper Pleistocene boundary, 130 kya. There are, however, temporal patterns in the geography of Pleistocene human occupation in Europe that are most clearly visible if site distributions before and after 200 kya are compared.

In lowland mosaics, however, compact shape promotes resilience. Compactness counters the tendency of small patches towards high edge:interior ratios so that small, compact patches may combine high species diversity and resistance to disturbance and catastrophe with high levels of interaction, thereby maximising resilience and productivity. This would be enhanced further in a fine grained mosaic in which large numbers of small patches were in close proximity and interactions between them only minimally inhibited by distance. The conditions for a persistent, resilient, diverse and productive mammoth steppe are closely approached in landscapes of this spatial character, although this may be compromised if patch size increases and patch number declines.

Since the purpose of the review is to assess the evidence for the presence or absence of human occupation in central and eastern Europe before 200 kya, the distribution of sites in France, Spain, Britain and Italy has not been considered. The systematic settlement of these regions through glacial cycles in the period before 200 kya has been noted already. Sites from the rest of Europe, however, were considered according to the following criteria:

Of course, it should be recognised that the spatial mosaic dynamic typical of lowland fine grained mosaics may also apply in extensive upland plateaux. The problems for human populations here are firstly that the plateau must be accessed from the lowland, and this may be inhibited if altitudinally-zoned elongate patches inhibit passage across them, as dense thickets may. This should not be overemphasised since access to plateaux is usually afforded by corridor patches associated with stream and river drainage systems. Coarse grained, large patch approaches to plateaux may pose more of a barrier to access, but the greatest obstacle to plateau occupation may be a change in matrix relative to the lowlands. Adiabatic precipitation may promote a woodland matrix, even when the lowlands are semi-arid open grasslands (Rossignol-Strick and Planchais 1989) or the plateau may have a heather matrix, for example. In both cases, plateaux may have been of value primarily as landscapes supplementary to the lowlands for both humans and grazers in the Pleistocene.

1. Undoubted authenticity of the lithic artefacts – sites or horizons where claims for the artefactuality of Middle Pleistocene stone flakes have been challenged were excluded from the catalogue. There are several such sites that figure prominently in the literature on the earliest occupation of central and eastern Europe, including Prezletice (Fridrich 1989) and Beroun (Fridrich 1991) in Bohemia, Kärlich A and BA (Würges 1986) in the German Rhineland, and Stránská Skála I (Musil and Valoch 1968) and Mušov (Valoch and Zeman 1979) in Moravia. Most involve claims for great antiquity but have failed to gain acceptance as anything other than naturally flaked pieces (Roebroeks and van Kolfschoten 1995b).

37

exclusively used archaeologically invisible perishable materials for artefact manufacture, one is forced to the conclusion that the human presence in the climatically continental regions of Europe was exceedingly thin before the late Middle Pleistocene. It is relevant that the greater number of occupation traces known are found in Germany, in view of Rousseau’s evidence that the boundary between oceanic and continental climatic provinces oscillated in such a way as to periodically incorporate the Rhineland at least in the western oceanic province (Rousseau et al 1990). If this was indeed the case then Germany would have been subject to oceanic climate for significant stretches of time before reverting to climatic continentality. Occupations there may relate specifically to these oceanic episodes.

2. Number of artefacts – sites or horizons that have produced fewer than 100 lithic artefacts, including all debitage, were excluded. A significant number of sites and horizons in central and eastern Europe have yielded very small collections of undisputed artefacts and are indeed evidence of human presence. These include Červený kopec (Svoboda et al 1994) and Růženin dvůr (Valoch 1977) in Moravia, Šandalja I in Croatia (Malez 1976), Wangen (Toepfer 1968), Kärlich BB and G (Würges 1986, 1991) and Kochstedt (Toepfer 1970) in Germany. However, the comparison with western Europe, where much larger assemblages from before 200 kya are common, would be diminished if all such impoverished collections were included. 3. Date – only sites and horizons unequivocally attributable to the period before 200 kya were included in the catalogue. To include assemblages of uncertain date would be a source of confusion at best, and factual error at worst. This condition excludes a considerable number of isolated surface finds, mostly handaxes, from all over central and eastern Europe; the complete impossibility of assigning reliable ages to them renders them little better than useless as evidence of Middle (not Upper) Pleistocene human occupation.

The pattern of settlement in Europe before 200 kya also clearly reveals that human groups tended to avoid upland regions with broken terrain. None of the sites shown in Figure 3.2 are located in the southern German uplands, and none are in the Carpathians, the Moravian Karst or the Balkans. It is of the greatest importance that this pattern is repeated also in western Europe. The great preponderance of French sites over 200 kya in age are at low and moderate elevations in the terraces of the northern rivers, particularly the Somme and the Seine. There is no occupation in the Massif Central before the Middle Palaeolithic (Raynal et al 1995), but there is a small number of Périgordian cave and rock shelter sites with Middle Pleistocene horizons. At La Micoque, Pech de l’Azé II, and Combe Grenal, the horizons are penultimate glacial in date and therefore younger than 200 kya, whereas at Fontéchevade the ‘Tayacian’ artefacts are in poor condition and of doubtful status. At Abri Vaufrey, though, a typologically Middle Palaeolithic assemblage lies beneath a stalagmite floor dated to 200 kya by U-Th (Gamble 1986, 149).

The results (Table 3.1 and Figure 3.2) are revealing. In the whole of continental Europe east of the Rhine and the Mediterranean east of Italy, the review identified only twelve sites that meet the above criteria. Of these, nine (Memleben, Schöningen 12, Schöningen 13, Bilzingsleben II Travertine, Ehringsdorf, Markkleeburg, Wallendorf, Ariendorf and Bad Canstatt) are in Germany, at the westernmost edge of central Europe. In addition, an age of less than 200 kya is just within one standard deviation of the 225 kya radiometric date obtained at Ehringsdorf. Only Bilzingsleben II, Schöningen 13, Vértesszöllös and Korol’evo can be regarded as major sites. There are more in southern England alone (Wymer 1999). It is true that there are abundant artefacts at Petralona Cave, plus of course a hominid skull of the first importance (Stringer 1974). The site clearly represents good evidence of a significant and persistent local human presence in coastal Greece in the Middle Pleistocene, but problems with the quality of publication and difficulties in establishing the age of the deposits with reasonable precision diminish its archaeological value. Account must also be taken of other sites where human remains of Middle Pleistocene age lacking association with any (or at least convincing) artefacts have been recovered in the study area, notably Mauer and Steinheim (Howell 1960) in Germany. Kartstein, also in Germany, is another travertine site with abundant Middle Pleistocene artefacts (Brunnacker et al 1982), but they are cemented in the travertine matrix and have not yet been recovered for examination.

There are, however, very significant upland occupations in Italy at this time (Mussi 1995). Possible reasons for this are discussed below. Petralona is also a cave site, but it is a low altitude coastal cave not located in a broken karstic upland. The Geography of Occupation in Time and Space 2: Later than 200 Kya. The Middle Palaeolithic, taken as beginning at 200 kya, marks a remarkable shift in European human settlement. From around 200 kya there is a sustained eastward movement of population into regions of climatic continentality. Late Middle Pleistocene occupations are known at Hannover-Döhren, Hundisburg, Reutersruh and Weddersleben in Germany; Bohuslavice, Odrou and Polanka in Silesia; Piekary IIb in southern Poland; and Bečov and Stvolínky in Bohemia (Svoboda et al 1996). The Middle Pleistocene occupation of the Dordogne may well be contemporaneous with this as discussed above.

Unless one were to argue that human populations in central and eastern Europe before 200 kya almost

38

REGION/SITE

DATE *

REFERENCES

Germany Memleben

ca 500 kya

Mania 1984

Schöningen 12

>400 kya

Thieme & Maier 1995

Schöningen 13

>400 kya

Thieme & Maier 1995

Wallendorf

MIS 11 – 430-360 kya, or MIS 9 – 340-301 kya

Mania 1988

Ariendorf 1

MIS 8 – 301-242 kya

Turner 1991

Markkleeburg

MIS 8 – 301-242 kya

Mania & Baumann 1980

Bilzingsleben II, Travertine

350-320 kya (U-Th) 414-280 kya (ESR)

Schwarcz et al 1988; Mania 1995

Ehringsdorf

225±26 kya (ESR)

Cook et al 1982

Bad Canstatt

MIS 7 – 242-186 kya

Wagner 1984

MIS 8 – 301-242 kya (OSL)

Foltyn et al 2004

350-245 kya (U-Th)

Kretzoi & Dobosi 1990

VI: 360±50 kya (TL); VII: 650±90 kya (TL) VIII: 850±100 kya (TL)

Gladilin 1989

Poland Rozumice C Hungary Vértesszöllös Ukraine Korol’evo VI, VII, VIII

Table 3.1. Sites older than 200 kya from central and eastern Europe. * Stratigraphically derived dates unless otherwise specified.

Figure .3.2. Distribution map of sites listed in Table 3.1.

39

From the beginning of the Upper Pleistocene at 130 kya, occupation intensifies in Poland (e.g. Raj, Zwolen, Zwierzyniec); Moravia (e.g. Kůlna, Předmostí); Slovakia (Gánovce); and Hungary (e.g. Büdöspest, Solyomkuti, Subalyuk) (Svoboda et al 1996; Féblot-Augustins 1993). The earliest traces of Palaeolithic occupation in the Upper Danube (Schambach, Sesselfelsgrotte, Bockstein) are post-Eemian in date (Müller-Beck 1988), although there is disagreement as to whether the lowest archaeological horizon at the Sesselfelsgrotte is of Eemian or early Würm date (Richter 1997).

because the topographically-forced nature of the spatial arrangement of patches in such landscapes could rarely deliver proximity, diversity and resilience to the same degree as the lowland steppe. This was not an iron law, as the occupation of the Italian Apennines in the Middle Pleistocene demonstrates. Here the palaeoenvironmental record clearly shows that human settlement was restricted to Apennine landscapes with fine grained vegetation patch clustering and that where local environments were not diverse, as in the Riano lake basin, human occupation did not develop (Mussi 1995). Clearly regional factors specific to the Apennines in the Pleistocene, possibly vulcanism, promoted a persistent fine grained mosaic in a broken upland landscape. It was not altitude or broken terrain per se that inhibited human occupation before 200 kya, but rather the temporal weakness and coarse grained character of the of the vegetation mosaic that broken topography promotes. Where local conditions mitigated this and resulted in a fine grained shifting mosaic in broken uplands, as in Italy, then human occupation was perfectly possible.

Occupation also originates no earlier than the last interglacial in the Meuse Basin, Belgium, e.g. the Grotte Scladina (Straus and Otte 1995), and in Romania, for example Boroşteni – although the radiocarbon dates imply an even later occupation (Mertens 1996; Honea 1986). Neanderthals appear to replace early Modern Humans in the Levant from the beginning of the first Würmian-Weichselian pleniglacial at 74 kya. At 40 kya Neanderthals reach highly continental Uzbekistani central Asia (Vishnyatsky 1999).

Raw material transport patterns for Dordogne Lower Palaeolithic sites, all attributable to the Middle Riss in the French terminology, illustrate this point (Figure 3.3). In the valleys of the Tarn, Aveyron, Dadou, Agout and Fresquel rivers, all tributaries of the Garonne, raw material was obtained exclusively from the lowland Garonne flood plain and carried upstream along the tributary corridors into the plateau. The sites of lithic discard in the tributary valleys almost all occur at elevations of below 200m; those at higher elevations are located only on the plateau edge. There is no evidence in the raw materials for any crossing of the plateau (FéblotAugustins 1997b). All connections between raw material source and the point of discard can be parsimoniously routed directly along the tributary river valley corridors, which were in fact extensions or lobes of the lowland plain ecosystem. In effect, for Lower Palaeolithic hominids in this region, the boundary between valley bottom and plateau was the boundary of their landscape, an edge that afforded access to supplementary resources. The plateau itself, it seems, lay outside the hominid landscape.

Not only does the systematic occupation of climatically continental regions begin after 200 kya, but so does the systematic occupation of caves. The Dordogne and Levantine caves have tended to dominate the thinking of Western archaeologists, but the sites in the newlycolonised areas just mentioned are located largely in caves. The Cantabrian mountains of northern Spain may not have been occupied before 200 kya (Raposo and Santonja 1995). The relevance of this lies in the typical geomorphology and topography of the regions in which caves occur, that is in broken karstic uplands dissected by river valleys, often steeply cut. These are precisely the kinds of upland topography that seems to have been almost entirely avoided before 200 kya. In terms of tolerance of both high amplitude intra-annual climate cyclicity, and of persistent, topographically-forced spatial distanciation of resources, the contrast between the geography of human occupation in Europe before 200 kya and that which followed is stark. This in turn implies a transformation in human ecology. Seasons and Millennia

The primary controlling factor barring the settlement of eastern Europe, however, was not the absence of a lowland mammoth steppe, which appears to have been well developed in this region. The factor specific to the east that matches the persistence of human absence there in scale terms is climatic oceanicity and consequent high levels of seasonality. The corrollaries of this are long, harsh winters and high levels of temporal heterogeneity in resource distribution - a seasonal environment is coarse grained on annual time scales just as a large patch or ‘zonated’ landscape is coarse grained in space. This would explain, as suggested earlier, the status of Germany as a semi-settled region, given its vulnerability

What factor unites the occupation of Europe during the era of apparent settlement stasis before 200 kya? As suggested earlier, it appears to have been a strong coupling of human ecology with fine grained mosaic environments. The palaeoenvironmental contexts of Middle Pleistocene occupation invariably seem to have been landscapes of diverse vegetation patches in close mutual proximity. The shifting mosaic, lowland character of this fine grained landscape underwrote the diversity of flora and fauna, and was itself underwritten by the persistence of Dansgaard-Oeschger scale climate perturbations. Broken upland landscapes were barriers to human occupation in the Middle Pleistocene precisely

40

Figure 3.3. Lower Palaeolithic raw material transport patterns in the valleys of the Garonne, south west France, and some of its tributaries. After Féblot-Augustins 1997b, Figure 23

41

to change in the levels of continentality to which it was subject in the Pleistocene.

can be made in an hour, carried for a few days and then discarded. Even distances of 100 km over which stone raw material very occasionally moved in the Middle Pleistocene (Féblot-Augustins 1997a) can be covered in three or four days’ walk.

The precise nature of the seasonality-coupled barrier to occupation is difficult to pinpoint. It may have been a coarser mosaic grain controlled by a shorter growing season, with consequent loss of resource proximity, as Gamble has suggested (see discussion above). This is quite impossible to demonstrate empirically given the current state of development of palaeoenvironmental reconstruction. Even if the mammoth steppe in climatically continental Europe was somewhat more ‘loosely woven’ than was the case to the west, that may not in itself have inhibited human settlement. It may have been the negative impact of a coarse grain on a key resource that excluded Middle Pleistocene populations from seasonal environments. The obvious candidates are prey ungulates, but settlement may just as easily have been limited by high distanciation of key vegetal resources, possibly roots that provided essential vitamin C in long winters when other plant food was scarce.

The social transmission of knowledge, in contrast, seems to have conferred on the bodies of knowledge so transmitted, Middle Pleistocene projects for living, an enormous persistence in time. The occupation of regions, landscapes and even sites through multiple glacial cycles, with little or no visible re-ordering of mobility strategies or resource recognition and no transformation in lithic technologies that could be attributed to directional, adaptive accommodation of landscape perturbation into new schemes of knowledge, is testimony to the powerful damping of input signals from the scale domain of knowledgeable action – i.e. practices or behaviour – as they were delivered upwards to the constraining, structural level of the project by the processes of knowledge transmission then in operation. In the Middle Pleistocene, the relation between knowledgeable action and socially transmitted knowledge, between agency and structure, was not the powerful symmetrical mutualism described by Giddens, but a strongly asymmetrical constraint of practice by structure. Human action remained coupled to, and constrained by, landscapes whose heterogeneity on time scales of days and weeks and spatial scales of kilometres was fine grained and which persistently retained that spatial and temporal fine grain for tens of millennia. Loss of that fine grained spatiotemporal heterogeneity, either in seasonal environments, in coarse grained uplands or through the overwhelming of mosaic resilience at the peaks of glacials and interglacials, rendered occupation untenable, precisely because the social repertoire of knowledge was separated from the grain responsiveness of perceptive individual action by a scalar gulf that placed each outside the other’s dynamic scale range.

But it is not necessary to imagine a coarser grained mosaic in the continental centre and east of Europe to understand the weakness of its Middle Pleistocene human settlement. The apparent reliance on a fine grained steppe with high levels of resource spatial proximity necessarily implies a reliance on high levels of resource temporal proximity. Given a characteristic rate of movement through the landscape – putting aside for now the resistance to movement afforded by closed landscapes – the time required to travel between synchronic resources is largely governed by their distanciation in space. Of course, distant key resources need not pose a problem providing that time scheduling can ensure their procurement without compromising the supply of other key resources (Binford 1979); the controlling factor may therefore be the temporal reach of knowledgeable action in the landscape. This is akin to Binford’s concept of planning depth except that it need not imply the conscious design and execution of a plan. Tacit and intuitive knowledge can equally well afford a degree of temporal scope to knowledgeable action. It seems, however, not to have been of a sufficient degree before 200 kya to overcome the scale of the difference between continental summer and winter in terms of affordance distribution in time. Middle Pleistocene life appears to have been lived very much in the here-and-now, but human survival in highly seasonal landscapes may depend on acting in the here-and-now in a way that implies the there-and-then. Otherwise, a long hard winter is an annual catastrophe, as for the hypothetical evergreen oak discussed earlier.

The Middle Palaeolithic and the Integration of the World The colonisation that began around 200 kya of Europe’s broken karstic uplands and eastern seasonal environments, taken together with the tendency, more marked from the onset of the Upper Pleistocene, for lithic practices to show structured change in time, is an unequivocal indication of a scale transformation in the ecology of human action and knowledge in the world. In inhabiting seasonal landscapes Pleistocene humans were exercising a broader spatiotemporal scope in their knowledgeable action than hitherto. It is important to recognise that survival in a seasonal environment need not necessarily imply an explicit concept of ‘the year’ or a consciously scheduled annual round within the landscape. If that were true, there would be few animals living in such environments other than those whose physiologies are so coupled. It does, however, imply that the temporality of knowledgeable action had expanded

Human knowledgeable action in the Middle Pleistocene, then, seems to have operated on very short time scales. On scales of minutes, hours and days Middle Pleistocene hominids were resourceful and skilful, as has recently been powerfully demonstrated by recovery of the Schöningen spears (Thieme 1997). But wooden spears

42

sufficiently to bridge crucial off-season gaps in resource availability. This could be achieved by predicting and intercepting herd movements, or by the temporary fissioning of task-specific groups, for example. Both of these devices involve knowledgeable action in the hereand-now in which action in the there-and-then is implicit.

detached by force, reveal sharp edges that afford cutting, piercing and scraping. But the identification of the planes of attachment, and thus of the force and direction of hammer blows, lies in the dialogue between the knapper’s social eye and the properties of the material. Just as the tool is a product of this dialogue, it is itself implicated in a further dialogue with the material toward which it is directed. The matrix of constraining socially mediated knowledge structures, individual goal directed action and the material properties of the world is the nexus of practice, and it is only in practice that knowledge is transmitted. Changes in the material world – available raw material, distribution of resources etc. – change the terms of the dialogue and promote new practices. Given the mutuality of organism and environment, then, transformations in the properties of the world and of its spatiotemporal character transform affordances and thus challenge the project for living. As an accommodation of a perturbation, this can be thought of as adaptation at the level of the social individual. Knowledgeable action, as an appropriate application by the human individual of a socially transmitted repertoire of knowledge, is always liable to generate new or innovative practices when transformations in the world push at the boundaries of the repertoire’s effectiveness. The extent to which new or innovative lithic practices become institutionalised in social knowledge and thus impact upon constraining structures of knowledge – that is, the extent to which the scale relation between agency and structure is symmetrical – will therefore be expressed in the extent to which changes in knowledgeable technical action, in industrial tradition, follow the pace of environmental change.

This is most clearly visible in a comparison of FéblotAugustins’ plots of raw material movement in the late Middle Palaeolithic (i.e. Upper Pleistocene) of the Garonne Valley with the Lower Palaeolithic patterns discussed above. In Figure 3.4, the transformation is clear. Sites tend to be located at elevations of around 200m, on the valley-plateau ecosystem boundary, although some peripheral sites are found on the high plateau. Raw material moved both along river valley corridor extensions of the lowland ecosystem, but also systematically across plateau tops, connecting river valley and plateau in a network of movement. (FéblotAugustins 1997b). The various connections shown are not intended to be seen as synchronic, but they do indicate that the plateau top had been incorporated into the landscape in the Upper Pleistocene. This is a crucial point, because the plateau is different from the lowland. Its patch grain is different in time and space, its matrix may be different. Its incorporation into a landscape together with the lowland plain and tributary river valleys represents the incorporation of difference, and an expansion in the scale of human grain responsiveness. This is not limited to the Périgord. At the Trou de l’Abime, Couvin, Belgium, raw material occurs at least 50km, perhaps as much as 80km, from its sources in the north and must have been carried across the axial River Meuse and up a tributary corridor to the cave site in the Ardennes (Ulrix-Closset et al 1988). In central Europe raw material moved preferentially between valleys and over plateaux, rather than both between and along the valleys, as was the case in the Périgord (Féblot-Augustins 1993).

It is not easy to spot this in the Middle Pleistocene industries of Europe. Former typological-cultural evolutionary schemes for chronological subdivision of the European Lower Palaeolithic handaxe industries, for example ‘Abbevillian’ and ‘Chellian’, have fallen into disuse as developments in dating techniques have demonstrated the weakness of patterned directional development through time in Middle Pleistocene knapped stone industries (Gamble 1986; Bosinski 1995). Detailed typological classifications of handaxe types (e.g. Roe 1968) have succeeded only in imposing spurious structured difference where none exists. The overwhelming impression is of a highly persistent social repertoire of technological knowledge that found different expressions at particular times and places according to raw material constraints, as with pointed and ovate handaxe forms (e.g. M. J. White 1998) and allometric relations of size (Gowlett and Crompton, 1994). Where persistent regional differences in handaxe form do exist, the distinctions are only visible statistically and on very gross spatiotemporal scales (Wynn and Tierson 1990). This is not to deny the existence of apparently idiosyncratic aspects of Middle Pleistocene handaxe form such as the ‘S’ and ‘Z’ twists, but if they defy explanation through raw material factors, they are

Industrial Variability Stone knapping techniques are the best proxy palaeolithic archaeologists have for past practices, and for the relationship between structure and practice. Each flake was struck in an instant by an individual with and for a purpose. At the level of the individual piece in functional context a teleology of the artefact is justified. At the same time the knapper is not free to work the nodule entirely as he or she chooses. As a member of society the knapper will have been exposed to the ways of knapping, striking, and of holding the core and the hammer used by his or her peers, much of which he or she will have absorbed into tacit or intuitive knowledge mediated by habit. The social repertoire of knapping techniques, of ways of relating purposefully with the stone, constrains the knapper’s choices. In Gibson’s terms, a core may be considered as a series of attached objects that, when

43

Figure 3.4. Late Middle Palaeolithic raw material transport patterns in the valley of the Garonne After Féblot-Augustins 1997b, Figure 39.

44

However, despite these caveats, there are apparently real temporal industrial trends within regions in the Middle Palaeolithic, the best known of which are Bordes’ Mousterian facies in south west France. Statistical analysis has shown the facies to possess discrete reality relative to each other (Mellars 1967; Callow and Webb 1981); they seem also to be chronologically patterned (Mellars 1967, 1969; but Ashton and Cook 1986). In fact significant industrial variability temporally structured on timescales of millennia or tens of millennia is a notable property of the European Middle Palaeolithic record. Although Bosinski’s chronology for the Middle Palaeolithic industrial succession in Germany (Bosinski 1967) has not stood the test of time, as will be shown in Chapter 4, clear Middle Palaeolithic temporal successions are apparent at many central and eastern European sites. At the Sesselfelsgrotte, Bavaria, the Eem/Early Würm levels, with a scraper-rich, low Levallois Mousterian, lie beneath a ‘Micoquian’ industry rich in bifacially worked pieces (Weiβmüller 1995a; Richter 1997). Comparable variability through time occurs at, for example, Kůlna Cave in Moravia (Valoch 1988) and the open air sites Königsaue in Germany (Mania and Toepfer 1973) and Ripiceni-Izvor, Romania (Păunescu 1993).

best understood either as stochastic variation, or as the expression of micro-scale events, such as local extinction followed by recolonisation with technological ‘founder effect’. One is forced to the conclusion that the Acheulean’s most striking feature is its persistence as an idea for 1.5 Myr in a vast province stretching from the Cape of Good Hope to Britain to India. Within Europe the major industrial opposition is that between the Acheulean and the non-biface industries such as the Clactonian and the small-tool industries from Bilzingsleben and Vérteszöllös, and this dichotomy persists through the Middle Pleistocene with no indication of any intra-regional trend in time from one to the other, at least before 200 kya. From around the late Middle Pleistocene, however, there are trends towards the regionalisation of industries. Blade industries employing techniques close to or identical with Upper Palaeolithic prismatic core reduction occur in some regions of Europe, notably southern Poland (Svoboda et al 1996) and the north Rhine, Picardie and the Pas de Calais (Conard 1990). Biface fabrication declines markedly in France, Spain and Italy but becomes an important feature of industries in parts of central and eastern Europe (Bosinski 1967; Allsworth-Jones 1986).

It is true that this structured variability in time of Middle Palaeolithic industries is not so clearly defined as is the Upper Palaeolithic industrial sequence, but this may be a consequence of typological methodologies that emphasise the fossile directeur. Type fossils underwrite the wellestablished Upper Palaeolithic culture-industrial chronology of Europe in which the AurignacianGravettian-Solutrean-Magdalenian sequence, and its regional variants, occupies just 30,000 years or so of Pleistocene time. They are, however, much less important (although certainly not entirely absent) and less clearly defined in the Middle Palaeolithic. Typological systems in which the fossile directeur plays an important role – including the Bordesian system (Mellars 1992) - are thus vulnerable to the charge that they are insufficiently sensitive to other dimensions of lithic-industrial variability.

The comparison of rates of lithic industrial change through time, it must be said, is fraught with difficulty since one is frequently not comparing like with like. This is especially true of comparisons between assemblages across the Lower-Middle Palaeolithic divide. How is one to compare, say, a handaxe-rich Acheulean assemblage with a Denticulate Mousterian? The problem resides in establishing points of comparison that are not implicitly arbitrary and subjective, and thus liable to undermine the comparative principle. Indeed, there is an urgent requirement for the development of methods of lithic assemblage analysis that permit quantitative comparisons between assemblages but which avoid the reliance on subjective attributions of pieces to arbitrarily defined type categories that bedevils the Bordesian system. Furthermore, even if such an ideal lithic-analytical methodology were to demonstrate an acceleration in industrial change through Pleistocene time, one would have to consider the possibility that the acceleration may be more apparent than real, a consequence of poor resolution in older sites and horizons damping out real industrial change. At the same time the stacking of artefact-bearing deposits and their vertical division into stratigraphically defined horizons in stratified sites – which, as discussed above, tend strongly to date to less than 200 kya – may literally make ‘vertical’ change in time appear to be a pervasive feature of the Middle Palaeolithic record. Finally, it remains the case that in all regions of Europe and what became the wider Neanderthal world, typologically and technologically unremarkable Mousterian assemblages occur throughout the Upper Pleistocene prior to the Upper Palaeolithic.

There is, then, good reason to believe that, as expressions of socially transmitted knowledge, lithic industrial practices in the European Middle Palaeolithic were liable to undergo systematically structured change on much shorter timescales than is apparent in the Lower Palaeolithic record. It is true that, even if one accepts the reality of the Bordesian Mousterian facies and accepts that they are chronologically patterned, it remains difficult to match the shift from Ferrassie to Quina to MTA, or from Taubachian to Micoquian at Kůlna, with any wavelength of climatic or environmental wavelength of perturbation visible in any proxy record. Nevertheless, although one would not expect to find perfect correlation with any single scale of proxy in a complex, multi-scale realm such as socially transmitted knowledge, there is a tendency in the Upper Pleistocene towards significant chronological variability on ecological timescales of

45

millennia, indicating that the mechanisms of transmission of knowledge operated at similar process rates to Milankovitch scale climate variation. Structured variability in time of this character is very difficult to see in the Middle Pleistocene archaeological record, and its emergence coincidentally with the transformation in settlement noted earlier suggests that the two phenomena are organically linked.

rate of change in and regionalisation of technological practices. The specific task to be undertaken in this dissertation, then, is an examination of the late Middle Palaeolithic leaf points of the Upper Danube region from this perspective. This entails not only an approach founded on a hierarchical ecology of knowledge as developed in Chapter 2, but also an awareness of the history of scale shifts in human coupling with the world revealed in this Chapter. To this end, Chapter 5 will examine the multiple wavelengths and amplitudes of climate and landscape change in the European Upper Pleistocene with a view to understanding the scalar character of transformations in the affording environments of which late Middle Palaeolithic leaf point fabrication, use and discard were component parts. In Chapter 6, late Middle Palaeolithic Upper Danubian lithic assemblages, including some with leaf points, will be analysed in order to characterise leaf point formal specificity as a proxy for the rate of emergence of innovative social-technological practices in that period and in that region. First, Chapter 4 will consider the spatiotemporal distribution of late Middle Palaeolithic leaf points, and the landscape context of the Upper Danube examples in particular, with the aim of identifying the spatial and temporal scale domains to which the leaf point phenomenon relates.

Implications and Directions This picture of the ecology of human engagement with the world in the European Lower and Middle Palaeolithic has been painted with a very broad brush. Nevertheless, it shows that a history of profound transformations in the scalar terms of human ecology commenced at about 200 kya, and that these transformations are visible as scale shifts in the grain responsiveness of knowledgeable action in the landscape and in the process rate of coupling between socially transmitted knowledge and knowledgeable action. This is evidenced in the emergence of new ecological geographies of settlement predicated on an integration of difference over temporal and spatial extents that had defied integration earlier in the European Middle Pleistocene, and in an accelerated

46

CHAPTER 4

THE DISTRIBUTION AND DATING OF MIDDLE PALAEOLITHIC LEAF POINT INDUSTRIES IN EUROPE variable size. The only feature that Obermaier and Wernert regard as universal rather than usual is the convergence of the edges to a point at each extremity, but even this no longer holds given the presence of platform remnant butts on leaf points from, for example, Jankovich and Musselievo, or of apparent hafting notches on a few of the Mauern examples.

Introduction The project for an ecological geography of the Palaeolithic, as developed in Chapters 2 and 3, is comparative in character in that it seeks to map possible changes through the European Lower and Middle Palaeolithic in the relation between the scale domains of socially transmitted knowledge and knowledgeable action. If the Middle Palaeolithic leaf point phenomenon is to be incorporated into this comparison, the spatiotemporal extent of the Middle Palaeolithic leaf point phenomenon must be determined. This Chapter, then, will seek to identify its boundaries in space and time. It should be pointed out now that leaf points as defined below are vanishingly rare in the Middle Palaeolithic of southern and western France, Iberia and Italy (Ringer 1995), and that no long list of sites and assemblages from those regions will be adduced merely to demonstrate this absence.

In the present analysis leaf points are regarded as biconvex, more or less symmetrically so, and as distally pointed, where pointedness entails a convergence of the edges even if the tip itself is rounded (Figs 4.1, 4.2). Pieces whose tip is so rounded as to preclude a convergence of two edges are excluded (Fig. 4.3). Leaf points may or may not be proximally pointed, and the point of maximum breadth might be located anywhere from the base to the mid-point of the length, or even somewhat more distally. The presence of flat scalar retouch on most or all of one surface (the dorsal surface where the point is fabricated on a flake or blade blank rather than on a slab or plaquette) is essential, as is the presence of low edge angles (40,000 for III. This would place the first cold maximum following the last interglacial within the OIS 3 interpleniglacial when in fact it is known to have occurred in OIS 4 (ca 74-59 kya) immediately prior to the onset of OIS 3. Assuming that cycles Ia2 and Ib relate to the two major interstadials within the early last glacial as the excavators suggest, one would expect the pleniglacial to be located between cycles Ib and II. The published chronology for Königsaue therefore depends upon a local first Weichselian pleniglacial maximum much later than is apparent elsewhere. An interpleniglacial, rather than early glacial, date for Cycle Ib is therefore possible and would be consistent with infinite radiocarbon dates of >49,000 BP and >55,800 BP and with more recently-obtained radiocarbon dates of 43,800±2100 BP for Micoquian horizon A and 48,400±3700 BP for Mousterian level B (Hedges et al 1998). In short, there is no conclusive chronology for the Königsaue sequence, and the cycle Ib Micoquian leaf points might date either to a late glacial interstadial older than 74 kya or to sometime within the earlier part of OIS 3 between around 60 and 43 kya.

The bifaces from Grotte du Docteur include 14 handaxes, as well as backed biface knives and bifacial scrapers. Ten whole pieces and five fragments are described as foliates, although most are asymmetrical Faustkeilblätter. Only three pieces might be called leaf points, although one of these (Fig. A2.9.i) has a thick, triangular cross section and is another example of a bifacial limace. This leaves just two claimed leaf points, one of which (Fig. A2.9.ii) is asymmetrical and similar to a Moravany-Dlha point; the other (Fig. A2.9.iii) is very thick proximally. In addition there are two probable leaf point fragments (Fig. A2.9.iv and v). There are no unifacial leaf points or Jerzmanowice points. Grotte du Docteur is therefore the most convincing example among Belgian Middle Palaeolithic sites for a leaf point presence although, as at Spy, the numbers are few and not clearly distinguishable typologically from other bifaces also present in the assemblage. This is a frequently-encountered characteristic of Middle Palaeolithic biface industries, especially those described as Micoquian, in northwest, central and eastern Europe (see Table A3.1).

Another important leaf point industry in central Germany is that from Layer X at the cave site of Ilsenhöhle, Ranis, Thuringia (Hülle 1977; Allsworth-Jones 1986, 68-70). This ‘Ranis 2’ industry consists of 60 tools of which 28 are unifacial leaf points and 25 bifacial leaf points (Allsworth-Jones 1986, 69). The bifacial leaf points are variable in form but the unifacial examples are strongly laminar and conform to the Jerzmanowice type (Figs. A2.16, 17). There are no handaxes, biface knives or any other bifaces of Micoquian type, no cores and virtually no debitage. Chmielewski (1961, 1972), Hülle (1977), Müller-Beck (1966), Campbell (1980), Kozlowski (1983), Otte (1984) and Allsworth-Jones (1986, 1990) are all agreed that the Ranis 2 industry is best thought of as early Upper Palaeolithic. It has been described as ‘Ranisian’ and linked with the Jerzmanowician and the

Northern and Central Germany Unequivocal leaf points in a Micoquian context are known from Königsaue in the former DDR (Mania and

56

British blade points (Campbell 1980, Kozlowski 1983). This attribution is accepted here. Layer X has been dated to the Hengelo interstadial, which, if correct, is consistent with either a Middle or an Upper Palaeolithic age for the industry.

Zotz describes 15 tools out of 239 as leaf points, but Allsworth-Jones (1986, Table A3.1) claims 57 leaf points (of which one is unifacial) out of 145 tools. The reasons for the discrepancies are not clear, although a reclassification of some of Zotz’s bifaces as cores, and of his Blattformen as leaf points, would account for some of the disagreement. Either way, the assemblage contains unequivocal bifacial leaf points (Fig. A2.22) in a probable workshop context, although their formal discreteness from other bifaces is not quite so clear as at Rörshain. The Kösten leaf points are invariably made on flake-blades or plaquettes and, as at Rörshain, most are fragments.

Southern Germany There can be no doubt that the hill country of southern Germany boasts the richest inventory of Middle Palaeolithic leaf points in Europe, although dating remains a problem at many sites. The identification of such industries is also simplified by the relative rarity of early Upper Palaeolithic leaf points in the region. Rörshain, in Hesse in the northern part of this region, is a sandpit site that produced 40,000 artefacts, made predominantly on local quartzite, in the course of excavations in 1965 (Luttrop and Bosinski 1967) and 1973-4 (Campen and Hahn 1975). The archaeological horizon occurred in solifluction deposits, the stratigraphy of which is difficult or impossible to determine and over which the various excavators cannot agree (Luttrop 1970; Bosinski 1973; Campen and Hahn 1975). Of the 1206 tools recovered in the 1965 excavation 331 are handaxes and biface knives (Fig. A2.21) and 118 are thick, elongate and parallel-sided bifacial leaf points (Fig. A2.20) that are quite distinct in form from the other bifaces (Allsworth-Jones 1986, Table A3.1). All but four are made on quartzite. The typological distinctiveness of the leaf points is all the more noteworthy in view of the variation in blanks employed for their fabrication. Hahn has shown that they are variously made on flakes or blades, on tabular plaquettes or on thermal fragments, with quite different reduction strategies applied to each (Hahn 1990). It has been suggested that the site represents a manufacturing workshop, at least for the leaf point component of the industry, in view of the large amounts of debitage, the local outcropping of the material on which the leaf points are overwhelmingly made, and the great preponderance of leaf point fragments (possibly manufacturing rejects) over whole examples (AllsworthJones 1986). Unfortunately dating is not possible, although Bosinski attributes the material on typological grounds to a final ‘Rörshain’ facies of the German Micoquian, immediately preceding the Altmühlian (Bosinski 1967). All that can be said is that, both typologically and technologically, the Rörshain leaf point industry is certainly Middle Palaeolithic.

The most important concentration of sites with Middle Palaeolithic leaf points in southern Germany is to be found in and around the valley of the Altmühl River, Bavaria, which flows into the Danube from the northwest at Kelheim. The term ‘Altmühlian’, coined by Bohmers (1951), has entered into widespread use in the literature as a cultural designation for these leaf point industries. There are also a number of Micoquian sites in which bifacial working is common but leaf points are absent or virtually so. The Middle Palaeolithic of the Altmühl Valley will be considered in more detail in Chapter 6, so discussion will here be restricted to the key sites (see also Tables A3.1 and A3.2). By far the most significant Altmühlian site is the Weinberghöhlen cave complex near Mauern (Bohmers 1951; Zotz 1955; von Koenigswald et al 1974) in the Wellheimer Trockental (dry valley). Bohmers’ 1937 excavation recovered some 33 unequivocal leaf points (30 bifacial and three unifacial; see Figs. A2.23, 24) from his Layer F, plus two handaxes (Fig. A2.25) and 59 other tools, mostly scrapers, of which several bear bifacial retouch. In addition to this ‘Altmühlian’ industry Bohmers identified a ‘Mousterian’, lacking leaf points but with one handaxe and several bifacially retouched scrapers, from the underlying Layers G' and G. Zotz’s 1951 excavation yielded a further 18 leaf points from Layer F and his F2 (equivalent to Bohmers’ G'), Layer G and other parts of the cave system (Figs. A2.26, 27). It should be borne in mind that Zotz was strongly committed to the idea of a Palaeolithic biface ‘phylum’ that led from the Acheulean to the Solutrean, and renamed Bohmers’ ‘Mousterian’ and ‘Altmühlian’ as ‘Praesolutréen I and II’ respectively. It is possible that the leaf points he attributed to his Layer G fell into that horizon from Layer F in the course of excavation. Further work in the 1970s confirmed Bohmers’ stratigraphy (von Koenigswald et al 1974). The industries from both Layers F and G are undoubtedly Middle Palaeolithic, small and, in view of the paucity of debitage and the absence of food refuse and hearths, seem to derive from ephemeral occupations.

The only other site attributed by Bosinski to the Rörshain Micoquian is the open-air site of Kösten, which should be considered together with the nearby site of Schönsreuth (Zotz 1959; Freund 1963) in the Upper Main Valley, in north Bavaria today but historically in Upper Franconia. Again, the site has no meaningful stratigraphy and dating is difficult. The inventory includes artefacts of Middle and Upper Palaeolithic, as well as Neolithic, type. 30 Handaxes and Fäustel are present, and the leaf points possess no attributes that bar a Middle Palaeolithic origin. 99% of the artefacts are made on Kieselschiefer or lydite.

The Mauern leaf points, although variable in terms of form and size, are nevertheless quite unlike any other tools in any industry from the site. They include

57

fragments and both unifacial and bifacial examples, and some are apparently unfinished, but are defined as a group by bifacial working, somewhat denticulated edges, distal pointedness, thinness and biconvexity. They are made on both tabular Plattensilex and on flakes. The unifacial examples are not Jerzmanowice points (Figs. A2.24.ii, 27.ii). Müller-Beck (1974) has placed Bohmers’ Layers G and G' in a cold stage immediately prior to Hengelo, and Layer F in the Hengelo warm event. Layer E, which overlies F without any indication of a sedimentary hiatus, has produced a Gravettian industry, as has Layer C. This is consistent with radiocarbon dates of 28,265±325 BP for the archaeologically sterile intervening warm-period deposits of Layer D and of 29,410±470 BP for Zotz’s Layer C2 on the terrace. An interpleniglacial age for the Middle Palaeolithic horizons is also suggested by correlations drawn between the cave sediments and those in the dry valley outside the cave (Brande 1975).

type. There is certainly nothing like an Altmühlian leaf point present. The Bockstein Micoquian has been placed in the final part of the last interglacial or the very early last glacial by Bosinski, on faunal, palynological and typological grounds. However, Bosinski and his fellow analysts assumed a short chronology for the Upper Pleistocene (Filzer 1969, Abb. 41) that is no longer tenable; it is therefore perfectly possible that the Bockstein Micoquian, which directly underlies a small and undiagnostic Upper Palaeolithic industry, dates to the OIS 3 interpleniglacial and is thus contemporaneous, broadly or strictly, with the Mauern Altmühlian. More south German Micoquian sites with bifaces including leaf points or leaf point-like pieces are given in Table A3.1.

A good case for an Altmühlian leaf point industry can also be made at the Obernederhöhle (Freund 1987), a small cave in the Ziegeltal dry valley on the north bank of the Altmühl some 5 km upstream from its confluence with the Danube. The stratigraphy is unclear, in part because of the sedimentary regime, but also because of the idiosyncratic excavation techniques of the site’s amateur discoverer, Oberneder. Freund’s re-excavation established a working stratigraphy, but it is, to a degree, schematic and the archaeological horizons cannot be regarded as sealed. 278 artefacts, including nine bifaces, six bifacial leaf points (Figs. A2.35, 36) and one Jerzmanowice point, were recovered from levels referred to the Middle Palaeolithic by Freund (1987, Tab. 2). However, Freund points out that Middle Palaeolithic pieces also occur in the Upper Palaeolithic levels, the lower of which produced a second Jerzmanowice point and a bone point. The leaf points are bifacial and morphologically Altmühlian, and bear no formal resemblance to the Jerzmanowice points or to the other bifaces; the latter are also similar to those from Mauern Layer F. The rich fauna is no more than generally Würmian in character and a more precise chronological attribution is impossible. Cores and debitage are rare and there are no hearths, suggesting that, as at Mauern, the cave was occupied only ephemerally, although the number of occupations represented cannot be determined.

Allsworth-Jones (1986, 1990) and Svoboda et al (1996) detail over 30 sites in Moravia from which leaf points have been recovered. The bulk of the assemblages with which the leaf points are associated are unequivocally Early Upper Palaeolithic (i.e. Aurignacian or ‘transitional’ in character (Table A3.4), or cannot be assigned with confidence to either the Middle or Upper Palaeolithic (Table A3.3). This places Moravia at the centre of debates on the nature, extent and origin of the Szeletian (e.g. Allsworth-Jones 1986, 1990; Oliva 1979; Valoch 1990a), a problem which lies beyond the scope of the present work.

Central Europe Moravia

One site that has been classified as Early Upper Palaeolithic, but which deserves attention here, is Bohunice in the Brno basin (Valoch 1976b). The finds, amounting to over 600 tools, occur in a fossil soil overlying a loess horizon some 2-3 metres in thickness, and combine a Levallois-dominated technology with a supposedly Upper Palaeolithic typology. 13 bifacial leaf points, some of which are rather thick (Fig A2.42) are the only bifaces present. 55% of the tools are unretouched Levallois blanks and the laminar index reaches 45%. Endscrapers account for 14%, and burins 12%, of the retouched tools (Svoboda et al 1996, Table 5.5). Czech archaeologists regard this industry as the type for the ‘Bohunician’ of southern Moravia, which they recognise also at Stránská skála II and III, Líšeň, Ondratice and Mohelno (Svoboda 1990). There is disagreement as to whether the leaf points are an intrusive Szeletian element (Valoch 1982) or integral to the industry (Svoboda 1990; Svoboda et al 1996).

In the Lonetal, some 90 km to the south west of Mauern in the Upper Danube of Baden-Württemberg, is a complex of localities at the Bockstein caves. By far the most important horizon is ‘Bockstein III’ in the ‘Bocksteinschmiede’ terrace, from which a large Micoquian assemblage was excavated by Wetzel (1958; Bosinski 1969a,b). The industry is heavily dominated by bifaces, and both cores and flake tools are rare. The bifaces include handaxes, flat handaxes and biface knives and scrapers (Fig, A2.31, 32); five or six approach leaf point form (Fig. A2.33). Once again, these grade into other biface forms and cannot be regarded as a discrete

Radiocarbon dates place the Bohunice fossil soil at between 40,173±1200 BP and 42,900+1700-1400 BP. Valoch suggests a date at the beginning of the OIS 3 Hengelo warm event but Svoboda puzzlingly argues that the dates correspond to the end of the first Würmian pleniglacial maximum. This is quite at odds with a termination of OIS 4 at around 59 kya. However the thick

58

loess horizon, which overlies two fossil soils attributed to the OIS 5 interstadials Odderade and Brørup-Amersfoort, can quite reasonably be placed in OIS 4. This would in turn place the immediately overlying archaeological horizon in the early OIS 3 interpleniglacial, with an age perhaps as great as 60 kya. This age for the artefacts is further supported by Valoch’s contention that they were discarded on the surface of the loess which only subsequently underwent pedogenesis. In this case the radiocarbon dates should be regarded as minimal and the Bohunice leaf points as unequivocally Middle Palaeolithic. All other sites attributed to the Bohunician consist entirely of undatable surface collections, not all of which contain leaf points; even those that do feature leaf points cannot therefore be regarded as unequivocally Middle Palaeolithic and are listed in Table A3.3.

Jerzmanowice (Chmielewski 1961; Allsworth-Jones 1986), industries rich in the Jerzmanowice points named after the site occur in Layers 4, 5 and 6. Jerzmanowice points are excluded from the leaf point type here, as discussed above; nevertheless, they are treated as such in much of the literature, so some attention should be paid to these industries. The ‘Lower Jerzmanowician’ of Layer 6 is radiocarbon dated to 38,500±1240 BP and attributed to the OIS 3 ‘Hengelo’ warm event, while the ‘Upper Jerzmanowician’ of Layer 4, which is separated from Layer 6 by a cryoturbated Layer 5, has been placed in ‘Denekamp’. The latter is therefore too young to be of Middle Palaeolithic origin. In the Jerzmanowician assemblages as a whole there are 10 bifacial leaf points and 69 Jerzmanowice points (Fig. A2.50) out of a tool total of 131; there are also 23 unifacially retouched pointed blades that require only some ventral thinning to qualify as Jerzmanowice points. The remainder of the collection includes one backed lunate but no scrapers, and there are 5 prismatic blade cores. There is very little debitage, so the site appears to have been used – and reused – as a cache. The industry possesses no Middle Palaeolithic features, and the evidence points strongly to an Early Upper Palaeolithic origin. As has already been mentioned, similar points are also important in the Ranis 2 industry and in the Belgian and British Early Upper Palaeolithic. Other Jerzmanowician sites in Poland are listed in Table A3.4.

The most significant Middle Palaeolithic site in Moravia is Kůlna Cave (Valoch 1988, 1995), in the Moravian Karst to the north of the Brno region. Well-controlled excavation has revealed 13 discrete archaeological horizons. The uppermost five are Upper Palaeolithic, but basal Layer 14 is Mousterian, Layer 11 ‘Taubachian’ and Layers 9b, 8a, 7d, 7c, 7a and 6a Micoquian. Layer 7b is archaeologically sterile. Valoch places the Layer 11 ‘Taubachian’ (a small-tool Mousterian lacking flat surface retouch but with some bifacial working) in the stadial following the last interglacial, the Layer 9b Micoquian in the Amersfoort interstadial, Layer 7c in Odderade and Layer 7a in ‘Moershoofd’. An OIS 3 date for the latter layer is consistent with the Layer 7b occupational hiatus if that horizon is placed in OIS 4. It is also consistent with radiocarbon dates for 7a of 38600+950-800 BP and 45600+2850-2200 BP, although these must regarded as minimal. Fragmentary human fossils classified as Neanderthal have been found in the same layer (Vlček 1996). This chronology would place the Micoquian assemblages of Layers 9, 7d and 7c in OIS 5.

Mention should also be made here of Wylotne (Chmielewski 1969), a stratified cave site in the Sąspówka Valley. Middle Palaeolithic ‘MicoquoProndnikian’ industries - with numerous biface knives frequently subjected to distal tranchet or ‘para-burin’ blows to produce what Chmielewski calls a prondnik knife – occur in Layers 7/8, 6 and 5. Layer 7/8 has been tentatively placed in Amersfoort and 5 in Brørup, but this ignores the unity of these events in most of Europe. In fact the industries could date from any time in the Upper Pleistocene before about 40 kya. There are no leaf points, although certain bifaces approach leaf point form.

The Micoquian horizons at Kůlna are characterised by the typical range of bifaces: there are handaxes (Fig. A2.45), flat handaxes, biface knives and biface scrapers, all worked with flat surface retouch, and the biface index is around 10%. Leaf points are both rare and doubtful; ‘foliates’ (i.e. Blatt-typen) amount to only four whole pieces and four fragments in the entire Middle Palaeolithic sequence. All are small, thick and better described as flat Faüstel than as leaf points (Fig. A2.46). Kůlna is therefore another example of Micoquian assemblages from stratified sites in which leaf points as a discrete type are absent but the range of biface form includes a pole that approaches leaf point form in some, but not all, respects.

Slovakia There are no unequivocally Middle Palaeolithic leaf point sites in Slovakia. There are, however, several which defy attribution to either the Middle or Upper Palaeolithic on account of the lack of stratigraphical context, assemblages so small as to be unclassifiable, the possibility of admixture of late Middle and early Upper Palaeolithic artefacts, or combinations of these factors (see Table A3.3). The most important of these sites is Moravany-Dlhá, an open air location in the Váh valley in the mountains of western Slovakia (Zotz 1951; Absolon 1947; Barta 1960; Allsworth-Jones 1986). Unfortunately details of the site have never been adequately published and much of the material seems to have been lost. Zotz found some 200 leaf points at a depth of 1m or less and over an area of some 25-30m2, in association with unretouched blades and endscrapers, including some

Poland The Middle and Early Upper Palaeolithic of Poland is restricted to the south of the country, and especially the Krakow region. At Nietoperzowa Cave, near

59

carinates. The leaf points are characteristically triangular and often asymmetrical; both bifacial and unifacial (Fig. A2.53) examples occur. Absolon also reported finding Mousterian artefacts. Moravany-Dlhá can therefore be regarded as no more than a possible Middle Palaeolithic leaf point site.

it, and industries like it, are excluded from the Middle Palaeolithic. It is much more difficult to justify the exclusion from the Middle Palaeolithic of the industry from Jankovich Cave, in the Öregkő Hills of northwestern Hungary, by these criteria, although both Vértes (1955a) and Allsworth-Jones (1986) do so. Hillebrand (1917, 1935) found a small assemblage with leaf points in the cave, as well as split-based bone points and ivory tools. Vertés subsequently argued that 19 out of 21 bone and ivory tools had originated from the upper part of the sequence and were thus typical only of the later phase of what he called the ‘Trans-Danubian Szeletian’; Gábori-Csánk (1974) went further and claimed that all came from levels higher than the lithics and were therefore not associated with them in any way. She consequently believes the stone tools should be placed in a late Middle Palaeolithic ‘Jankovichian’ industry. The assemblage is small, with just 121 tools according to Vertés and 102 according to AllsworthJones. This includes 32 bifacial leaf points (Fig. A2.54) and two doubtful unifacial examples. There are only four endscrapers and burins and most of the retouched tools, including some of the leaf points, are made on Levallois blanks. Prismatic blades are absent, and there are no cores and virtually no debitage; there is nothing Upper Palaeolithic about the industry. The site appears, then, to have been only ephemerally occupied, although there are two hearths. Dating is difficult. The fauna, according to Gábori-Csánk, is undiagnostic and could relate to any phase of the Upper Pleistocene other than OIS 4. However, the stratigraphic proximity of the ‘Jankovichian’ to clearly Aurignacian bone artefacts suggests that it is OIS 3 in date.

Hungary The leaf point sites of the Carpathians of northern Hungary, and especially those of the Bükk mountains in the northeast of the country, have played pivotal roles in the history and development of our understanding of Palaeolithic leaf points. Indeed, the notion that the leaf points of central and eastern Europe represent a phenomenon older than and quite distinct from the Solutrean found expression in the culture-industrial taxon ‘Szeletian’ (Červinka 1927; Prošek 1953), for which Szeleta Cave in the Bükk is the type site. The reality and extent of the Szeletian are problems that continue to be debated (Vértes 1962-3; Siman 1990; Adams 1998) but they are not central to the questions under consideration here. It is sufficient to note that, contrary to AllsworthJones (1986, 1990), there are very few leaf point sites in Hungary that can be described as Early Upper Palaeolithic other than Aurignacian, and which have produced more than one leaf point. In fact most Hungarian leaf point sites, like those in Slovakia, cannot be attributed with confidence to one or other side of the Middle-Upper Palaeolithic boundary (see Table A3.3). Of those that can be described as Upper Palaeolithic (Table A3.4), Szeleta Cave itself is by far the most important (Vértes 1959, 1965; Allsworth-Jones 1986). There are two stratigraphically distinct leaf point industries represented at the site: an ‘Early Szeletian’ in Layer 4, and a ‘Developed Szeletian’ in overlying Layers 5, 6 and 7. As at Nietoperzowa, the upper industry is unquestionably Upper Palaeolithic. It includes backed blades and Gravette points, and is radiocarbon dated (Layer 7) to 32,620±400 BP. The lower industry, in addition to 113 leaf points, many with cryoturbation damage (Fig. A2.55), includes scrapers and denticulates (although these are probably the products of cryoturbation and thus of no cultural-diagnostic value), endscrapers, burins and, more significantly, truncated blades, retouched pointed blades, blades with continuous retouch, lames à crête and two bone points. Of the nine cores, one is Levallois, and might derive from a small Mousterian industry recovered from Layer 3 but curated together with the Layer 4 material; seven are single platform prismatic blade cores. Radiocarbon dates of 43,000±1100 BP for uppermost Layer 3 and >41,700 BP for Layer 4, together with charcoal of deciduous species, place the lower industry in an OIS 3 warm episode close to Hengelo.

Southeastern Europe Romania The Middle Palaeolithic of Romania remains poorly characterised both in terms of its techno-typological variability and of its chronology. Romanian archaeologists recognise Typical, MTA, Denticulate and Charentian variants of the Mousterian (Mertens 1996), although this scheme owes as much to the Francophile nature of Romanian culture as it does to the properties of the lithic industries. This is indicated by the variable presence of bifaces (sometimes including leaf points) and/or bifacial retouch in every variant. In addition, leaf points are important in putatively Upper Palaeolithic assemblages from a few sites (see Tables A3.3, A3.4). This, together with a number of erroneously late radiocarbon dates for the country’s Middle Palaeolithic (e.g. Păunescu 1988a) and a dubious Upper Pleistocene chronology constructed from short, cave-derived pollen sequences (Cârciumaru 1979, 1988, 1995), has underpinned the claim that the Romanian Middle Palaeolithic persisted until some time after 30 kya (Păunescu 1988b; Chirica 1990, 1995). Such a chronology is no longer tenable given the radiometric

This, then, is the archetype of the ‘transitional’ Early Upper Palaeolithic industry in central and eastern Europe. It is discussed here only to indicate the criteria whereby

60

dates produced by Honea (1984, 1986, 1990; Mertens 1996).

dates for these horizons and the age attribution cannot be properly scrutinised in the absence of clear publications.

Given these difficulties, the only Romanian Middle Palaeolithic leaf point site worthy of consideration here is the open air location of Ripiceni-Izvor, on the right bank of the River Prut in northeastern Romania (Păunescu 1965, 1989, 1993). Six stratified Middle Palaeolithic horizons have been defined. The lower layers (I-III) feature no bifaces or leaf points, but the overlying ‘MTA’ levels (IV and V) contain both. Levallois reduction is important in all five horizons. The final Mousterian Layer VI has yielded a small, typologically impoverished industry lacking leaf points, and is overlain by an archaeologically sterile layer above which an Aurignacian without leaf points occurs. The leaf points of Layers IV and V are therefore certainly Middle Palaeolithic. The site at these times seems to have been repeatedly used as an occupation centre, since there is a rich faunal assemblage, hearths and even, Păunescu claims, evidence for mammoth-bone huts. According to Păunescu (1993, 92-3 and 118) there are 57 leaf points out of 1421 tools (excluding handaxes) in Layer IV and 4 out of 362 in Layer V; the comparable figures for other bifaces are 200 in Layer IV and 23 in Layer V. Unequivocal leaf points are certainly present but in both layers they grade into a highly variable biface category in which flat handaxes are important, as well as into bifacial scrapers (Figs. A2.62, 63, 64). Ripiceni-Izvor IV and V, then, are further examples of Middle Palaeolithic industries with bifaces that grade into leaf point form. Honea (1984, 1986) has radiocarbon dated Layers IV and V to between 44,800+1300-1100 BP and 40,200+11001000 BP, so an OIS 3 date is strongly favoured. Both layers seem associated with a cold, dry climate with little tree cover.

The overlying levels IVa, IV, III, IIa, II, Ia and I are referred to as ‘Mousterian’; IIa, II and Ia lie between fossil soils placed in the last interglacial and the Brørup interstadial. The industries in levels IVa to IIa are described as Levallois Mousterian and lacking leaf points, but that from II is called Micoquian by Kulakovskaya (1990). Gladilin et al refer to leaf points in Level II but Kulakovskaya, whose strict definition of the type has already been mentioned, denies this. The key to Korolevo, therefore, is the dating. That the supposedly pre-Brørup industry in Ia is said to be Upper Palaeolithic with a prismatic blade technique and numerous endscrapers, while the overlying Level I industry is Mousterian, suggests either that the stratigraphy is rather more complex than the publications allow, or the chronology is much shorter, or both. Either way, where leaf points occur at Korolevo they do so in very small numbers and in the context of other biface forms with which they form a continuum. Bulgaria Perhaps the most spectacular collection of Middle Palaeolithic leaf points in Europe is that from the open air site of Musselievo, in northern Bulgaria some 10 km south of the Danube (Dzambazov 1968; Chmielewski 1977). Haesarts and Sirakova (1979, Table 3) identify 257 whole and fragmentary leaf points and 28 unfinished examples out of 421 retouched tools. Allsworth-Jones (1986, 65) on the other hand, identifies 273 leaf points (of which five are unifacial) and 12 biface knives and handaxes out of 411 tools. Neither publication, indeed no publication, analyses all the material recovered from the site; Chmielewski also identified 30 leaf points and three small handaxes out of 138 tools in the material he excavated in 1970, whereas Sirakova (1990) makes unamplified reference to over 500 complete examples. The leaf points are overwhelmingly bifacial (Figs. A2.65, 66, 67); there are only five unifacial examples. There is some variability in silhouette, butt form and initial blank, but the leaf points are unified by a combination of symmetry, pointedness, thinness and flat surface retouch. This is all the more noteworthy in view of the paucity of other bifaces and of bifacial retouch in other tool categories. All authorities are now agreed that the industry is unequivocally Middle Palaeolithic in character with a predominance of Levallois blanks and cores, although there is dispute as to whether it should be classified as ‘Mousterian with leaf points’ (Ivanova 1979), ‘Mousterian-Levallois culture with leaf points’ (Kozlowski 1975) or Micoquian (Allsworth-Jones 1986). All authorities also agree that the material should be interpreted as a leaf point workshop, the flint on which the pieces were made being derived from nodules present at the site (Sirakova 1990). The leaf points are unequivocal.

Ukraine Early Upper Palaeolithic leaf point assemblages do not occur in the Ukraine (other than Crimea) but there are several ‘Eastern Micoquian’ sites at which Middle Palaeolithic horizons have produced leaf points in a biface-rich industrial context (Table A3.1). The deeply stratified cave site of Korolevo, in the Transcarpathian far west of Ukraine close to the Hungarian border (Gladilin 1989; Adamenko and Gladilin 1990; Kulakovskaya 1990; Gladilin et al 1995), is the most significant of these sites, although its value is limited by inadequate publication. Gladilin et al claim, in addition to a few elongate handaxes, a total of 24 leaf points in ‘Late Acheulean’ levels Vb, Va and V, which they place in the ‘Riss 1/2’ (Vb) and ‘Riss 2/3’ (Va and V) warm events; these are uncertain chronological terms that might fall within OIS 7, between ca 240 and 190 kya. However, they employ a broad definition of a leaf point which includes backed and partially worked pieces. If these are excluded one is left with just two leaf points in Level Vb, eight in Va and one in V (Fig A2.57). Unfortunately there are no absolute

61

The dating of the leaf points is difficult because the finds occur in secondary position, having been carried by solifluction to their present position from a nearby locus of original discard some time around the last glacial maximum (although the presence of both small debitage flakes and larger tools and cores indicates that little of the assemblage has been lost in the process). The bulk of the leaf points were found in layers FA1 and FA2 in the main ‘Champ de Fouilles’ locality, beneath layer FB whose equivalent fossil soils FB1 and FB2 have been identified in a nearby profile and correlated with the OIS 3 Kalabaki, Krinides and Heraklitsa climatic oscillations at Tenaghi Philippon. Fossil soils KC1, KC2 and KC3 in the Kariera locality are also correlated with these events. Haesarts and Sirakova (1979) concluded that the leaf points therefore date to the end of the OIS 4 ‘Lower Pleniglacial’, although Sirakova (1990) subsequently amended this view, placing the industry prior to Kalabaki and Krinides but not Heraklitsa and dating it to between 45 and 50 kya. An OIS 3 date, then, seems highly probable.

points out, the incidence of bifaces at the site is no more than 10% of the tool total and no more than 11 of these are leaf points. To these may be added a further 6 recovered in the course of excavations from 1993-5 (Marks and Monegal 1998). In fact the reality of several of these pieces as leaf points is open to doubt. Many of those from both Formozov’s (Fig. A2.71) and Marks’ (Fig. A2.72) excavations are too thick, too asymmetrical, or both, to qualify as leaf points here; rather, they constitute a range of bifaces including flat handaxes of various silhouettes. In addition, a number of scrapers at the site are bifacially worked and bear flat surface retouch to some degree. Starosel’e seems better described as a Middle Palaeolithic site with bifaces, some of which tend toward leaf point form. Others in Crimea are given in Table A3.1. With regard to dating, Level I has been placed at >41.2±3.6 kya by ESR (Rink et al 1998), and radiocarbon dated to 41,200±1800 BP and 42,500±3600 BP (Hedges et al 1996). According to Kolosov, leaf points occur in significant numbers in the Crimean Middle Palaeolithic only in the so-called ‘Ak-Kai’ or ‘Akkaiskaya’ culture. This is constituted by assemblages with relatively frequent bifaces (20-50%) known from rock shelter and open air sites in eastern Crimea (see Table A3.2) (Chabai 1998). The most important of these is the multiple rock shelter site of Zaskalnaya (Kolosov 1990, 1995), where bifacerich industries with bifacial leaf points occur in the second and third archaeological horizons at both Zaskalnaya V and VI. These levels are radiocarbon dated to ‘more than 40,000-50,000 BP’ (Kolosov 1990, 54). Although Kolosov illustrates 19 unequivocal leaf points from Zaskalnaya V and VI (1990, Plates 1, 2 and 3; see Fig. A2.73), he also points out that this is a small number and that they are greatly outnumbered by other bifaces including a range of knives (Fig. A2.74) as well as several hundred broken bifaces with rounded or pointed ends. So, while leaf points certainly do occur at Zaskalnaya and seem to form a discrete typological category, they form only a small part of a wider biface category. The importance of bifaces in Ak-Kai assemblages might be explained by their preferential use and location close to sources of tabular flint which, as has already been observed, is readily transformed into tools by the removal of cortex from both faces of the tablet or plaquette (Kolosov 1990, 53). Zaskalnaya has been characterised as a ‘base camp’ by Chabai and Marks (1998) on account of the high artefact densities, low tooldebitage ratio, and the presence of hearths and food refuse.

Knowledge of the Bulgarian Palaeolithic remains at a rudimentary level, so the paucity of leaf point sites elsewhere in the country should not be taken as an indicator of their absence. In fact several such sites are now known in open air contexts in the Rhodopes Mountains of southern Bulgaria, although they are only minimally published. Of more interest here, then, is the cave site of Samuilitsa 2 some 120 km southwest of Musselievo (Dzambazov 1964, 1967; Chmielewski 1977; Ivanova 1979). Levels A (basal) through F have produced industries described as ‘Clactonian’ and Mousterian by Dzambazov; all lack leaf points. The industries from the overlying levels G through K, on the other hand, have produced a very few leaf points; Allsworth-Jones (1986, 75-6) identified just 12 out of 640 tools in the entire Middle Palaeolithic sequence (Fig. A2.68). No other bifaces are present. Dzambazov suggested cultural subdivisions into Lower and Upper Szeletian, but this is now rejected in favour of an exclusively Middle Palaeolithic attribution; in fact Allsworth-Jones regards all the Middle Palaeolithic industries (except the very small assemblage from Level A) as Levallois Mousterian and differing only in the presence of leaf points higher in the sequence. There is some evidence for Gravettian presence in levels K and M. A radiocarbon date of 42,780±1270 BP (Vogel and Waterbolk 1972) refers to a level 0.5 metres below the surface of the deposits and thus to the top of the Middle Palaeolithic sequence. With the exception of the absence of bifaces other than leaf points, then, Samuilitsa 2 bears close comparison with Ripiceni-Izvor.

Southwestern Russia A few Middle Palaeolithic sites with bifaces and occasional leaf points are known in southwestern Russia and the northern Caucasus (see Tables A3.1 and A3.2), although most are poorly understood and are only published in detail in Russian. At Kostienki, on the River Don, there is also a complex of open air Early Upper

Crimea Claims for an industry with abundant leaf points have been made for Level I at Starosel’e rock shelter in southwestern Crimea (Formozov 1958; Allsworth-Jones 1986; Kulakovskaya 1990). However, as Kolosov (1990)

62

If the spatial boundaries of this particular Middle Palaeolithic lithic-industrial phenomenon are clear, its boundaries in time are somewhat less so. Many of the site industries are undated or undatable, or can only be placed sometime in the Upper Pleistocene. However, cases certainly attributable to OIS 3 occur at Sesselfelsgrotte Level G, Kůlna levels 7a and 6a, Stinka I Upper, Ripiceni Izvor IV and V, Samuilitsa 2, Mitoc la Sărături, Mitoc Valea Izvorului and Mezmaiskaya and Starosel’e, while examples certainly dating to either OIS 3 or OIS 4 include Champlost, Büdöspest, Stinka I Lower, Boutechty, Bouzdoujany I and Kabazi II. Predmostí can be placed in OIS 4. In addition, the Micoquian-like industry from Grotte du Docteur probably dates from either OIS 4 or OIS 3. In addition, several claims for earlier dates for these industries are open to challenge, as at Salzgitter-Lebenstedt, Königsaue, the Bockstein and Wylotne. The Ukrainian sites deserve mention since there are claims for biface industries with leaf points as old as OIS 7 at Korolevo and Velikij Glubotchok I. Doubts over the validity of the Korolevo chronology were expressed earlier, and the Level II industry might properly be referable to OIS 3. At Velikij Glubotchok I, the OIS 7 date is based upon a magnetic susceptibility date of 212 kya for a ‘glacial soil’ (Bogutskij et al 1999). Given that the reliability of magnetic susceptibility as a dating method is far from established and that glacials are associated with loess deposition rather than soil formation even to the south of Velikij Glubotchok I, the OIS 7 date must be treated with scepticism, although a true age younger than early OIS 4 is unlikely.

Palaeolithic sites often cited in the literature as rich in leaf points. The industry from Kostienki-Tel’manskaya Layer 1, for example (Allsworth-Jones 1986) has produced points, some tanged and all on blades, that are best described as Jerzmanowice points or, in some cases, as pointes à face plane (Fig. A2.76). Several bear multiple distal burinations, and the accompanying industry is strongly laminar and undoubtedly Early Upper Palaeolithic in origin. Chmielewski (1961) regards the industry as an southeastern extension of the Polish Jerzmanowician which, as was discussed earlier, might also be linked with industries from Ranis 2, Spy, Goyet and the ‘Lincombian’ industries of the British Early Upper Palaeolithic.

Discussion Spatiotemporal Distributions This review confirms a pattern, the outlines of which were recognised by McBurney in The Geographical Study and which has since been widely noted; namely, that European Middle Palaeolithic industries with leaf points or leaf point-like bifaces are distributed geographically from northern France and Belgium in the northwest, through central Europe and the Balkans to the Crimea and southwestern Russia in the southeast. Most such assemblages are assigned to the Micoquian as the term is used by archaeologists in central and eastern Europe, although there is dispute as to the cultural affiliation of some. Salzgitter Lebenstedt, for example, is described as ‘Jungacheuléen’ by Bosinski (1967) and the industries from Ripiceni-Izvor levels IV and V are placed in a ‘Mousterian of Acheulean Tradition of Levallois debitage’ by Păunescu (1993). These disputes need not concern us here. The essential point is that, in the Middle Palaeolithic of Europe, the practice of bifacial working leading to the production of a range of biface forms, some of which can be described as foliate or as Blatt-typen or as ‘true’ leaf points as defined here, but extending also to include the ‘Micoquian’ types defined by Bosinski, was a characteristic of stone tool fabrication in these particular regions. This should not be taken to imply that all of these industries can be assigned to a single archaeological ‘culture’, whether or not it be called the Micoquian. At the same time, broadly (and possibly strictly) contemporaneous industries lacking significant levels of bifacial working are also known from the same zone; in Hungary, Érd (Gábori-Csánk 1968) is close to Tata both in space and time (Schwarcz and Skoflek 1982) but lacks bifacial working almost entirely, as does the Levallois Mousterian industry from Level 12 at Bacho Kiro (Kozlowski 1975), Bulgaria, which is only 150 km from and contemporaneous with the biface industry of Samuilitsa 2.

When these factors are taken into account, one can reliably place Middle Palaeolithic industries with bifaces of or approaching leaf point form in periods older than OIS 4 or 3 only at Tata and Kůlna levels 9, 7d and 7c. So, although caution is necessary in view of the number of occurrences of unknown or uncertain date, it appears that the Micoquian industries of central and eastern Europe, and those with similar arrays of biface forms in the same region, date primarily to the OIS 4 and 3 last interpleniglacial period, although a few older occurrences are demonstrated. The spatiotemporal distribution of Middle Palaeolithic industries in which unequivocal leaf points are present as a discrete type, rather than as one pole of a biface morphological continuum– referred to hereafter as leaf point industries or assemblages - is revealed as a subset of the above. Middle Palaeolithic leaf point assemblages are rare, but they are wholly absent from northwestern Europe and, with the exception of Rörshain, from Germany north of Bavaria. In central Europe, northern outliers occur at Bohunice in Moravia and at Jankovich and Máriaremete in Hungary; otherwise, they are restricted to southeastern Europe. Even if one takes into consideration those leaf point assemblages that might be either Middle or Early Upper Palaeolithic, this spatial distribution is not significantly changed. The Lužná localities in Bohemia are somewhat to the north of

63

Bohunice, but the Moravian, Slovakian and Hungarian sites all confirm the northern limit of the leaf point industries. One can confidently conclude, then, that Middle Palaeolithic leaf point industries are found in the south and central German uplands as far north as about 51°N; on the southern flanks of the Northern Carpathians no further north than 49°N; in the Balkans south of 46°30'N; and in Crimea and the northern Caucasus. They are absent from the north European plain, and also from the Hungarian plain. There is therefore a tendency for Middle Palaeolithic leaf point sites to be found in upland and mountain regions, although this should not be pushed too far since their absence in the Hungarian lowlands might reflect the paucity of survey for Palaeolithic sites, and Mamaia in eastern Romania is in the Dobrogean plain close to the Black Sea shore.

this category. The assemblages from the Belgian sites are smaller and judgement is often rendered difficult by the lack of stratigraphic control and poor curation; however, the biface Middle Palaeolithic at Spy, and possibly Goyet, seem to conform to this pattern. The open air sites of northern France and central Germany are more difficult to characterise. Middle Palaeolithic leaf point assemblages, on the other hand, are small to medium-sized and occur in contexts that suggest task-specific, rather than residential, occupation. Rörshain, Kösten, Bohunice and Musselievo, all open air locations, have been interpreted as workshop or leaf point manufacturing sites on account of a preponderance of broken pieces, proximity to the raw material on which the leaf points are made, large numbers of waste flakes and a paucity of hearths and food refuse. There are two stratigraphically distinct hearths at Jankovich, but there is little waste debitage and no cores at all; most of the fauna is cave bear. Given that one third of the 100 or so tools at Jankovich are leaf points, the site is best described as a short term camp or temporary halt; indeed, the evidence is consistent with the assemblage being a leaf point cache, a key place in the landscape at which specific tools were placed for future use. The archaeology of Mauern Layer F and, more problematically, Layer G, is very similar. Again, a small assemblage contains numerous typologically distinctive leaf points, but no evidence for domestic occupation. There are no hearths known at Mauern, and there are numerous cores present; nevertheless the overall archaeological structure of the site is very close to that at Jankovich, and an interpretation of the site as a leaf point cache and ‘gearing up’ site is indicated. The rarity of cores and waste debitage and the absence of hearths at the Obernederhöhle extends this pattern to a second Altmühl Valley site. Other Middle Palaeolithic leaf point assemblages are more difficult to characterise, but all share the assocation between small – sometimes tiny – assemblages and the absence of evidence for domestic or extended occupation.

There is a strong tendency for Middle Palaeolithic leaf point assemblages to date to OIS 3. In fact, where their age is known with any confidence and precision, as at Mauern, Haldenstein, Bohunice and Il’skaya, they date unequivocally to OIS 3. In addition, the leaf points from Máriaremete date to either OIS 3 or OIS 4. Mamaia and Kokkinopilos β relate either to OIS 3 or OIS 5. Of the leaf point assemblages whose age is uncertain, OIS 3 dates seem likely at the Obernederhöhle and Jankovich in view of the stratigraphic proximity of and artefact admixture between Middle Palaeolithic leaf point and Early Upper Palaeolithic levels. A similar guess might reasonably be made for the Crimean leaf point sites in view of an array of radiocarbon, ESR and MSUS dates for the Crimean Middle Palaeolithic, all of which suggest ages of 70 kya or younger (Marks and Chabai 1998). In fact, an OIS 3 date is confirmed or possible at every Middle Palaeolithic leaf point site. Leaf points in Landscape Context The review above reveals more than the Middle Palaeolithic distributions in time and space of industries with foliate bifaces and of leaf point industries; it also highlights a clear distinction between them in terms of the archaeological contexts in which they occur. Middle Palaeolithic industries with a range of bifaces, some of which approach or grade into with leaf point or leaf point-like form, generally occur in stratified, usually cave or rock shelter sites. They are associated with large or medium-sized lithic assemblages in which tools are greatly outnumbered by waste flakes, and often with hearths and, where preservation permits, considerable collections of animal bone food refuse. That is, these industries seem to be associated with repeated, apparently more-or-less extended occupations in which a wide range of tasks were carried out. The terms ‘residential site’, ‘home base’ or base camp’ seem appropriate. Königsaue, Schambach, Sesselfelsgrotte Level G, the Bockstein, Wylotne, Ciemna, Piekary, Kůlna Cave, Subalyuk, Tata, Büdöspest, Korolevo, Bouzdoujany I, Stinka I, Ripiceni Izvor, Boineşti, Zaskalnaya and Mezmaiskaya all fall into

Given the contemporaneity between Middle Palaeolithic leaf point and biface industries in central and southeastern Europe, their proximity in space and their shared use of bifacial working, it is quite inappropriate to place the leaf point industries in separate cultures, such as ‘the Altmühlian’, ‘the Jankovichian’ or ‘the Levallois Mousterian with leaf points’ on the basis of the presence of leaf points. It is proposed instead to consider the Middle Palaeolithic leaf point industries within a particular region together with the biface industries in the same region as aspects of the same landscape settlement system, in which leaf points were associated specifically with the performance of particular activities or tasks at particular places in the landscape beyond the centres of social life. This idea will be explored further in Chapter 6, where the best-known regional cluster of OIS 3 Middle Palaeolithic Micoquian and leaf point sites, that from the Altmühl Valley, will be examined in more detail.

64

If the Middle Palaeolithic leaf point phenomenon can be shown to relate to a particular period within the Upper Pleistocene, i.e. OIS 3 between ca 60 and 40 kya, and to be geographically restricted to central and southeastern Europe, then the contrast with the Lower Palaeolithic prior to 200 kya, in which spatiotemporal specificity in stone tool form and technique on these scales is unknown (Chapter 3), is stark. Given that socially transmitted structures of lithic-technical knowledge in the European Lower Palaeolithic were insensitive to transformations in affordance environments operative even on the wavelength of glacial-interglacial cycles, the question

arises as to whether the spatiotemporal specificity of the Middle Palaeolithic leaf point industries is connected with a greater sensitivity of socially transmitted knowledge to environment transformation. That the Middle Palaeolithic after 200 kya seems to indicate that a trend towards just such an increase in sensitivity was in operation has already been highlighted in Chapter 3; Chapter 5 will now consider the environment history of Upper Pleistocene Europe with the geographical and temporal boundaries of the Middle Palaeolithic leaf point phenomenon in mind.

65

66

CHAPTER 5

UPPER PLEISTOCENE ENVIRONMENT CHANGE IN THE REGIONS OF EUROPE Introduction

both deep sea and terrestrial ice cores, for the following reasons:

An ecological geography of late Middle Palaeolithic leaf points in Europe must be based upon a rigorous understanding of European Upper Pleistocene climate and environmental history. More specifically, the hierarchical approach developed in Chapters 2 and 3 demands an emphasis on the various spatial and temporal scales on which climatic and environmental-ecological processes operated in that period, their mutual sensitivities, and the consequent impact on landscapes in time and space. This entails the elucidation and subsequent comparison of regional variations in landscape character, and their persistence in time, through the course of the Upper Pleistocene. It must be admitted that neither the proxy records available, nor the analytical methods available to apply to them, permit the detailed reconstruction of mosaic landscape grain on spatial scales commensurate with human grain responsiveness. Neither is it possible to resolve temporal developments in landscape character much below the millennial scale. However, the remarkable expansion in Quaternary science’s knowledge of the histories of climate and environment change in Europe during the Upper Pleistocene over the last twenty five years or so does now permit an investigation of this problem to proceed on the basis of a degree of resolution that would have been unimaginable for McBurney in 1950. Given that the late Middle Palaeolithic leaf point phenomenon appears to have been sub-continental in extent and might have endured for as much as twenty millennia or longer, it is possible to conclude that European Pleistocene climate and environment history is now understood in sufficient spatiotemporal detail to render a hierarchical ecological geography of the phenomenon practicable. It is on this basis that the implications of leaf points for relations between socially transmitted knowledge, knowledgeable action and scale domains of variability in the world in the late Middle Palaeolithic can be understood.

i. The Pollen Record. Despite the well known difficulties attendant on the interpretation of pollen sequences (Lowe and Walker 1997, 163-75), the pollen record remains the biotic-environmental proxy that offers the best combination of high resolution, sensitivity, sheer numbers of individual sequences, geographical spread and temporal depth. The recovery and analysis over the last twenty five years or so of long, continuous pollen sequences from across the European continent and covering most or all of the Upper Pleistocene, together with calibration against the marine oxygen isotope curve, has revolutionised understanding of events in Upper Pleistocene vegetation history, their chronology and their correlation with climatic events. Palynology thus offers the potential for constructing regional and continental scale vegetation histories within chronological frameworks of much-improved robusticity. There are, of course, problems in relying too much on the pollen record. The difficulties involved in interpreting Quaternary ecological communities and their structures from contemporary analogues has been discussed already. In addition, palynologists’ preference for pollen sources with a large pollen catchment area (and, consequently, regional scale vegetation signal) militates against the possibility of reconstructing landscapes on spatial scales small enough to address the question of mosaic landscape grain. Nevertheless the pollen record, as the best proxy record of vegetation history, represents the most powerful first-order record of transformation in humanly experienced and known landscapes through the Pleistocene. The locations of the various pollen core sites, spectrum sequences from which are cited in this Chapter, are shown at Figure 5.1. Pollen diagrams subjected to the quantitative analyses described later are given in Appendix 1. ii. The Marine Stable Oxygen Isotope Curve. The deep sea oxygen isotope curve represents a record of global ice volume, and thus ultimately, but not directly, of global temperature (Emiliani 1955; Dansgaard and Tauber 1969; Shackleton 1967, 1987; Mix and Ruddiman 1985). The remarkable consistency in curves derived from individual cores from around the world (Patience and Kroon 1991) and the closeness of match between the curves and Milankovitch variations in solar insolation (Hays et al 1976; Imbrie and Imbrie 1979; Imbrie et al 1984) indicate the integrity of the global signal represented in this record

Proxy Indicators of Upper Pleistocene Climate and Environment The range of proxy indices of past climate and environment, both terrestrial and marine, is large. It is impossible here to mount an exhaustive analysis of them all. Instead this investigation will rely primarily on the pollen record and the oxygen isotope curves derived from

67

Figure 5.1: Locations of palynological sites discussed in the text: 1, Chelford; 2, Brumunddal; 3, Sejerø; 4, Brørup; 5, Rederstall; 6, Odderade; 7, Oerel; 8, Amersfoort; 9, Denekamp; 10, Hengelo; 11, Watten; 12, Scladina; 13, Tourbière de Biscaye; 14, Combe Grenal; 15, Le Moustier; 16, Lac du Bouchet; 17, Les Echets; 18, La Grande Pile; 19, Sulzberg; 20 Dürnten; 21, Mauern; 22, Samerberg; 23, Kittlitz; 24, Warsaw-Wola; 25, Gołkow; 26, Zgierz-Rudunki; 27, Imbramowice; 28, Kąty; 29, Ljubljana Moor; 30, Padul; 31, Carihuela; 32, Cova Beneito; 33, Abric Romani; 34, Valle di Castiglione; 35, Monticchio; 36, Tyrrenhian Sea KET 8003; 37, Ioannina; 38, Xinias; 39, Tenaghi Philippon.

(see Figure 5.2), although it must be emphasised that it reveals little of regional or even continental scale climate variability. Indeed, the utility of the deep sea oxygen isotope record both as a proxy of past global climate and as the benchmark Pleistocene chronology is now so widely accepted and familiar that no further amplification is necessary here. A summary account of its principles and methodology can be found in Lowe and Walker (1997, 149-53).

This explains their sensitivity to ‘Dansgsaard-Oeschger’ (D-O) temperature oscillations on wavelengths of 1-3 millennia (Figure 3.1). These are generally virtually invisible in long oxygen isotope curves derived from deep sea cores since very low rates of sea floor sediment accumulation in the deep ocean limit their resolution (Lowe and Walker 1997, 153). In the Antarctic Vostok core the δ18O curve is based upon the estimation of the 18 O content of air bubbles trapped in the ice, and as such measures the 18O:16O ratio in the atmosphere (δ18Oatm). This index is strongly coupled with marine isotopic ratios, so D-O scale climate processes are rather less well expressed in the Vostok than in the Greenland curves. Compaction under the weight of ice limits the resolution of the ice core record beyond 100 kya or so, although this does not compromise its applicability to the problem in hand here. The importance of high amplitude temperature oscillations on sub-Milankovitch timescales for the maintenance of the Eurasian mammoth steppe has been discussed at length in Chapter 3; their particular relevance in OIS 3 is discussed in more detail below. The central point here is that the value of the ice core oxygen isotope curves for investigations of Upper Pleistocene

iii. The Ice Core Stable Oxygen Isotope Curve. The oxygen isotope curves derived from the coring of terrestrial ice sheets, particularly the Greenland GRIP (Dansgaard et al 1993; Bond et al 1993) and GISP2 (Grootes 1993) cores but also the Antarctic Vostok core (Lorius et al 1985; Petit et al 1999) and the Tibetan Guliya core (Thompson et al 1997) provide high resolution records of isotopic ratios with very considerable time depth. The Greenland ice core δ18O curves are constructed on the basis of the 18O content of the ice itself (δ18Oice); this proxy records condensation temperature above the ice sheet, a variable with a very much higher process rate than the marine 18O:16O ratio. 68

ecological and environmental history lies not so much in chronology as in the insights they offer into the multiscalar character of climate change in that epoch. Taken together, then, these proxies permit a critical evaluation of the differing impacts on plant communities of climate oscillations on various wavelengths and in the various regions of Europe, and thus can provide a context for an examination of the scalar character of knowledgeable human relations with changing environments in Upper Pleistocene Europe. Other proxy indicators of climate and environment will be cited when relevant, but these three indices will form the basis of this analysis.

OIS 5: Global and European Synchroneity A review of the palaeoenvironmental and palaeoclimatic records for the 55,000 years or so of OIS 5 will clarify the extent to which it can be thought of as a period of high amplitude, global scale oscillations in high level climatic constraints that induced synchronous oscillations in many coupled systems, including ecosystems. The OIS 5 Oxygen Isotope Record The essential features of OIS 5 in the ‘stacked’ deep sea oxygen isotope curve shown at Figure 5.2. are: i. A peak of isotopic lightness following Termination II at 130 kya. This indicates a warm interglacial episode corresponding to OIS 5e. Figure 5.2. The stacked marine oxygen isotope curve for the last 300,000 years. The Oxygen Isotope Stages to which the peaks and troughs correspond are indicated. After Martinson et al 1987. Note that OIS sub-stages are identified using the ‘Decimal’ system of Pisias et al (1984) in which OIS 5a, 5c and 5e are represented as 5.1, 5.3 and 5.5 respectively. The system is not widely favoured.

ii. From around 118 kya, a general trend towards increased isotopic heaviness persisting for some 45,000 years up to an δ18O enrichment peak (i.e. temperature minimum) commencing some time shortly before 70 kya and marking the end of OIS 5 and the onset of the OIS 4 pleniglacial maximum. iii. Two major periods in which this trend towards isotopic enrichment is reversed are represented by two depletion troughs which, although less developed than that of OIS 5e, nevertheless represent extended deglaciation, and thus warming, events. The older, lasting from around 104 to 95 kya, is OIS 5c. Within this there is a short but marked isotopic enrichment peak where δ18O values rise by some 0.15‰ relative to the preceding and succeeding phases of the event, implying a brief but significant climatic cooling (Pisias et al 1984); the younger is OIS 5a, attributed to around 85-74 kya.

the isotopic shift from OIS 5e to 5d is around 0.8‰, i.e. approximately half that observed across the boundary at Termination 1 between the last glacial maximum OIS 2 and the Holocene OIS 1. Each of the five sub-stages of OIS 5 lasts for around 10,000 years, which implies the operation of a warming-cooling climatic cycle with a wavelength of around 20,000 years. This has been attributed to the operation of the precessional ‘wobble’ in the Earth’s rotation on its axis, which exhibits a wavelength of some 21,000 years, during an interval of high orbital eccentricity in which the Earth followed a

iv. These three OIS 5 depletion troughs/temperature maxima are separated by two enrichment peaks/temperature minima, sub-stages 5d (ca. 118-104 kya) and 5b (ca. 95-85 kya). In the stacked δ18O curve

69

strongly elliptical path around the sun (Shackleton 1969; Broecker et al 1968).

This coupling produced a history of high amplitude climatic cyclicity characterised by wavelengths of around 20,000 years and which was expressed in a wide range of climate-related parameters as a globally synchronous sequence of three warm events - the first of which was a full interglacial - separated by two cold events. OIS 5 thus appears to have been a period in which, in hierarchical terms, those climate systems and system components whose characteristic process rate wavelengths were of the order of 20,000 years or less were profoundly and synchronously affected by high amplitude oscillations of that wavelength order in high level constraining parameters. That these constraint excursions were probably orbitally-forced in character is suggested precisely by the global synchroneity observed and by their close correlation with Milankovitch variations in solar insolation. Given that ecosystem dynamics even on spatial scales as large as the continental generally operate with process rates significantly or much higher than this, one would expect that European vegetation history in OIS 5 would also show radical and synchronous responses to these overarching climate cyclicities.

The same pattern is unequivocally visible in the GRIP ice core δ18Oice curve (Figure 3.1), where three clusters of δ18Oice enrichment peaks - here signifying temperature peaks since in terrestrial ice cores the relation between 18 O:16O ratios and temperature is the reciprocal of that in the deep sea floor – occur within OIS 5, the first being appreciably higher than the second and third and corresponding to OIS 5e. The transition from the putative OIS 5e to 5d in the GRIP curve is marked by an isotopic shift of around 6‰, from which Dansgaard et al (1993) infer a temperature decline in the order of 9°C. The GRIP curve, like the stacked deep sea curve, also suggests a short cold episode within OIS 5c. The GISP2 δ18Oice curve (Grootes et al 1993) correlates almost perfectly with that from GRIP up to depths of around 2700 metres, with the first 22 out of 24 Upper Pleistocene ‘interstadials’ visible in the GRIP curve also unequivocally present in GISP2. At depths greater than 2700 metres, corresponding to around the OIS 5c/5b boundary, layer thicknesses between the two cores begin to diverge, implying that depth-stratigraphic correlations between the cores may be problematical beyond 2700 metres and that the undivided enrichment peak attributed to OIS 5c in the GISP2 curve might be questionable. Finally, the same pattern of three isotopic enrichment peaks occurs in both the Antarctic Vostok δ18Oatm curve (Jouzel et al 1987; Petit et al 1999) and the Tibetan Guliya δ18Oatm curve (Thompson et al 1997).

The OIS 5 Pollen Record in Europe Although there is an extensive palynological database in existence for Europe in OIS 5, the records from which it is derived are distributed essentially in the north, west and centre of the continent (see Figure 5.1). The Upper Pleistocene is scarcely represented at all in pollen sequences from that part of Europe south and east of southern Poland and north of Greece. Nevertheless the course of the continent’s vegetation history in OIS 5 is well established, and conforms closely to the template derived from the palaeoclimatic records discussed above.

OIS 5 in Other Palaeoclimate Records The consistent OIS 5 pattern of three warm events separated by two cool episodes extends also to other physico-chemical proxy records derived from deep sea and terrestrial ice cores. The calcium carbonate percentage curve in marine core RC11-120, for example, shows three warm episodes within OIS 5 (Martinson et al 1987), as does the dust content in core V21-146 (Hovan et al 1989). Percentage frequency curves for the obligate polar planktonic foram Neogloboquadrina pachyderma sinistralis in the North Atlantic cores DSPD Site 609 and V23-81 both show depletion troughs/temperature maxima corresponding to substages 5a and 5c, and enrichment peaks/temperature minima corresponding to OIS 5b (Bond et al 1993). Estimations of OIS 5 sea surface temperatures based on frequencies of the same foram in the chronologically deeper North Atlantic core V23-82 (Sancetta et al 1973) show the familiar fivefold pattern. The Vostok ice core exhibits the same pattern in its deuterium (δD), carbon dioxide and methane curves (Chappellaz et al 1990); the same is true of the Guliya chloride ion curve (Thompson et al 1997).

The OIS 5e Interglacial Where the OIS 5e interglacial is represented in European pollen sequences it is characterised by very similar sequences of pollen spectra throughout the continent. North and west of the Alps, an initial episode of Betula (birch) and Pinus (pine) expansion is followed by a long thermophilous forest period of very high AP% values (typically 90% or more) during which first Ulmus (elm), then Quercus (oak), followed by Corylus (hazel), Taxus (yew) and finally Carpinus (hornbeam) reach their acmae. The interglacial ends with a period in which Abies (fir) and Picea (spruce) forests dominate and Pinus and Betula again expand (Zagwijn 1989; van Andel and Tzedakis 1996). Although there are differences in detail between sites, this essential pattern is visible in sequences from, for example, La Grande Pile borings X (Woillard 1978) and XX (de Beaulieu and Reille 1992a), Les Echets G (de Beaulieu and Reille 1984) and Lac du Bouchet D (Reille and de Beaulieu 1990) in France, Samerberg in Bavaria (Grüger 1979), Oerel (Behre and Lade 1986) and Rederstall (Menke and Tynni 1984) in northern Germany, Machnacz (Kupryjanowicz 1991) and

The world’s climate in OIS 5, then, was characterised by a powerful coupling of multiple climatic-system components in the oceans, atmosphere and ice sheets.

70

Imbramowice (Mamakowa 1989) in southern Poland and sites in the Low Countries (Zagwijn 1996). Even as far east as Nizhnyaya Boyarschina in the western Russian Plain, close to the border with Belarus, the Mikulino interglacial pollen succession conforms to the northwestern Eemian template except in details such as a Quercus maximum marginally earlier than that of Ulmus and the presence of Tilia (lime) instead of Taxus (Grichuk 1984).

(Wijmstra 1969), and Valle di Castiglione (Follieri et al 1988, 1989) and Vico (Leroy et al 1990) in Italy. The pan-European scale of forest expansion in these interstadial events is therefore clearly evident in the palynological record. Perhaps the best indication of pan-continental synchroneity in European vegetation history in the early part of the last glacial is a brief but significant decline in thermophilous and mesophilous trees during OIS 5c. Woillard (1978, 1979) identified this as a cooling phase, which she named Montaigu, within St Germain I at La Grande Pile boring X. This episode is characterised by a rise in the pollen representation of the cool-tolerant taxa Pinus and Betula and a synchronous dramatic decline in that of other, more warmth-demanding genera including Picea, resulting in a small decline in total AP%. St Germain I at La Grande Pile is thus a tripartite interstadial divisible into St Germain Ia, Ib (Montaigu) and Ic. An identical cooling event occurs also at Lac du Bouchet D and Les Echets G (Pons et al 1992). Reille et al (1992) have characterised this episode as the Montaigu event. The Amersfoort and Brørup interstadials named after their type sites in The Netherlands and Denmark, though originally interpreted as two quite separate interstadials (Zagwijn 1961; Andersen 1961), are now regarded as the older and younger climatic optima respectively of this tripartite OIS 5c interstadial known as Amersfoort-Brørup in the Netherlands and Brørup in Denmark and Germany. The same structure to the first interstadial above the last interglacial has also been identified at, for example, Rederstall, Oerel, Watten and Samerberg, as well as at Sulzberg in the Alpine foreland of northern Switzerland (Welten 1982) and Kittlitz in eastern Germany (Erd 1973). Similarly, brief but significant episodes of decline in warmth-demanding trees during OIS 5c are apparent in the cores from Mediterranean Europe such as Padul 2, Ioannina 249, Valle di Castiglione and Tenaghi Philippon (where Wijmstra’s Drama and Doxaton interstadials are now held to equate to St Germain Ia and Ic respectively), and even in the pollen curve from Clear Lake, California (Adam and West 1983). That this clearly pan-European, even global event in vegetation history can be correlated with the similarly tripartite OIS 5c in the stacked deep sea oxygen isotope curve (Figure 5.2) indicates clearly the continuing synchroneity of climatic and vegetation history in Europe during this period.

The Eemian palynology of southern Europe shows a history of forest development very similar to that of the north and west of the continent. For example, the Eemian succession in the Ioannina 249 core from northwestern Greece – a Quercus/Ulmus phase, followed by Carpinus/Ostroya and finally Abies phases – corresponds well with the northern and western record. There are some unsurprising variations. The northern Taxus maximum referred to above is absent in Mediterranean sequences, where an apparently contemporaneous expansion of Olea (wild olive) is visible instead. Picea is rare or absent in southern European sequences, and there is a tendency for Eemian Quercus populations in the Mediterranean to have been evergreen rather than deciduous in character (Tzedakis 1994; van Andel and Tzedakis 1996). The OIS 5c and 5a Interstadials Like the OIS 5e interglacial, the two later OIS 5 warm phases described above in the global palaeoclimate record are clearly expressed in the pollen record as phases in which tree cover expanded throughout Europe. Similarly the OIS 5d and 5b stadials are unequivocally episodes of expansion in open country and of decline in arboreal species. In this respect OIS 5 subsequent to the OIS 5e/Riss-Würm/Eemian/Mikulino full interglacial substage (i.e. early Würm/Weichsel/Valdai) clearly conforms to the pattern of OIS 5 climatic-environmental synchroneity apparent in the marine and ice core oxygen isotope records. It is consistent also with evidence for two early glacial regressions of the Scandinavian ice sheet and increases in sea levels (van Andel and Tzedakis 1996). At La Grande Pile borings X and XX, for example, AP% in OIS 5c (St Germain I) and 5a (St Germain II) attains peak values of 93% and 95% respectively (Woillard 1978; de Beaulieu and Reille 1992a). These are comparable with Eemian values in the same core. AP% in OIS 5c and 5a also reaches high levels consistent with significant regional expansion of arboreal species at values close to those of OIS 5e at pollen boring sites as far apart as Watten in north eastern France (Emontspohl 1995), Odderade (Averdieck 1967), Rederstall (Menke and Tynni 1984) and Oerel OE61 (Behre and Lade 1986: Behre 1989) in northern Germany, Amersfoort in The Netherlands (Zagwijn 1961), Brørup in Denmark (Andersen 1961), Samerberg in Bavaria (Grüger 1979, 1989), Machnacz in Poland (Kupryjanowicz 1991), Padul 1 and 2 in southern Spain (Florschütz et al 1971; Pons and Reille 1988), Ioannina 249 (Tzedakis 1994) and Tenaghi Philippon in Greece

North-South Vegetation Gradients in OIS 5c and 5a The palynological record suggests strongly that the species composition of woodland in OIS 5c and 5a was rather more powerfully linked to latitude than in OIS 5e. Northwestern Scandinavia in these interstadials was a tundra landscape (van Andel and Tzedakis 1996). Further south, cores from Brumunddal, Norway (Helle et al 1981), the Danish sites Brørup and Sejerø (Andersen 1961), Odderade, Rederstall and Oerel in northern Germany and Amersfoort in The Netherlands – all north 71

In Mediterranean Europe, on the other hand, forest development in OIS 5c and 5a was truly thermophilous. Quantitative comparisons (derived from measurement of published pollen diagrams) between the mean percentage contribution made to total arboreal pollen by taxa other than Pinus, Betula and the mesophilous conifers Picea, Abies and Larix, i.e. by thermophilous, mainly broadleaved trees, indicates that north of 46°N this index never attains a value greater than that of 23% at La Grande Pile in OIS 5c3 (excluding the cave deposit sequence from Scladina), whereas it attains 70.58% at Ioannina 249 (39° 39'N) (Tzedakis 1994) and 62.73% at Padul 2 (37° 1'N) in southern Spain (Pons and Reille 1988). For OIS 5a the equivalent values are 48.52% at La Grande Pile XX, 61.55% at Ioannina 249 and 67.80% at Padul 2. The presence at Tenaghi Philippon and Ioannina 249 of, for example, Cistus, Olea and Pistacia (pistachio) alongside broad-leaved elements like Quercus and Ulmus, often evergreen, suggests the formation of fully thermophilous forests with typically Mediterranean components.

of 52º latitude– record OIS 5c and 5a woodland of a strongly boreal character with mesophilous and thermophilous species very weakly represented. Pollen spectra attributed to OIS 5c and 5a at these sites are dominated by Pinus and Betula; no other taxa show continuous curves. Boreal woodland also occurred in Britain in the Chelford interstadial which probably corresponds with the Brørup interstadial of northern Europe (West 1977) and thus with at least a part of St Germain I at La Grande Pile. However, deciduous trees are more strongly represented in these intervals further to the west and south. According to Behre (1989) the northern limit of thermophilous trees in Germany during Brørup was 4-6° latitude south of Oerel, i.e. between 49° 30' and 47° 30' North. In fact warmth-demanding trees seem to have occurred, even if sparingly, a little further to the north. At Watten in the Pas de Calais (50º 50'N) Pinus, Picea and Betula dominate but other species show continuous low-value curves (Emontspohl 1995). The northernmost known occurrence of mesophilous woodland in these Early Würm/Weichsel interstades appears to be at Scladina Cave (50º 30'N), southern Belgium, where pollen spectra strikingly rich in warmth-demanding tree species have been correlated with St Germain I and II (Bastin 1992). However, this sequence should certainly be treated with great caution, at least as an indicator of regional vegetation, owing to the tendency of plants growing in the immediate vicinity of cave entrances to be heavily over-represented in the pollen spectra recovered from cave deposits.

This analysis demonstrates that the distribution of European vegetation in the early part of the last glacial, at least in the two interstadials, displayed a very marked north-south gradient, much steeper than those of either the last interglacial or the Holocene. A climatic explanation for the observed palynological patterns appears necessary given that the durations of both the Brørup and Odderade interstadials were quite sufficient at 5-10,000 years to permit the establishment of warmthdemanding trees north of these latitudes if climatic conditions had been appropriate (Behre 1974, 1989). Frenzel (1980) has calculated from palynological data that mean annual temperatures in OIS 5 interstadials were 11° cooler than today in northern Germany but only 2° cooler at La Grande Pile. Ruddiman and McIntyre (1976) have also reported a southward shift relative to OIS 5e of some 10° latitude in OIS 5d and 5° in OIS 5c in the boundary between sub-polar and transitional waters in the North Atlantic.

At La Grande Pile boring XX at 47° 44'N, altitude 360m, St Germain I and II AP% values sometimes exceed 90% and show successional developments which differ from that of the Eemian in the same boring only in relatively minor ways. However, the interstadial pollen sequences in this boring are characterised by markedly lower pollen frequencies from taxa other than of Pinus and Betula than is the case for the Eemian (de Beaulieu and Reille 1992a). It also noteworthy that in St Germain Ic at La Grande Pile XX Gramineae (grasses) persist at pollen percentage values of at least 5%, often attaining 10-15%, whereas in Eem this parameter is depressed to 3% or below throughout the long thermophilous forest phase. In St Germain II Graminaea values remain as high as ca. 20%. OIS 5c/St Germain I and OIS 5a/St Germain II in the Vosges, then, appear to have seen a somewhat more significant persistence of open country taxa than was the case in the interglacial, from which it can be inferred that OIS 5c and 5a landscapes in the Vosges were rather more open and fragmented, with a finer grained mosaic character, than in OIS 5e. It would seem more accurate to talk in terms of mixed boreal/mesophilous woodland, rather than forest, development. This trend in France towards somewhat less thermophilous woodland in OIS 5c and 5a than in OIS 5e is also apparent, though significantly less well developed, further south at Les Echets G (45° 49'N, altitude 267m) (de Beaulieu and Reille 1984).

East-West Vegetation Gradients in OIS 5c and 5a Although the latitudinal vegetation gradients in OIS 5c and 5a Europe are well-established, much less attention has been paid to the possibility of longitudinal vegetation gradients in these periods. That the woodland of Europe in the interstadials of the early last glacial departed from the Eemian regional distribution pattern along a northsouth transect ought, however, to suggest the possibility that there were also regional variations along an east-west transect. Any attempt to address this problem is hampered by the paucity of the Upper Pleistocene palynological record in central and eastern Europe south of southern Poland and east of the German-Czech border. Nevertheless it is possible to make certain comparisons between pollen sequences that cast light on the matter. i). In Europe north of 52°N woodland development in the OIS5c and 5a interstadials seems to have been 72

everywhere strongly boreal in character. The palynological evidence from Brumunddal (60° 53'N), Brørup (55° 29'N), Sejerø (55° 52'N), Odderade (54° 8'N), Rederstall (54° 14'N), Oerel (53° 29'N), Chelford (53° 16'N) and Amersfoort (52° 9'N) have already been cited as yielding pollen sequences with no significant contribution from warmth-demanding trees in these intervals. Further to the east, sites north of 52°N which have yielded pollen sequences referable to OIS 5c, OIS 5a or both include Machnacz at 53° 8'N (Kupryjanowicz 1991), Horoszki (52° 15'N), Warsaw-Wola (52° 14'N) and Gołkow (52° 3'N), all in Poland (Mamakowa 1989, 135 and 138). At each of these sites the OIS 5c and 5a (EV2 and EV4 respectively in Mamakowa’s Polish pollen chronology) spectra indicate AP% averages in the region of 80% with heavy domination by Pinus and Betula. No other arboreal taxa show a continuous curve or exceed a representation of 3%. However the occurrence of Picea, Abies and Larix at a total percentage of 7.63% in the Oerel OE61 core (calculated from Behre and Lade 1986, Tafel 3) at 9° 3'E, when compared with a total percentage for the same taxa of just 0.50% at Machnacz MII (from Kupryjanowicz 1991, Figure 3) at 23° 9'E, may certainly be suggestive of some longitudinal gradient at these latitudes.

distortions that may arise from inclusion of the OIS 5c2 Montaigu climatic deterioration and the short-lived OIS 5c1 warm interval, in which migration time lags may have mitigated against the development of woodland succession) and OIS 5a. This was achieved by scanning each diagram at a resolution of 700 dpi and producing an enlarged print to improve resolution. AP%, Pinus%, Betula%, Picea%, Abies% and Larix% plots were then measured for each individual spectrum to the nearest 0.1mm and the percentage contribution of each to the spectrum calculated with reference to the measurement of the 100% interval. These taxa were selected for their tolerance of cool conditions; subtraction of Pinus and Betula from AP% yields a gross measure of the representation of mesophilous and thermophilous taxa to the spectrum, and further subtraction of the remaining mesophilous coniferous taxa produces a residual figure corresponding essentially to the contribution made by thermophilous taxa. The sites, with latitude and longitude, together with the results of the analysis (mean total arboreal pollen percentage, mean Pinus+Betula percentage, mean arboreal pollen percentage excluding Pinus and Betula, mean Picea+Abies+Larix percentage and mean arboreal pollen percentage excluding Pinus, Betula, Picea, Abies and Larix) are given in Table 5.1 for OIS 5c3 and Table 5.2 for OIS 5a. Pollen diagrams are presented in Appendix 1. Note that Gołkow and Kąty both lack OIS 5a sequences, and that the values for Scladina have been drawn directly from Bastin 1992 Table 1.

ii). There is little evidence in the palynological record for any significant longitudinal vegetation gradients in southern and Mediterranean Europe in OIS 5c and 5a. At Padul (3° 38'W), Valle di Castiglione (12° 45'E), Ioannina 249 (20° 55'E) and Tenaghi Philippon (24° 20'E) woodland development in these interstadials was, as is mentioned above, marked by the expansion of thermophilous trees, often evergreen, with Mediterranean elements. Comparison between the Greek records Ioannina 249 and Tenaghi Philippon does indeed reveal an east-west vegetation gradient referable to greater precipitation at the former, which is situated close to Greece’s western Adriatic coast, whereas Tenaghi Philippon in the northeast of the country is in the rain shadow of the southern Balkans. The consequence is a greater tendency to the development of lush forests at Ioannina than at Tenaghi Philippon (Tzedakis 1994, 422). Such a gradient, however, is strictly regional, rather than continental, in extent. At Padul, far to the west in southern Spain, the pollen sequence does not indicate greater precipitation than at Ioannina. The differences in the pollen curves from these Mediterranean cores are best explained in terms of local conditions – altitude, topography, local precipitation and soil, for example – rather than by reference to any pan-European gradient or trend.

The low number of sites included in the analysis, dictated primarily by the paucity of sites at these latitudes east of Germany, precludes the possibility of meaningful statistical treatment of the results, which should therefore be seen as suggestive rather than conclusive. In addition, the three easternmost sites, Gołkow, Zgierz-Rudunki and Kittlitz, are all north of 51°N, so care must be taken to avoid confusion between longitudinal and latitudinal variation. Nevertheless the results reveal the influence of both latitudinal and longitudinal gradients on the distribution of tree species at these latitudes in the OIS 5c3 and 5a interstadials. This can be seen in the graphical plots shown in Figure 5.3a-l. The relation between mean AP% in OIS 5c3 and longitude is shown in Figure 5.3a, and with latitude at Figure 5.3g. The mean AP% value of 81.57% at Watten (site 4) could be regarded as anomalously low if one accepts that it is middle Brørup in age (Emontspohl 1995, 236) and thus might include a part of the cool Montaigu event OIS 5c2 in which depressed AP% values could be expected, and which has been excluded from the analysis elsewhere. Equally the low value of 65.10% at Scladina (site 5) might be thought of as aberrant on account of the cave depositional context. If the values at these two sites are ignored then the suggestion of a decline in AP% with increasing northerliness in Figure 5.3g would be weakened, and a weak trend of declining AP% with increasing easterliness might then be thought apparent

iii). It is at mid-latitudes, between 45° and 52°N, that sustained east-west vegetation gradients are most clearly visible in OIS 5c and 5a. This is demonstrated by the quantitative analysis of pollen curves at relevant sites. As before, this analysis involved the identification of all individual pollen spectra referred to OIS 5c3 (chosen in preference to the whole of OIS 5c in order to avoid the

73

SITE*

LAT /°N

LONG /°E

MEAN AP%

MEAN Pinus+ Betula%

MEAN AP% EXCLUDING Pinus, Betula

MEAN Picea+Abies +Larix%

MEAN AP% EXCLUDING Pinus, Betula, Picea, Abies and Larix

1. Gołkow

52.05

20.98

88.13

83.94

4.19

1.87

2.32

2. Zgierz-Rudunki

51.87

19.42

88.93

84.07

4.86

2.80

2.06

3. Kittlitz

51.13

14.41

95.14

89.55

5.59

2.54

3.05

4. Watten

50.83

2.22

81.57

63.68

17.89

9.86

8.03

5. Scladina

50.50

5.03

65.10

22.00

43.10

3.90

39.20

6. Kąty

49.42

20.38

84.69

54.33

30.36

27.98

2.38

7. Samerberg

47.77

12.20

96.19

34.41

61.78

57.91

3.87

8. La Grande Pile XX

47.73

6.50

92.05

50.25

41.80

21.11

20.69

9. Sulzberg

47.50

9.00

93.86

18.60

75.26

65.32

9.94

10. Les Echets G

45.81

4.92

95.25

32.03

63.22

12.25

50.97

Table 5.1. Results by site of quantitative analysis of OIS 5c3 pollen sequences from mid-latitude Europe. Latitude and longitude are expressed in decimal fractions of degrees, rather than in degrees and minutes. *Gołkow; analysis performed on pollen zones EV2a,b in published diagram Mamakowa 1989, Figure 35, (reproduced in Appendix 1 Figure A1.7): Zgierz-Rudunki; zone ‘Amersfoort’, Mamakowa 1989, Figure 35 (Figure A1.8): Kittlitz, zone W IV, Erd 1973, Abb. 8 (Figure A1.14): Watten, zone 3, Emontspohl 1995, Figure 3 (Figure A1.1): Scladina, Bastin 1992 Table 1 (Figures A1.2a,b): Kąty II, zones B, C, D, Mamakowa et al 1975, Figure 5 (Figure A1.9): Samerberg ‘Kernsbohrung’, zones 18,19, Grüger 1979, Beilage 3 (Figure A1.10b): La Grande Pile XX, zone 4c, de Beaulieu and Reille 1992a, Figure 1 (Figure A1.3): Sulzberg, zone FWI 1b, Welten 1982, Dia. 44 (Figure A1.15): Les Echets G zones D6-11, de Beaulieu and Reille 1984, Figure 3 (Figure A1.4b).

SITE†

MEAN AP%

MEAN Pinus+ Betula%

MEAN AP% EXCLUDING Pinus, Betula

MEAN Picea+Abies +Larix%

MEAN AP% EXCLUDING Pinus, Betula, Picea, Abies and Larix

2. Zgierz-Rudunki

70.08

67.25

2.83

1.28

1.55

3. Kittlitz

80.96

78.85

2.11

1.18

0.93

4. Watten

65.37

46.83

18.54

7.65

10.89

5. Scladina

74.50

21.40

53.10

0.60

52.50

7. Samerberg

88.33

44.05

44.28

40.01

4.27

8. La Grande Pile XX

87.35

36.27

51.08

8.70

42.38

9. Sulzberg

86.59

30.62

55.97

31.97

24.00

10. Les Echets G

91.00

16.94

74.06

3.82

70.24

Table 5.2. Results by site of quantitative analysis of OIS 5a pollen sequences from mid-latitude Europe. †

Zgierz-Rudunki; zone ‘Rudunki’, Mamakowa 1989, Figure 35 (Figure A1.8): Kittlitz, zone W VI, Erd 1973, Abb. 8 (Figure A1.14): Watten, zone 5, Emontspohl 1995, Figure 3 (Figure A1.1): Scladina, Bastin 1992 Table 1 (Figures A1.2a,b): Samerberg ‘Kernsbohrung’, zones 22-25, Grüger 1979, Beilage 3 (Figure A1.10b): La Grande Pile XX, zone 6, de Beaulieu and Reille 1992a, Figure 1 (Figure A1.3): Sulzberg, zone FWI 1b, Welten 1982, Dia. 44 (Figure A1.15): Les Echets G zones F2-5, de Beaulieu and Reille 1984, Figure 3 (Figure A1.4b).

74

Figure 5.3: graphical plots of data given in Tables 5.1 and 5.2. 100

10 8

7

9

100

3

10

2

90

1

90

8

6

7

9

4

3

80

80

70

70

5 2 4

5

60

60

50

50

40

40

30

30 0

5

10

15

20

0

25

Figure 5.3a: Mean AP% in OIS 5c3 against longitude

80

5

10

15

20

25

Figure 5.3b: Mean AP% in OIS 5a against longitude

80

9

10 70

70 10

7

60

60

5

9 8

50

50 5

7

8

40

40 6

30

30 4

20

4

20

10

3

10

2

1

2

3 0

0 0

5

10

15

20

0

25

Figure 5.3c: Mean AP% excluding Pinus and Betula in OIS 5c3 against longitude

5

10

15

20

25

Figure 5.3d: Mean AP% excluding Pinus and Betula in OIS 5a against longitude

80

80

70

70

10

60

60

5

10

50

50

8 5

40

40

30

30

9 8

20

20

4

9

4

10

10 7

3

7

2 61

0

5

10

15

20

0

25

Figure 5.3e: Mean AP% excluding Pinus, Betula, Abies and Larix in OIS 5c3 against longitude

2

3

0

0

5

10

15

20

Figure 5.3f: Mean AP% excluding Pinus, Betula, Picea, Abies and Larix in OIS 5a against longitude

75

25

Figure 5.3 continued 100

10

9

7

100

3

8

10

21

90

7 9 8

90

6 4

3 80

80

5 70

2

70

5

60

60

50

50

40

40

4

30

30 45

46

47

48

49

50

51

52

45

53

46

47

48

49

50

51

52

53

Figure 5.3g: Mean AP% in OIS 5c3 against latitude Figure 5.3h: Mean AP% in OIS 5a against latitude

80

80

9

10

70

70 10

7

60

60

50

50

9

5

8 7

5

8 40

40 6

30

30 4

20

10

4

20

3

10

21

3

0

2

0

45

46

47

48

49

50

51

52

53

45

Figure 5.3i: Mean AP% excluding Pinus and Betula in OIS 5c3 against latitude

46

47

48

49

50

51

52

53

Figure 5.3j: Mean AP% excluding Pinus and Betula in OIS 5a against latitude

80

80

70

70

60

10

60 5

10 50

50 8 5

40

40

30

30 9 8

20

20 9

10

4

4 7

6

10 3

21

51

52

0

7 3

0

45

46

47

48

49

50

53

45

Figure 5.3k: Mean AP% excluding Pinus, Betula, Picea, Abies and Larix in OIS 5c3 against latitude

46

47

48

49

50

51

2 52

Figure 5.3l: Mean AP% excluding Pinus, Betula, Picea, Abies and Larix in OIS 5a against latitude

76

53

from Figure 5.3a. However it is highly questionable to regard the mean AP% value from Watten as low; it is much higher than that at Oerel OE61 in northern Germany which, when subjected to the same analysis, produces a mean AP% of 70.94% (from published diagram Behre and Lade 1986 Tafel 2; see Appendix 1 Figure A1.11b). No clear relation between AP% and either latitude or longitude could therefore be claimed in OIS 5c3 on the basis of these results.

values for mean AP% excluding Pinus and Betula in OIS 5a than in OIS 5c3. The significance of this pattern in the absence of OIS 5a records from Kąty and Gołkow is, however, open to question, and Figure 5.3j indicates that the correlation between this index and latitude may also be stronger in OIS 5a than in OIS 5c3. The data from Scladina and Samerberg are of relevance here. The former in OIS 5a generates a value for this index that is a high outlier in terms of latitude but which conforms much more closely to trend in terms of longitude. At Samerberg (12° 12'E) mean AP% excluding Pinus and Betula falls from 61.78% in OIS 5c3 to 44.28% in OIS 5a, in stark contrast to the more westerly sites which, as has been already observed, show an increase in OIS 5a relative to OIS 5c3. One may conclude, then, that the contribution of trees other than the cool-tolerant taxa Pinus and Betula to woodland expansion in OIS 5a at European mid latitudes was, unsurprisingly, greater in the south than the north of the zone, but that there was also some tendency for increasing easterliness to militate against the development of these taxa more than had been the case in OIS 5c3.

In OIS 5a the relation between mean AP% and longitude (Figure 5.3b) appears virtually identical with, and thus equally as weak as, that in OIS 5c3. There does, however, seem to have been a rather stronger latitudinal effect on this index. Figure 5.3h shows site mean AP% in this interstadial forming northern and southern clusters, with the values in the southern cluster never attaining that of any site in the northern cluster. The gradient is not steep. Mean AP% at Sulzberg is, at 87.59%, comparable with the value of 80.96% obtained for Kittlitz over 3° 30' further north. Nevertheless the data do support the inference that, within mid-latitude Europe, a gradient of increasingly arboreal vegetation development from north to south was in operation in OIS 5a, that this latitudinal gradient was considerably more significant than in OIS 5c3, and that it was also more significant than any longitudinal gradient of tree cover in either interstadial.

The clearest indication of unequivocal west-east vegetation gradients in OIS 5c3 and 5a is to be found in the values for mean AP% excluding not only Pinus and Betula but also Picea, Abies and Larix, an index essentially of the contribution made to pollen spectra by truly thermophilous arboreal taxa. Figures 5.3e and 5.3f show that this contribution declines dramatically with easterliness in both OIS 5c3 and 5a. Watten again appears to be an exception with values for this index considerably lower (8.03% in OIS 5c3 and 10.89% in OIS 5a) than the curves described by the remaining sites might suggest, and the possibility that a part of the Montaigu event OIS 5c2 might be included in the analysed spectra attributed here to OIS 5c3 should be borne in mind. When this site’s latitude is taken into account, however, the strength of the longitudinal influence on its pollen spectra becomes clear. In OIS 5c3 mean AP% excluding Pinus, Betula, Picea, Abies and Larix at Watten exceeds that of Kąty, 1°25' to the south, by a factor of more than three, and is more than double that of Samerberg, 3°4' to the south. Kąty possesses no OIS 5a record, but this index at Watten is again over twice that of Samerberg in that interstadial. Also, the values from Watten are greater than those from the eastern site of Zgierz-Rudunki, just 1°2' to the north but 17° 12' to the east, by a factor of nearly four in OIS 5c3 and seven in OIS 5a. The values from Kittlitz, 12°11' to the east but just 18' to the north of Watten, are exceeded by a factor of 2.6 in OIS 5c3 and almost 12 in OIS 5a. From Samerberg (12°12'E) eastward, thermophilous trees never contribute more than 4.27% of total pollen in either interstadial. The severity of this eastwest gradient in thermophilous tree development is probably attributable to the exacerbation of a general trend towards greater climatic continentality in these interstadials (van Andel and Tzedakis 1996), by a shortening of the summer growth season and a decline in spring precipitation in the more climatically continental east. Whatever the underlying cause, the impact of

With respect to the percentage of pollen from arboreal taxa other than Pinus and Betula, the data again suggest little longitudinal influence in OIS 5c3. Figure 5.3c could, like Figure 5.3a, be interpreted as revealing a weak longitudinal gradient if Watten and Scladina are ignored, but this would be contentious for the reasons given earlier. In any event these exclusions would simply produce a data plot with two clusters, one with high values of AP% excluding Pinus and Betula, and the other with low, the former consisting entirely of the more southerly sites. That the more important factor in the geographical patterning of this index in OIS 5c3 was latitude is apparent from Figure 5.3i, in which there is a clear, if variable, decrease in values with latitude. It should be noted that, in this plot, both Watten and Scladina are associated with values that conform to the gradient, as opposed to their outlier character in longitudinal plot Figure 5.3c. This is further evidence that, at mid latitudes in OIS 5c3, latitude exerted a more powerful influence than longitude on the development of trees less tolerant of cold than Pinus and Betula. In OIS 5a the correlation between mean AP% excluding Pinus and Betula and longitude (Figure 5.3d) appears somewhat closer then in OIS 5c3. Values cluster more closely around a negative gradient and those from ZgierzRudunki and Kittlitz, the most easterly sites, are both less than half that which they attain in OIS 5c3 (2.83% as compared with 4.86% at the former and 1.18% as opposed to 5.59% at the latter). At the same time the four most westerly records, Watten, Scladina, Les Echets G and La Grande Pile XX, all show significantly higher

77

longitude on the development of thermophilous trees in the OIS 5 interstadials is very clear, and indeed seems to increase in OIS 5a. Figures 5.3k and 5.3l, by contrast, reveal no correlation between mean AP% excluding Pinus, Betula, Picea, Abies and Larix and latitude in either OIS 5c3 or OIS 5a.

1989; van Andel and Tzedakis 1996) separating the OIS 4 and OIS 2 pleniglacial maxima. That there is and remains disagreement as to how these events in climate history should best be categorised or divided into meaningful blocs illustrates the fundamental feature which, from a scalar point of view, distinguishes them from OIS 5; namely, the collapse following OIS 5a of a clearly defined global sequence of warm-cold climatic oscillations with well-expressed and synchronous correlates in terrestrial palaeoenvironmental proxy records including the pollen record. Instead OIS 4 and 3 (OIS 2 need not be explicitly considered here since it postdates the Middle Palaeolithic) appear as a period of time in which general but ill-defined and often contradictory trends are visible but synchroneity between proxies, and between regions in the European pollen record, is impossible to establish. OIS 4 and 3 are treated here as a single unit in view of the virtual invisibility, discussed below, of any boundary between the two in the pollen record of Europe.

Vegetation Development in OIS 5c and 5a: Conclusions In OIS 5c and 5a, then, interstadial woodland development in Europe was clearly subject to north-south gradients much steeper than had been the case in the preceding interglacial. Although the pollen record affords little in the way of evidence for east-west vegetation gradients in Europe north of 52°N and south of 45°N in these interstadials, between these latitudes there is good palynological reason to believe that thermophilous tree development was subject to just such gradients. In midlatitude Europe woodland succession was more highly developed in the west than the east, with pine-birch forests strongly dominant east of around 12°E. As a consequence of these gradients, although European vegetation development in early Würm/Weichsel/Valdai retained the temporal synchroneity apparent in palaeoclimatic records, even as far as exhibiting a virtually pan-continental expression of the Montaigu event, the pan-continental geographic scale of successional homogeneity so characteristic of OIS 5e was lost. Instead vegetational landscapes in OIS 5c and 5a were more fragmented and more regionalised, with quite different vegetation regimes developing in different parts of the continent along both latitudinal and longitudinal axes. Insofar as human repertoires of socially transmitted knowledge might have been coupled with those scalar aspects of vegetational landscapes that varied with the character of tree cover – visibility, ease or difficulty of transit, herd aggregation and dispersal, mosaic spatial grain, species diversity and availability of edible fruit, for example – this regionalisation could potentially have underwritten the regionalisation of lithic industries visible in the Upper Pleistocene Middle Palaeolithic record of Europe (Bosinski 1982; Mellars 1996, 342-54) although this need not imply any simple causal relation between ‘types’ of environment and lithic industry.

OIS 4 and OIS 3 in the Oxygen Isotope Record The OIS 5a-OIS 4 boundary is placed at 74 kya in the orbitally tuned marine oxygen isotope chronology (Martinson et al 1987), and the OIS 4-OIS 3 boundary at 59 kya. The duration of OIS 4, at around 15,000 years, is comparable with that of the sub-stages of OIS 5 and can be seen, like them, as an expression of the 21,000 year orbital precessional cycle. In the marine oxygen isotope curve the OIS 4 maximum represents an isotopic enrichment peak (temperature minimum) of some 0.60.7‰ relative to the depletion trough at the OIS 5a apogee; relative to OIS 5e the enrichment is in the region of 1.2‰. This implies a decline of around 10˚C in mean annual temperatures from the OIS 5a temperature maximum (Lowe and Walker 1997, 333) and possibly of as much as 15˚C by comparison with the last interglacial. OIS 3, beginning at 59 kya and ending at 23 kya, is identified in the stacked marine oxygen isotope curve as a period of isotopic lightness intervening between the enrichment peaks of OIS 4 and OIS 2. It is not comparable in magnitude as a depletion trough with any of the warm sub-stages of OIS 5. In fact the interval appears in the stacked curve essentially as a baseline continuation of OIS 4 δ18O values, punctuated by shortlived, ill-defined and low amplitude isotopic depletion episodes. These never constitute an excursion of more than 0.3‰ towards isotopic lightness relative to the OIS 4 enrichment maximum, and usually much less (Figure 5.2). In some individual deep sea cores such as East Pacific core V19-30 (Shackleton et al 1983; Shackleton 1987) the isotopic oscillations are rather better defined. Still, in the V19-30 curve, in which six such isotopic depletion troughs/temperature maxima are visible within OIS 3, isotopic weight never falls to the levels even of the OIS 5d and 5b cool stadials, while the isotopic enrichment peaks separating these excursions are all greater than or equal to that at the peak of OIS 4. The

OIS 4 and OIS 3: Localisation and Asynchrony The OIS 4 first Würmian/Weichselian pleniglacial maximum and the ensuing OIS 3 together constitute the Middle Würm/Weichsel/Valdai according to some authorities (e.g. Donner 1996; Krzyszkowski 1990), distinguished from the preceding warmer OIS 5 and the succeeding very cold last glacial maximum OIS 2. For other authorities OIS 4, 3 and 2 together comprise the last pleniglacial (e.g. Behre 1989; Behre and van der Plicht 1992; Pons et al 1992), while others describe OIS 3 alone as the interpleniglacial or Middle Glacial (e.g. Guiot et al 78

status of OIS 3 as an interstadial is therefore problematic. Rather, the marine oxygen isotope record suggests that it should be understood as a cold period interrupted by short lived warm episodes.

under examination, by ill-defined climate ameliorations. The fact that OIS 4 in the GRIP and GISP2 oxygen isotope records incorporates D-O events (19 and 20) but, unlike OIS 3, shows no shorter-lived isotopic depletion troughs in the stacked marine oxygen isotope curve, is a further indication of the decline in OIS 4 and 3 of the cross-proxy synchroneity typical of OIS 5.

The significance of the Greenland GRIP and GISP2 ice core oxygen isotope records for our understanding of climate history in OIS 4 and 3 has already been touched upon. It reveals the operation of high amplitude D-O warming events, from D-O event 20 at 74 kya to event 3 at 25 kya inclusive – i.e. in OIS 4 and 3 – there are as many as 18 such events. (Dansgaard et al 1993). These ‘interstadials’ lasted between 500 and 2000 years and typically consisted of a very abrupt onset in which temperatures above the ice sheet increased by around 58°C (i.e. isotopic depletions of around 4-6‰) in 50 years, perhaps even as little as 10 years (Johnsen et al 1992), followed by a rather slower stepwise cooling. Temperatures at the peaks of these warming events reached levels some 5-6°C below today. The curves from the Antarctic Vostok (Jouzel et al 1987; Petit et al 1999) and Tibetan Guliya (Thompson et al 1997) ice cores are less informative on this matter; still, there are unambiguous isotopic enrichment peaks/temperature maxima within OIS 4 and 3 in both curves.

OIS 4 and OIS 3 in Other Palaeoclimate Records The millennial scale climate fluctuations visible as D-O events in the Greenland ice core oxygen isotope record are also apparent in other palaeoclimatic proxies referable to OIS 3, including the atmospheric dust records from the GRIP and GISP2 ice cores. Recent evidence from marine core ENAM93-21, drilled to the north east of the Faroe islands in the eastern North Atlantic, shows that not only N. pachyderma (s.) but also the benthonic cold water foram Cassidulina teretis and ice rafted debris (IRD) values oscillate within OIS 3 on D-O wavelengths (Rasmussen et al 1997). Beyond the North Atlantic, millennial scale climate oscillations in OIS 3 have now been demonstrated in the magnetic susceptibility curve from the Lac du Bouchet (Thouveny et al 1994), in marine sediments from the Arabian Sea (Schultz et al 1998), in planktonic foram curves from the Benguela Upwelling in the eastern South Atlantic offshore from Namibia (Little et al 1997), in Chinese loess sequences (Chen et al 1997) and in the structure of dunes in the Kalahari (Stokes et al 1998). As discussed in Chapter 3, there seems to be no reason to believe that the D-O events recognised in the GRIP and GISP2 oxygen isotope curves were an exclusively North Atlantic phenomenon, and every reason to accept their global character in OIS 3

The systemic causes underlying D-O wavelength climate fluctuations in the Upper Pleistocene are not well understood. It appears that feedback relations between North Atlantic deep water production, the thermohaline pump (the process whereby surface waters enriched in salinity by evaporation sink and thereby drive major ocean circulation phenomena such as the Gulf Stream), the latitudinal boundaries between polar and temperate sea surface temperatures, ice sheet extent and northern hemisphere atmospheric circulation are implicated in these fluctuations, although there may be no single factor that might be regarded as a controlling prime mover (Lehman 1993). Attempts to establish a causal link with major episodes of ice sheet collapse in the North Atlantic (Bond et al 1993) as evidenced by layers of ice rafted debris (IRD) in marine cores – so-called Heinrich Events – have not commanded widespread support among oceanographers and palaeoclimatologists. Whatever the causes of D-O wavelength climate instability, the phenomenon does not seem to have been restricted to Greenland, the North Atlantic or even the northern Hemisphere; neither does it seem to be unique to the Upper Pleistocene. Recent evidence shows that similar high amplitude climate fluctuations on such wavelengths were global in reach and were a pervasive and persistent feature of Pleistocene climate, as discussed in Chapter 3 and below.

Sea level experienced a eustatic fall of more than 60m across the OIS 5a/4 boundary (Shackleton 1987), an unequivocal indication that OIS 4 was a period of significant decline in global temperatures. Further corroboration is to be found in evidence for the expansion of the Fennoscandian ice sheet to the west, south and south east (Drozdowski and Fedorowicz 1987; HoumarkNielsen 1989). Whereas the decline in global mean temperatures in OIS 4 from the OIS 5a maximum was in the region of 10°C according to the marine oxygen isotope record, at Clear Lake, California, the fall in mean annual temperature estimated from palynological data was 7°C (Adam and West 1983), and the decline in sea surface temperatures in the southern North Atlantic was in the region of 5-6°C (Eglinton et al 1992). North-south temperature gradients in OIS 4 seem therefore to have been even steeper than those in OIS 5c and 5a.

The Greenland ice core oxygen isotope curves, then, clearly indicate why it is so difficult to describe OIS 3 as either an interstadial or as a cold period. Considered on millennial timescales it is neither, but a sequence of alternating warm and cold events. This is manifest on longer Milankovitch timescales as a cold period interrupted variably, depending on the particular record

It is not clear that the Fennoscandian ice sheet receded with the onset of OIS 3. According to Larsen et al (1987) the maximum extent of the ice sheet occurred some time after 63 kya, i.e. close to the beginning of OIS 3 at 59 kya, and there was no significant regional deglaciation until 54 kya. Yet OIS 3 begins with an increase in ice rafted debris in the Norwegian Sea, evidence for a 79

significant collapse of the ice sheet margins (Baumann et al 1995). In fact sea level across the OIS 4/3 boundary rose by 25m or more, reaching a level some 50m below that of today, before slowly receding to –80m by 30 kya (Shackleton 1987; van Andel and Tzedakis 1996), implying an associated decline and subsequent increase in global ice volume. No such slow descent into full glacial conditions in the course of OIS 3 is, however, visible in the European permafrost record. The boundary between continuous and discontinuous permafrost in Europe receded northeastwards between 60 and 45 kya, leaving only Scandinavia, the Baltic region and mid-latitude eastern Europe in the grip of continuous permafrost (van Vliet-Lanoë 1989). The enigmatic character of OIS 3 is again apparent in these cross-proxy comparisons.

Grande Pile. The analysts identify an interstadial in zone 21 and suggest that this event may correspond to zone F7 at Les Echets and Ognon I at LGP X, but above this recognise only two brief warming events before the onset of the OIS 2 glacial maximum (Reille and de Beaulieu 1990). This breakdown of regional synchroneity in OIS 4 and OIS 3 is even more clearly expressed at the continental scale. Zones P2k-P2o in the Padul 2 core consist of fundamentally cold and arid pollen spectra within which there is a succession of weakly expressed episodes of uncertain number and palaeoclimatic significance, and zone P2k (corresponding to OIS 4) includes no warming event comparable with the Ognon I event at La Grande Pile (Pons and Reille 1988). Velichko describes OIS 3 in European Russia as the ‘Middle Valdai Interstade’ with ill-defined events (Velichko 1984). By typifying OIS 3 as an interstadial Velichko implies that it was essentially a mild period climatically, in sharp contrast with de Beaulieu and Reille’s interpretation of the same period in France. At Tenaghi Philippon Wijmstra identifies no interstadial events in zone V (OIS 4) (Wijmstra 1969, 519). The Valle di Castiglione core lacks the first part of OIS 4, but zones VdC16 and 15, which comprise OIS 3 and the later part of OIS 4, feature four major peaks in AP%, with contributions from thermophilous species including evergreen oaks (Follieri et al 1989).

OIS 3 lasted for some 36,000 years before the onset of the last pleniglacial at 23 kya on the orbitally tuned timescale. From a hierarchical point of view this is of the first importance, since it signals the end of the prevailing dominance of the 21,000-year precessional cycle in which well-defined warm and cold events of around 10,000 years duration alternated and upon which the system synchroneity characteristic of OIS 5 was predicated. In OIS 3 this wavelength of Milankovitch cyclicity in climatic constraints on life systems is lost and replaced by a much more persistent cold period. If OIS 4 is taken into account the period of cold climate lasted for around 50,000 years. The vegetation history of Europe in OIS 4 and 3 must therefore be understood as structured primarily by sub-Milankovitch scale climate shifts against a background of orbitally-forced climate stability lasting some 50 millennia. This transformation in the scalar character of climatic fluctuations relative to OIS 5 has profound implications for terrestrial ecosystems in OIS 3 Europe.

In northern Europe the problem is compounded by an absence of pollen sequences covering the whole of OIS 4 and OIS 3; however, a number of interstadial sequences from these periods have been recognised in pollen cores from the region. At Oerel two interstadials overlying OIS 5a/Odderade have been identified and named Oerel and Glinde (Behre and Lade 1986; Behre 1989). Radiocarbon dates for these episodes place the former at 58-54,000 BP and the latter at 51-48,000 BP (Behre and van der Plicht 1992), but these must be regarded as minimum dates. In these circumstances it is simply impossible to place Oerel in particular on either side of the OIS 4-3 boundary with any confidence. Their correlation with other palynological interstadials in the region is highly problematical. Behre suggests that the older interstadial, Oerel, might be synchronous with Ognon II at La Grande Pile, the reality of which has been questioned by de Beaulieu and Reille. Behre believes both Oerel and Glinde to be older than the OIS 3 interstadials defined in the Netherlands, but points out that the latter – Moershoofd, Hengelo and Denekamp from oldest to youngest – have in fact been constructed from ‘complexes’ of local, short lived and even asynchronous episodes of shrub-tundra development or peat formation, the lithological boundaries of which are unclear, which never occur together in a single core, and which have been ‘clustered’ into these three episodes solely on the basis of a scatter of radiocarbon dates (Behre 1989, 41). It seems likely, then, that the Dutch OIS 3 pollen chronology, despite its widespread use throughout Europe (Chapter 4) is in fact a contentious amalgamation of ill-defined, localised events whose proximate causes

The OIS 4 and OIS 3 Pollen Record in Europe: Localisation, Asynchroneity and Migration Lag The most visible consequence of this transformation on the European pollen record in OIS 4 and 3 is the loss of well developed interstadial forest and woodland spectra that can be stratigraphically and chronologically correlated on a continental scale. Indeed, even regional and local correlations are fraught with difficulty. Woillard (1979) identified seven temperate interstadial episodes in that part of the La Grande Pile X curve now regarded as corresponding to OIS 4 and 3 – the Ognon 1, 2 and 3 interstadials within the Lanterne I zone, and the Goulotte, Pile, Charbon and Grand-Bois ameliorations in the overlying Lanterne II. In boring XX de Beaulieu and Reille (1992a) recognised only the four Lanterne II warming events. At Les Echets, the same analysts correlate zone F7 with Ognon I at LGP X but the overlying zones G and M, corresponding chronologically to Lanterne I and II at LGP X, feature only three episodes that might be termed interstadials, none of which are well defined (de Beaulieu and Reille 1984). At Lac du Bouchet D there is still less comparability with La 80

lay in local edaphic conditions and vegetation history rather than in climate change. de Beaulieu and Reille make this point with particular clarity in relation to the comparison between the La Grande Pile and Les Echets pollen sequences:

mammoth steppe ‘steady state mosaic’, and the duration of climatic events. The larger the spatial extent of the area under consideration (conceived in terms of the distances between seed sources or refugia and localities potentially capable of being colonised by species from these seed sources), the longer the characteristic minimum wavelength of climatic fluctuation to which the species composition of the vegetation community as a whole is capable of responding becomes. Proximity to refugia was therefore a key factor in determining the propensity of vegetation in any particular region of Europe to approach woodland or forest in a warming lasting just several centuries or a very few millennia. It is now clear that the refugia for mesophilous and thermophilous trees in the cold stages of the European Pleistocene were located in the mid-altitude zones of the mountains of southern Europe, most particularly the western Balkans and the Italian Apennines (Bennett et al 1991), with the Alps and Carpathians playing a subordinate refugial role (van der Hammen et al 1971; Huntley and Birks 1983). Warm events lasting ten millennia or so, such as those of OIS 5 and the Holocene, were clearly sufficiently long to permit the migration even of thermophilous trees as far north as their ecological tolerances permitted, leading to plant community transformation on a pan-continental spatial scale. The impact of the much shorter D-O warming events in OIS 4 and OIS 3 will now be addressed by a comparison of the vegetation histories of the various regions of Europe in those periods.

…it should be noted that during climatically not wellcharacterized periods – that Welten (1981) refers to as ‘interphases’ – there must have existed gradients establishing local thresholds favourable to the transitory and short expansion of arboreal vegetation types whose development was mostly dependent on the situation and importance of refuges inherited from a regional past. So, the point is not to explain the fact that six interstadials were recognised at La Grande Pile, whereas only three are found at Les Echets: it is probable that in both sites these episodes cannot be used as biostratigraphical references, as during the period considered, the local character of the vegetation may have been very different from one site to the other. (de Beaulieu and Reille 1984, 126)

The difficulties encountered in attempting to make crosssite correlations of OIS 4 and OIS 3 warm events are therefore not merely technical in character; rather, they are expressions of the breakdown of continental and even regional coupling between climate change and vegetation development in OIS 4 and OIS 3 and a consequent loss of synchroneity and increase in localisation. An ‘interstadial’ at one site might have been contemporaneous with a ‘stadial’ at another.

Vegetation Histories of the Regions of Europe in OIS 4 and OIS 3

However, de Beaulieu and Reille’s attribution of this pattern to the low amplitude nature of climatic ameliorations in OIS 4 and OIS 3 seems difficult to sustain in view of the ice core oxygen isotope evidence, which clearly shows that climate events in these periods were of high amplitude in terms of shifts in mean annual temperatures. A better explanation is that these ameliorations were short lived punctuations within a long period of cold climate. In these circumstances the course of vegetation history at any locality could be expected to proceed from the basis of the local and regional species pool established during the pleniglacial cold of OIS 4, and vegetation development would be strongly influenced by proximity to seed sources. If seed-producing trees, for example, were far distant, then a climatic amelioration lasting 1000 years might have ended before woodland was established locally and regionally. That is, migration lag must be taken into account. Equally, a climatic deterioration, if sufficiently short lived, need not have resulted in the local and regional extinction of mesophilous and thermophilous taxa providing that the temperature decline was not so great as to kill individuals outright but instead merely inhibited the optimal expression of their life cycles, as with the Cambridge evergreen oak of Chapter 2 that lost its leaves in the cold January of 1982.

North Western Europe Interstadial pollen spectra of OIS 4 and OIS 3 age are known from northern Germany, Belgium, Denmark and, especially, the Netherlands. Leaving aside the problems of correlation already discussed, a clear regional picture emerges; variation in pollen frequencies of the various taxa are slight and interstadial events are invariably treeless. At Oerel OE61, the only site from which pleniWeichselian (i.e. OIS 4 and OIS 3) interstadials conformably overlie the OIS 5 interstadials, the older Oerel and younger Glinde interstadials both represent shrub-tundra landscapes. In the Oerel interstadial, which occurs above a sterile sand attributed to the Schalkholz stadial, NAP% averages 81.3%. Betula pollen reaches a peak of 23.6% but macrofossils indicate that this was generated entirely by the dwarf birch Betula nana. Willow (Salix), Juniperus and heathers (Ericales) are also present, together with a rich spectrum of herbaceous species. One may conclude that, despite the absence of trees, temperatures in Oerel were considerably higher than in Schalkholz. In the peat of the Glinde interstadial, which is separated from the Oerel event by a zone of sterile sand (the Ebersdorf stadial), NAP% averages 78.2%. Betula and Salix are better represented than in the Oerel event but

The crux of the matter, then, is the scalar relations between the cyclical upgrade-downgrade dynamics of the

81

Juniperus and Ericales pollen is rarer. The Glinde is overlain in OE61 by sterile layers of gravel and sand. Pinus averages just 2.9% in Oerel and 0.95% in Glinde, indicating that the nearest pine stands were some distance away from the locality, possibly as far south as the Alps (Behre 1989). It seems, then, that in north western Germany landscapes in OIS 4 and probably the earlier part of OIS 3 were barren periglacial deserts and that climatic amelioration led to brief episodes of open shrub tundra formation.

and the possibility that this has generated AP% peaks lacking palaeoclimatic significance, must be addressed. Woillard’s seven interstadials in Lanterne I and II at La Grande Pile X are open to challenge on these grounds, with de Beaulieu and Reille considering that Woillard’s Ognon II and III events in La Grande Pile X core are products of this process. Equally, they are sceptical about correlations between other Lanterne interstadials visible in the various LGP borings on account of the homogeneity of the pollen spectra and the low amplitude variations in taxon frequency. They conclude that, when long-distance pollen transport is taken into consideration, almost all of the fluctuations in AP% in LGP XX zone 8 (late OIS 4 and OIS 3) can be discounted as palaeoenvironmental and palaeoclimatic indicators, and that only the Pile event (zone 8c) unequivocally demonstrates the regional presence of mesophilous trees. Even in this event AP% averages only some 40%, exceeding 50% only briefly, and arboreal taxa other than Pinus and Betula are restricted to Picea and deciduous Quercus, neither of which exceed 5%. de Beaulieu and Reille interpret the vegetation history of the Vosges in OIS 4 and OIS 3 as a persistent steppe-tundra landscape punctuated by poorly defined expansions of sparse Betula and Pinus woodland or of improved pollen production by these taxa in periods of moderate climatic improvement (de Beaulieu and Reille 1992a, 433-5).

This pattern is confirmed by pollen cores from other sites. Spectra attributed to the putative Moershoofd interstadial complex (50-42 kya; Zagwijn 1989) in the Netherlands, such as Denekamp core 20a (van der Hammen 1971), Hengelo 1 and 2a (Zagwijn 1974), Duckenburg (Teunissen and Teunissen-van Oorschot 1974), Voorhuitzen and Eerbeek (Kolstrup and Wijmstra 1977) all indicate open vegetational landscapes similar to that of the Glinde interstadial at Oerel OE61, with high values of heliophytes such as Artemisia and Helianthemum. The same is true of spectra placed in the putative Hengelo and Denekamp interstadials (39-36 kya and 32-28 kya respectively; van Andel and Tzedakis 1996, 494), including those from Ruigekluft, Hengelo (Zagwijn 1974), Bussloo and Laarhuis (Kolstrup and Wijmstra 1977). In these the landscape was apparently yet more open, with Betula nana still less important than in either Glinde or Oerel. Unreliable radiocarbon dates preclude the correlation of similar events known from the pollen sequences of Brugge, Belgium (Vandenberghe et al 1974) and Sejerø, Denmark (Houmark-Nielsen and Kolstrup 1981) with the Dutch complexes, but both yield unequivocal evidence for treeless interstadials in OIS 3.

The Les Echets G pollen core provides a very similar picture to La Grande Pile. Pollen zones F7, G, H, I, J, K and L are dated to OIS 4 and OIS 3, and in this region of the core AP% fluctuates rapidly, primarily in consequence of fluctuating Pinus values. Even so, AP% is generally low, rarely surpassing 50% and never more than 80%. De Beaulieu and Reille (1984), however, do not interpret this as evidence for pulses of forestation and deforestation; instead they argue that the constancy of the floristic composition excludes significant changes in plant communities in this period and that the variations in Pinus% are the result of ‘sudden and repeated changes in pollen transport and deposition in the lake’ (de Beaulieu and Reille 1984, 122). Against this background they identify spectra indicative of real local and regional vegetation response to climate warming in pollen zones F7 (which they equate with Ognon I at LGP X), H, J and L. In F7 AP% attains values of around 75% and, in addition to Pinus and Betula, Picea, Alnus and deciduous Quercus are present at levels of 10%, 90%. Picea dominates at 6065% with pine at 30-40%; Betula, Alnus and Larix are also present. Grass and sedge pollen each decline to 40,000 BP (in reverse stratigraphic order) from pollen zones T1 and T2 respectively, and of 10,190 BP from the zone X1 much higher in the sequence, indicate that the sequence extends into OIS 3 at least as far as 40 kya. A major forested event is indicated by AP% maxima of >90% and ca 85% in zone T1, tentatively correlated with V1 and V2 at Xinias respectively. In zone T2 AP% falls to around 35% but it recovers to a level of around 90% in T3, an episode that might be equivalent to Xinias V3. The amplitude of younger oscillations is somewhat less, as at Xinias, but nevertheless exceeds those at the latter site.

Southeastern Europe Numerous studies have shown that warmth-demanding tree taxa survived through Pleistocene glacial periods in the mountains of south eastern Europe (Huntley and Birks 1983; Bennett et al 1991; Tzedakis 1993; Willis 1994). In the Balkans 60% of the land surface is more than 1000m above sea level, providing great altitudinal and topographical diversity, together with winter snows and summer adiabatic rainfall, in a region that was always ice free. Pollen sequences from southeastern Europe therefore exhibit a continuous and significant presence of both coniferous and broad leaved trees through glacial periods, including OIS 2 (Willis 1994; Tzedakis 1994). One would therefore expect to see expansions of coniferous and broad-leaved woodland from local refugia in synchrony with millennial-scale climatic improvements. Although many pollen records exist from south eastern Europe only a handful extend into OIS 4 and OIS 3. Nevertheless, these confirm this view. The sequence from Lake Xinias, at an altitude of 500m in Thessaly, south central Greece, extends from the Holocene to around 50 kya. Radiocarbon dates of 46,900+5000–3060 BP for the basal zone V1 and 25,620±400 BP for the centre of zone X have been obtained (Bottema 1979). The sequence between these points therefore relates to the greater part of OIS 3 and may even incorporate part of OIS 4 if the radiocarbon date is minimal. Zone V is divisible into three sub-zones V1, V2 and V3, each of which represents a major increase in arboreal pollen frequency. V1 commences with AP% at 55%; values decline thereafter, implying that arboreal pollen in this zone is declining from a peak not represented in the core. Pinus is the most important taxon at 34.1% but evergreen oak reaches >20%and a wide range of thermophiles is also present. AP% falls to 30% but climbs to 65% in V2. Evergreen oak attains a level of 35% whereas pine rises only to 20%, with synchronous increases in Carpinus, Corylus,

The west-east decline in OIS 3 forest development in Greece is confirmed by the record from Tenaghi Philippon. Zone V marks the end of a long sequence of moderately forested phases attributed to OIS 5. AP% is very low (10% or less) and chenopods and Artemisia dominate; this zone can be assigned to OIS 4. In the overlying zone P, there are numerous low-amplitude oscillations in AP%, although this never exceeds 30%. These fluctuations are primarily driven by Pinus pollen frequencies, which reach twin maxima in sub-zones P1 and P3, and a single maximum in P5. Quercus is continuously present in P1-P5, but shows twin maxima of around 10% each in P3 and a smaller maximum in P5, each coincident with a rise in Pinus and AP%. Oak then disappears from the record in P6, showing only occasional low frequency presence until zone Y, corresponding to the beginning of the deglaciation (Wijmstra 1969). Radiocarbon dates and presumed sedimentation rates place the onset of zone P at 39,000 and its end at 30,000 BP (Wijmstra 1969, 523), although 87

these, especially the former, should be regarded as minimum dates. Indeed, Bottema cites a date of 49,070 BP for the base of P1 and 32,140 BP for P5 (Bottema 1979, Figure 7). Wijmstra identifies three interstadials within zone P. The twin AP% maxima of sub-zone P1 he names Heraklitsa and dates to between 42 and 39,000 BP, but, given the minimal nature of the radiocarbon dates and the position of sub-zone P3 towards the base of the OIS 3 sequence a real date of around 50,000 BP for Heraklitsa seems more likely. The twin Quercus peaks of P3 he calls Kalabaki and the P5 Quercus maximum he terms Krinides. As at Xinias and Ioannina, then, the amplitude of AP% maxima in OIS 3 at Tenaghi Philippon

is greatest in its earlier part and then declines, although arboreal pollen is never as important at Tenaghi Philippon as at the two more westerly sites. Macedonia’s rain shadow location is the primary explanation for this pattern, but the low altitude of Tenaghi Philippon – 40m above sea level compared with Ioannina I and Xinias each at 500m - should also be emphasised. Of all the regions of Europe, then, Greece affords the strongest evidence for regionally synchronous vegetation events and trends in OIS 3, although the amplitude of their expression varied with longitude and altitude. Summary pollen diagrams from the three sites are shown in Figure 5.5.

Figure 5.5. Summary pollen diagrams of OIS 3 sequences from Ioannina I, Xinias and Tenaghi Philippon, showing curves for Quercus (solid line), Pinus (dotted line), AP/NAP, Artemisia (stippled), chenopods (fine hatching) and other herbs (coarse hatching). After Bottema 1979, Figure 7

88

The OIS 4 and OIS 3 pollen record in the Balkans other than Greece is much less well known. There are no records extending beyond the last deglaciation at around 10,000 years ago in Croatia, no reliably dated records in Romania and none at all in Albania, European Turkey, Yugoslavia or Bosnia (Willis 1994). However, at Ljubljana Moor, at 300m elevation in western Slovenia, spectra from the last glacial (i.e. subsequent to the OIS 5 interstadials) indicate landscape alternations between steppe mosaics with Pinus and Picea on the one hand and open mixed broad-leaved woodland on the other, with thermophilous taxa continuously present (Šercelj 1966). The record from Suchero Ezero, in the Rhodopes Mountains of southern Bulgaria, yields a narrower range of taxa, as one would expect from a site at an altitude of 1900m above sea level. Nevertheless Quercus, Alnus and Fagus show periodic expansions (Willis 1994; Bozilova and Smit 1982).

Discussion This investigation of Upper Pleistocene climate and vegetation history in Europe clearly indicates that, following high levels of synchroneity at pan-continental spatial scales in the OIS 5e interglacial, the character of vegetational landscapes tended towards increasing regionalisation through the OIS 5d – 5a sub-stages and OIS 4 and 3. In OIS 5c and 5a the continental extent of Milankovitch-scale, warm-event tree expansion typical of OIS 5e continued, but steep north-south vegetation gradients emerged. In consequence the thermophilous woodlands of the south did not develop in the north during these interstadials; rather, northern woodlands were boreal in character and dominated by pine and birch with variable development of fir and spruce. In addition, the development of thermophilous trees at European midlatitudes in these interstadials declined strongly from west to east. Both the latitudinal and longitudinal vegetational gradients seem to have been somewhat steeper in OIS 5a than in OIS 5c.

Palaeoenvironmental reconstructions based on data from archaeological sites are also of some relevance. At Musselievo, northern Bulgaria, the leaf point bearing level FA is sandwiched between underlying reworked soils (upper level FX) tentatively dated to OIS 5 and overlying fossil soils (FB) possibly correlated with Heraklitsa, Kalabaki or Krinides at Tenaghi Philippon (see Chapter 4). There has been no pollen analysis of the sediments, but the stratigraphy suggests strongly that mature soils, and thus possibly woodland, developed episodically during OIS 3 in northern Bulgaria.

In OIS 4 and 3 the loss of the 20,000-year precessional wavelength of orbitally-forced climate change resulted in the loss of pan-European synchroneity in episodes of tree expansion. In the context of warm events of subMilankovitch, D-O duration local vegetation history and proximity to refugial seed sources became primary factors in local and regional woodland development, along with levels of moisture availability sufficient to permit the survival of saplings, a factor more readily met at higher altitudes. At the same time north-south vegetational gradients became even steeper than in the OIS 5 interstadials, with treeless interstadials in northern Europe. Only in regions close to cold-period tree refugia – conifers in the Alpine foreland and northern Carpathians, and both conifers and broad-leaved taxa in the Italian Apennines, southern Carpathians and the Balkans – did significant expansion occur in pulses comparable in duration and number with D-O scale warm events. Wooded interstadials of centuries or a very few millennia in duration during OIS 3 are evidenced in the pollen record only in the southeastern portion of Europe from Bavaria and the Alpine foreland to Italy, the Balkans and the Black Sea.

With regard to Romania, the limitations of Cârciumaru’s Upper Pleistocene cave pollen chronology (Cârciumaru 1979, 1988), in which Amersfoort, Brørup and Odderade are regarded as younger than the first Würmian pleniglacial maximum, have already been highlighted, and his proposed interstadial events (e.g. ‘complexe interstadial Nandru’ and ‘complexe interstadial Ohaba’) must therefore be regarded as provisional at best. At the same time a series of conventional radiocarbon dates for Middle Palaeolithic horizons from throughout Romania and performed at the Gröningen laboratory places the entire Romanian Middle Palaeolithic in OIS 3 (Honea 1984, 1986: Mertens 1996). A number of finite dates in the range 40-46,000 BP suggest that some at least of the pollen events described by Cârciumaru do indeed relate to OIS 3. In fact Pinus and Picea are present in all of these events, even in those considered glacial; in ‘interstadial’ events thermophiles are important, expanding from Carpathian montane refugia (Cârciumaru 1979, 25). However, although it is conceivable that pulses of afforestation occurred in the Carpathian mountains of Romania during OIS 3, it is much more difficult to argue for this at the northern site of Ripiceni-Izvor at just 100m altitude in the Prut valley and in the lee of the Carpathians. The OIS 3 leaf point levels IV and V are associated with treeless environments (Păunescu 1988a, 1993). Pine pollen is more important in the fossil soil of the uppermost Mousterian level III (Păunescu 1965), but this could easily have been derived from upland regions to the west by wind transport.

The northern limit of this zone of OIS 4 and OIS 3 forest pulses is unclear. However, if one considers that, during OIS 5c and 5a forest development was weaker in the east than the west of Europe at mid latitudes, as demonstrated earlier, and that there is no evidence for the weakening of vegetation gradients in OIS 4 and 3, then D-O interstadials in Germany and Poland are most unlikely to have been wooded at 52˚ N, the latitude of Hengelo where the interstadials were treeless. Even if Hengelo were the southernmost limit of treeless interstadials in these periods – an unlikely proposition given the paucity of arboreal pollen in the interstades at that site – then one would still expect that limit to have been located further to the south in more eastern parts of Europe. The 89

possibility that wooded interstadials occurred at Königsaue can be ruled out. Similarly, sedimentological studies undertaken on profiles from Wola Grzymalina, near Bełchatów, central Poland, indicate only tundra and desert environments in a sequence radiocarbon dated to between 43,700+3700-2400 and 31,800±700 BP (Krzyszkowski 1990). Taking into account the known latitudes of both treeless and forested interstadials in western and central Europe, the east-west vegetation gradient and the location of known refugial areas, one can conclude that the probable northernmost boundary of significant woodland expansion in response to D-O warming events in OIS 4 and OIS 3 is likely to have been the northern edge of the central German uplands at around 51˚N. To the east, the northern limit must have been further south; Moravia’s proximity to the Alpine ice sheet probably militated against OIS 4 and 3 forest expansion there, but it probably occurred as far north as the current Carpathian border between Slovakia and Hungary at around 49˚N. Yet further east, the Carpathian arc seems likely to have delimited the OIS 4 and 3 woodland zone, with a northern limit of around 47˚N in the lowland east of Romania.

that leaf points are not a feature of the Italian late Middle Palaeolithic, despite the clear evidence for OIS 3 wooded interstades in at least the upland regions of the peninsula. However, that does not detract from the association of OIS 3 Middle Palaeolithic leaf points with centennial and millennial scale oscillations between open and wooded landscapes; rather, it shows only that such oscillations did not necessarily induce the practice of leaf point production. Italy is in any case separated from the leaf point zone to the north and east by the Alps, which may have acted as a barrier to the transmission of the practice of leaf point fabrication. We are left, then, with a late Middle Palaeolithic leaf point zone unified by a particular temporal wavelength of radical environment transformation and, west of the Black Sea, physically linked by the Danube and the upland regions of its drainage basin. In sharp contrast to the Lower Palaeolithic insensitivity of socially transmitted lithic-technical knowledge to even highamplitude climate and environment change through glacial-interglacial cycles, the OIS 3 (and possibly 4) Middle Palaeolithic social practice of leaf point fabrication was explicitly coupled to relatively short wavelength, high process-rate cyclical transformations in the material world and its affordances. A systematic relation of this character between a stone tool type and a sub-Milankovitch rate of environmental transformation has not, to the writer’s knowledge, hitherto been demonstrated for the European Palaeolithic older than OIS 4. This suggests strongly that, whereas situationally appropriate applications of socially transmitted knowledge to landscapes changing on Milankovitch wavelengths induced no persistent transformation in the structured social repertoire of lithic-technical knowledge in the Lower Palaeolithic, and whereas the same was true of situationally appropriate responses to subMilankovitch scales of landscape transformation in the earlier Middle Palaeolithic, in the late Middle Palaeolithic knowledgeable lithic-technical responses to subMilankovitch scale environment change could become structurally institutionalised. The scalar environmental context of Middle Palaeolithic leaf points in Europe therefore indicates that the trend towards an increasing susceptibility of socially transmitted knowledge to transformation by knowledgeable action in the Middle (relative to the Lower) Palaeolithic as discussed in Chapter 3, intensified in the late Middle Palaeolithic. In Chapter 6 the landscape context of the Altmühl Valley leaf points, and the lithic industries with which they are associated, will be examined in detail with a view to illuminating this process and its implications.

This geographical distribution of wooded interstadials in OIS 4 and 3 Europe is of the greatest significance in that it is essentially identical with the geographical distribution of late Middle Palaeolithic leaf point sites. From Mauern in the north west to Jankovich in the north to the Greek leaf point sites in the south to Il’skaya in the east, these OIS 3 Middle Palaeolithic leaf point assemblages occur in regions where landscapes, and thus their affordances, were repeatedly transformed on subMilankovitch wavelengths. Even within the southern Balkans leaf points occur in Serbia, northwest Greece and the Rhodopes Mountains but not, as far as is currently known, in Greek Macedonia where OIS 3 interstadials were not forested. The environmental history of the northern Black Sea region, including The Crimea and The Caucasus, has not been considered here, largely because of the paucity of the palaeoenvironmental record in those regions. Speculation as to whether the late Middle Palaeolithic leaf points of that region shared a similar scalarenvironmental context is therefore not justified, although one should note that the Black Sea was, in the Upper Pleistocene, a below-sea-level wetland basin, fed by several major rivers, and its drainage basin to the north might have provided sufficient moisture to permit tree development in favourable conditions even in the absence of mountainous uplands. Of course, it must also be noted

90

CHAPTER 6

THE MIDDLE PALAEOLITHIC OF THE ALTMÜHL VALLEY at Wellheim, roughly equidistant between the Altmühl and the Danube, and flows eastsoutheast for 20km to its confluence with the Danube at Ingolstadt.

Introduction As was noted in Chapter 4, the Altmühl Valley, Bavaria, boasts the best-known cluster of sites with Middle Palaeolithic leaf points in Europe. Within the Altmühl Valley proper there are six such sites: Weinberghöhlen, Mauern; Obernederhöhle; Groβe Ofnet; Kleine Ofnet; Steinerner Rosenkranz, Mörnsheim; and Biesenhard. A further four are found in adjacent regions of the Upper Danube: Haldensteinhöhle, Urspring, to the west; and Zeitlarn, Albersdorf and Flintsbach-Hardt to the east. Finally, Kösten is situated some 135 km north of the bend in the Altmühl that marks its most southerly point, and connected directly to it by the courses of the rivers Rednitz, Regnitz and Main. This amounts to 11 Middle Palaeolithic leaf point sites in the region. In addition to the leaf point sites, there are, within the Altmühl Valley proper, 14 Middle Palaeolithic sites with Mousterian (i.e. biface-poor) or Micoquian/‘Jungacheuléen’ (biface-rich) industries, or both; there are a number of other such sites in adjacent regions of the Upper Danube, including the Bockstein localities.

The Altmühl Valley today is a national park, incorporating rolling hills at lower altitudes, together with more rugged country in the upland massifs. Much of the river in the Lower Valley has been incorporated into the Main-Danube canal which, further north, follows the courses of the Rednitz and Regnitz rivers; this indicates that those watercourses represent a favoured route between the Altmühl and Kösten, in the Upper Main. The national park authorities carry out a programme of regularly felling saplings since, following the collapse of sheep farming in the valley at the end of the 19th Century and the consequent decline in grazing, the park would otherwise become forested. It is for this reason that the only photographs of the Weinberghöhlen, Mauern, in which the cave entrances are not obscured by trees are those taken a hundred years ago (Weiβmüller, pers comm). Clearly, forestation can proceed quickly in this region when conditions are favourable. Geology and Geological History

The Altmühl Valley

The Frankische Alb is the upland region delimited by the Danube to the south, the Bavarian and Bohemian Forests to the east, the Thuringian Forest to the north and the Swabian Jura to the west. Geologically, it is a complex of Jurassic limestone formations that form a ‘dog-leg’, extending northwards from the Kelheim region to Nuremberg and eastwards to Nordlingen, where it is interrupted by the Ries meteorite impact crater. The Jurassic formations then continue south and west into Baden-Wurttemberg, where they are known as the Swabian Jura. Three distinct facies occur within the Jurassic complex; a Lower Jurassic component on its northern and western margins, a Middle Jurassic facies towards the centre, and a more extensive Upper Jurassic unit that constitutes the centre and the southern and eastern flank. To the north and west of the Jurassic formation the geology is Triassic in origin; it is in this region, extending north into the Thuringian Forest, that the Altmühl rises. The Altmühl therefore passes through progressively younger formations from its source to its confluence with the Danube, south of which the geology consists of unconsolidated Tertiary sediments. At Kelheim, and running northwards to the east of the Jurassic limestones, outcrops of Cretaceous quartzite occur.

Geography The river Altmühl rises in the northwest of the Frankische Alb at 49°30'N, 10°20'E and at an altitude of some 500 metres above sea level. From there it flows southeast for 85km, passing between massifs to the east and west that reach altitudes of some 700m, until it reaches its most southerly point at Dollnstein, 48°52'N, 11°6'E. The river then turns northeastwards for 40km, skirting the southern flank of the eastern massif and passing to the north and west of a third massif before turning southeast and flowing for a further 20km to its confluence with the Danube at Kelheim, at 48°55'N, 11°52'E and an altitude of 350m. This last 20km constitutes the ‘Unteren Altmühltal’ or Lower Altmühl Valley. The Danube is to the south of the Altmühl and flows due east past Neuburg and Ingolstadt before turning northeast towards Kelheim (see map at Figure 6.1). An important geographical feature of the Altmühl Valley is the Wellheimer Trockental, a shallow dry valley that runs some 17km from the Altmühl at Dollnstein in the north to the Danube at Rennertshofen, some 7km west of Neuburg, in the south. The Schutter, a small tributary of the Danube, rises

91

92

During the Middle Pleistocene, the Danube in the region of what is now the Altmühl underwent a series of radical changes in its course (Meyer and Schmitt-Kaler 1991). In the early part of the Riss glaciation (i.e. OIS 8) it flowed north, rather than east, at Rennertshofen and followed the course of the modern-day Altmühl from there to Kelheim, from where its course was the same as that today. The term ‘Altmühl-Donau’ denotes the river at this point in its development. At the same time, that part of the modern Altmühl northwest of Dollnstein was connected to the Main and is referred to as the ‘Urmain’; the Schutter existed as the ‘Urschutter’, but did not flow into the Altmühl-Donau. In the middle part of the Riss, the Danube broke through the rock barrier at Wellheim, abandoning the course of the present-day Altmühl and instead following the course of the Urschutter, creating the ‘Schutter-Donau’; at the same time the connection between the Urmain and the Main was broken. The Altmühl therefore attained its present form, but the Danube flowed north of Neuburg, past Wellheim, and rejoined its present course only at Ingolstadt. In consequence, that portion of its former course between Wellheim and Dollnstein was no longer a watercourse but a dry valley, the northern part of the modern Wellheimer Trockental. Finally, in the late Riss, the Danube broke through at Rennertshofen and attained its present course as the ‘Neuburger-Donau’, creating the whole of the modern Wellheimer Trockental and leaving the Schutter as a minor tributary. This development, then, led to the progressive shortening of the course of the Danube as it

migrated southwards to its present position on the southern fringe of the Frankische Alb (see Figure 6.2). From the early last glacial, however, there remained one important difference from today; the Danube between Neuburg and Neustadt was a wide wetland fan. It is not clear when this ceased to be, but Meyer and SchmidtKaler (1991, 32) place the occupation of Mauern within the period of the wetland fan. Raw Materials A wide range of raw materials was available to Middle Palaeolithic knappers. Fine-grained Cretaceous quartzite (Kreidequartzit) was widely used, along with Jurassic crypto-crystalline rocks, including lydite, radiolarite and jasper, generally obtained from stream and river beds. A key resource was ‘Jurahornstein’ or Jurassic chert, which occurs commonly in southern Bavaria either as residual nodules in Tertiary soils (especially to the east of Kelheim; Weiβmüller 1996) or, in the Frankische Alb, in laminated or tabular form. The latter occurs widely in outcrops such as that at Baiersdorf, 4km to the northwest of the Sesselfelsgrotte, and is variously referred to by the German authorities as Plattenhornstein or Plattensilex. All of these raw materials occur widely throughout the Altmühl Valley and precise characterisation is only rarely possible. In fact, there is little or no evidence for long distance transport of any raw material in the Middle Palaeolithic of the region.

Figure 6.2. Evolution of the Danube from the ‘Altmühl-Donau’ through the ‘Schutter-Donau to its present ‘Neuburger’ course. After Meyer and Schmitt-Kaler 1991, Abb. 16.

93

discrete occupations. The comparison with Ranis 2 is doubtful since the Jerzmanowice point might easily have derived from the Early Upper Palaeolithic horizon layer 2 Lower. A stratigraphically higher Middle Palaeolithic industry occurs in layer 3 Upper (Oben) which Freund calls Charentian, although a single leaf point was recovered from this level. All six leaf points from the site are Altmühlian in that they are closely similar to those from Mauern. Given the problems of post-depositional movement of artefacts, the cave’s size and easterly aspect, the paucity of cores and debitage, the absence of hearths and the small number of artefacts recovered, all that can be said is that the cave was visited occasionally by Middle Palaeolithic knappers and was not a location for extended occupation by social or family groups. Despite the stratigraphical difficulties and the absence of any fauna chronologically diagnostic at a level of precision finer than generally Würmian, the admixture of Middle and Upper Palaeolithic artefacts suggests that the Middle Palaeolithic leaf points are late in date. Another small Palaeolithic cave site, Jagerloch, is situated about a kilometre further north in the Ziegeltal, but is exclusively Upper Palaeolithic (Kaulich et al 1978).

Middle Palaeolithic Sites in the Altmühl Valley The location of Middle Palaeolithic sites in the Altmühl Valley is shown in Figure 6.1. The Weinberghöhlen (also referred to as Mauern) and Obernederhöhle have been discussed already in Chapter 4, along with the Bockstein localities and Kösten; the information given there will not be repeated here except insofar as is necessary. For present purposes the sites will be considered in three geographical clusters: the Lower Altmühl; the Wellheimer Trockental; and the western Altmühl Valley. The Lower Altmühl Valley There are 10 Middle Palaeolithic sites in the final 2025km of the Altmühl’s course, including the multiple localities at the Klausen Caves and Schulerloch. The river here is broad and might have been fordable only occasionally or seasonally in the Upper Pleistocene. The palaeolithic archaeology of the Lower Altmühl has been a particular interest of the University of Erlangen for 50 years, under the leadership first of Zotz and subsequently of Freund; an important summary appeared some 29 years ago (Kaulich et al 1978) and has been superseded only by the publication of Freund’s re-excavation of Obernederhöhle and by the publication of the excavations at the Sesselfelsgrotte.

Groβes Schulerloch (Figure 6.1, site 2) is a 100m-long cave situated on the northern bank of the Altmühl some 1.5 km upstream from the Ziegeltal (Bosinski 1967, 156). The entrance was excavated in 1914/15 by Birkner (1916), who, although he described a simple stratigraphy, did not refer any artefacts to any particular level. The finds must therefore be considered as a whole. The inventory consists of some 200 tools in addition to 32 cores, of which 10 are described as ‘prepared’, and a number of bone hammers or ‘compresseurs’. There are no leaf points. The assemblage is dominated by unifacial scrapers, some of which are pointed and a few of which could be termed convergent or even Mousterian points. There is a single bifacially worked Kartstein knife, a term used to describe bifacial limaces such as that from Couvin illustrated at Figures A2.3.i and A2.4.ii. Bosinski concluded that the industry is Mousterian, as opposed to Micoquian, but regards the presence of three bifacial scrapers, a Prondnik knife, a small handaxe, 13 Faustkeilblätter and two fragments of bifacially-worked pointed artefacts as evidence for a Micoquian presence. Whether this combination of unifacial flake tools with bifaces should be interpreted in these terms is, of course, questionable; the waters are muddied further by the impossibility of establishing any chronological sequence. There are no grounds for attributing the archaeology, or any part of it, to any particular period within the Upper Pleistocene. Groβes Schulerloch, then, is a site of limited value that cannot usefully be compared with others in the Altmühl Valley.

Obernederhöhle (Figure 6.1, site 1). This leaf point site is described in Chapter 4, where it was pointed out that Freund’s re-excavation (Freund 1987) was only partly successful in solving the stratigraphical problems arising from the work of the amateur Oberneder. It is located in the small Ziegeltal dry valley on the north (left) bank of the Altmühl, some 5km upstream of the river’s confluence with the Danube. Freund describes an industrial succession which commences with a Micoquian with leaf points and bifaces in layers 4, 3/4 and 3 Lower (Unten); the industry is compared with that from Klausennische. In fact, as is discussed below, the presence of leaf points at that site is subject to serious doubt, whereas the three leaf points recovered from these layers at the Obernederhöhle (see Appendix 2 Figure A2.35.i, ii, iii) are quite unequivocal and very different from the two coarse handaxes (e.g. Appendix 2 Figure A2.36.i), one flat handaxe (Faustkeilblatt) and one Keilmesser (Figure A2.36.ii) found in the same layers. In the overlying layer 3 General (Allgemein) there are a further two unequivocal leaf points (e.g. Figure A2.35.iv), one handaxe and three Fäustel, as well as a single Jerzmanowice point; Freund compares the industry with that from ‘the elder leaf-point-industry of the Weinberghöhlen’ (Freund 1987, 213), by which she means Mauern Layers G'/F2/G (see below), and with Ranis 2 on account of the Jerzmanowice point. However, there is no good reason to attribute the industries to different ‘cultures’ nor, as Freund points out, to identify

Kleine Schulerloch (Figure 6.1, site 3) is a small rock shelter close to the main Schulerloch cave; a small inventory has been recovered but is awaiting publication (Böhner, in prep). No more can be said at this stage.

94

Sesselfelsgrotte (Figure 6.1, site 4). The Sesselfelsgrotte is the key site within the Lower Altmühl Valley, excavations having been carried out between 1964 and 1977, with a further season in 1981, under the direction firstly of Zotz and subsequently of Freund (Freund 1998; Weiβmüller 1995a; Richter 1997).

Glacial and Late Upper Palaeolithic archaeology; they are underlain by level D, an archaeologically sterile loessic unit deposited during the last glacial maximum OIS 2. Level E consists of limestone rubble; a sedimentary hiatus occurs above the lowest unit, E3, which has yielded a small, poorly defined Mousterian with denticulates (Böhner, in prep). The ‘G-Komplex’ comprises levels F, G, H and I, of which G is much the most important archaeologically. Despite a thickness of only 25-35 cm, level G is divisible into six discrete sedimentary units (1, 2, 3, 4, 4a and 5) and has produced a stone artefact inventory of 85,000 pieces, including some 3900 retouched pieces attributed to the Micoquian on account of the high frequency of bifaces and bifacial working, and 350 cores. The ungulate faunal assemblage, dominated by reindeer, horse and mammoth, is largely burnt, and there are numerous hearths. Cave bear occurs only in G4. These features point to a wide range of activities having been carried out at the site during the deposition of level G, and to the frequent and extended occupation of the site as a centre of social activity. There is one possible leaf point fragment; otherwise, leaf points are entirely absent (Richter 1997).

The site consists of a rock shelter in the Sesselfels limestone cliff on the north bank of the Altmühl, upstream 2km from Schulerloch and 10 km from the confluence of the Altmühl and the Danube. At the modern ground level the rock shelter extends some 8m beneath the rock overhang and is 374m above sea level. 15 metres to the east there is a second, much larger, rock shelter, Abri im Dorf, which has produced only Gravettian artefacts (Figure 6.3). Excavation revealed sediments 7m deep, with 20 geologically discrete levels, from which a series of Middle and Upper Palaeolithic, as well as more recent, industries were recovered (Figure 6.4). Levels A, B and C together account for the topmost metre of the sequence and have produced Mediaeval, Late

Figure 6.3. Plan of and cross-section through the Sesselfelsgrotte. G = main excavation; S = test pit; a = line of cross-section shown in inset. The larger rock shelter to the south and east of Sesselfelsgrotte is Abri im Dorf. After Richter 1997, Abb 3.

95

Figure 6.4. Schematic composite stratigraphy of the Sesselfelsgrotte. After Richter 1997, Abb 4.

96

Beneath the G-Komplex, archaeologically sterile layers K and L overlie the ‘Unteren Schichten’ or Lower Levels, which amount to a thickness of 3m. The topmost level within this complex is M1, and it continues through levels M2, M3, N, O, P and Q; below this, the sequence is divided by a limestone ridge into a southern portion, in which level Q is underlain by levels R and S, and a western portion, in which the sequence continues through R-West, 1-West, 2-West and 3-West. 9400 Middle Palaeolithic artefacts were recovered in the course of excavation of these levels, on the basis of which eight inventories have been recognised and grouped into three industrial-stratigraphical ‘complexes’: a basal ‘microlithic’ Mousterian comparable to the Taubachian of Kůlna 11; a Charentian Mousterian with Quina-type elements; and a final Typical Mousterian. Bifaces are virtually absent throughout. There are no apparent archaeological hiatuses within the Lower Levels and, although the evidence for intensity of occupation is rather less than in the G-Komplex, the occurrence of cave bear only in M2 and of wolf only in O implies that humans occupied the cave frequently throughout the period in which the Lower Levels accumulated (Weiβmüller 1995a).

artefacts in L and K, implying the end of the formerly continuous human presence at the Sesselfelsgrotte, the presence only of Pinus, Larix and Picea charcoal in lower L and upper K and the complete absence of tree macrofossils at the L/K boundary region, this strongly suggests that levels K and L correspond to the OIS 4 glacial maximum. This would place the G-Komplex in the earlier part of OIS 3. The alternative, if one accepts Reisch’s view, is that levels K and L correspond to the stadial OIS 5d, that the G-Komplex accumulated in the OIS 5c Amersfoort-Brørup interstadial, and that the sedimentary hiatus in level E spans OIS 5b, 5a, 4 and the first half of OIS 3, since the loessic level D is certainly attributable to the last glacial maximum OIS 2 (Weiβmüller 1995a, 78). In fact the G-Komplex is much better placed in OIS 3, rather than OIS 5c, in view of (uncalibrated) finite radiocarbon dates of 35200±260 BP, 34800±300 BP, 36,600±875 BP (all on bone) and 41840+1170-1020 BP (on charcoal) for unit G2 and 46600+980-880 BP (on bone) for unit G4. Level N in the Lower Levels has an infinite date of >45,900 (Richter 1997, 25). Richter accepts Weiβmüller’s chronology, a major implication of which is that Bosinski’s (1967) placement of the Micoquian of Germany in the early last glacial is mistaken; in any case, the industrial succession at the Sesselfelsgrotte deals a severe blow to Bosinski’s claim that the Micoquian in Germany is older than the Mousterian. Although there are problems with Weiβmüller’s chronology - there is nothing in the Lower Levels that could be referred to the OIS 5d or 5b stadials, a fact that Weiβmüller attributes to the virtual cessation of sediment accumulation in these periods – the balance of evidence is strongly in favour of the Mousterian of the Lower Levels dating to the OIS 5 interstadials and the GKomplex Micoquian dating to early OIS 3.

Reisch has argued, on the basis of the presence in 2-West, S, O, N and M3 of the putatively warmth-demanding aquatic molluscs Bithynia tentaculata, Lithoglyphus naticoides and Fagotia acicularis that the whole of the Lower Levels accumulated during the last interglacial ‘oder vielleicht besser des letzten warmzeitlichen Komplexes’ (Reisch 1985, 53), a view endorsed by Freund (1984). This position has been challenged powerfully by Weiβmüller (1995a), who points out that Reisch’s analysis did not take into account the fact that, in view of the development of the deep sea oxygen isotope curve and its correlation with European terrestrial palaeoclimatic and palaeoenvironmental records, the last interglacial is now known to correspond only to OIS 5e rather than the whole of stage 5. He also notes that aquatic molluscs are rather poorer indicators of climate than land molluscs and that both B. tentaculata and L. naticoides are known from cool contexts in the early and middle last glacial (Weiβmüller 1995a, 78). He argues instead that the Lower Levels were laid down in the OIS 5c Amersfoort-Brørup and OIS 5a Odderade interstadials, with some possibility of the very lowest levels, 2-West and 3-West, being derived from the late Eemian.

The dating and climatic evidence of the G-Komplex is a matter of some interest. Richter takes the view that, in the light of the thickness of only 25-35cm of level G and the absence of evidence for climatic change within the level, that the G-Komplex represents a single interstadial. Noting that the G2 radiocarbon dates on bone are systematically younger than that on charcoal by some 7000 years and probably suffer from contamination, he argues that the bone date of 46,600 BP for G4 represents a true age of 53-54 kya, that the apparent time span of 10,000 years for level G is too great and that the whole of the G-Komplex falls within the Oerel-Glinde interstadial. This must be challenged. Firstly, Oerel and Glinde, as defined in the Oerel 61 core (Behre and Lade 1986), are two separate palynological interstadial events, not a single event. Secondly, these treeless ‘interstadials’ are defined in northern Germany; as was demonstrated in Chapter 5, the vegetation of southern Germany in OIS 3 was very different. A better referent would have been Samerberg (Grüger 1979, 1989), which is also in Bavaria and at an altitude of 600m, similar to that of much of the Frankische Alb. There an early OIS 3 forested event in which AP% exceeds 90% and in which Picea and Pinus dominate is apparent in pollen zone 29, followed by a

There is much evidence to support Weiβmüller’s view. The first appearance of cold-climate indicator species, namely charcoal of Pinus cembra and the remains of mammoth, is to be found in level M3, the uppermost in the Lower Levels. The overlying levels L and K are rich in remains of cold steppic mammalian microfauna, especially the rodent Dicrostonyx torquatus, which suggest a development from a cool climate in the lower part of level L through a very cold climate at the L/K boundary to more moderate conditions in the upper part of K and in I. Taken together with the absence of

97

somewhat less forested event in zone 31. Another forested episode earlier than either might be concealed in the hiatus between zones 28 and 29 (see Chapter 5). Even in the ‘stadial’ episode of zones 27 and 28, AP% reaches 70%, and in those of zones 30 and 32 reaches 40%. All of this is entirely consistent with the presence of Pinus, Picea, Larix and Corylus charcoal, with some Prunus and Tilia, in level G. Similarly, in view of the discussions in Chapter 3 on the biodiversity of the mammoth steppe and in Chapter 5 on the nature of vegetation history within OIS 3, the presence of reindeer throughout the GKomplex, along with mammoth in every level except H, in no way militates against the existence of a series of climatic oscillations within the G-Komplex. The fact that more specifically arctic elements such as wolverine, arctic hare and arctic fox are restricted to the upper units within the G-Komplex (Richter 1997, Tab. 2.2) does not demonstrate that the climatic decline of a single interstadial is represented; it is equally consistent with a net climatic decline through a series of climate oscillations, as is suggested by the decline in AP% in Samerberg pollen zone 31 relative to zone 29. To reiterate: environment history in OIS 3 must be understood in terms of particular wavelengths of climate change inducing different responses in different regions and localities. To think in terms of an alternation between synchronous, pan-European stadials and interstadials as defined in the pollen record of northwestern Europe is misguided.

sediments he found three undiagnostic, heavily patinated artefacts that he regarded as Middle Palaeolithic (Bosinski 1967, 160). None are leaf points. The Klausen Caves (Figure 6.1, sites 6, 7 8 and 9). These sites are of historical importance in this investigation in that the first definition of the Blattspitze, that given by Obermaier and Wernert (1929, 308) and cited in Chapter 4, was developed in the light of excavation at these sites. Obere, Mittlere and Untere Klause are the upper, middle and lower chambers of a single hydrological system. Klausennische is a rock shelter located between the Mittlere and Untere Klause. The system is on the south bank of the Altmühl, less than a kilometre upstream from Heidenstein and directly facing the Sesselfelsgrotte; the two locations are clearly visible each from the other. Because of the hydrological linkages between the caves the stratigraphic integrity of each is suspect. In any case, the stratigraphies are very complex (Freund 1963). The Obere Klause is 25m long and 15 m broad and consists of a series of niches or alcoves which militate against a uniform stratigraphy (Bosinski 1967, 158). Nevertheless Obermaier and Fraunholz (1926) constructed a sequence in which two Upper Palaeolithic (‘Magdalenian’) horizons were underlain by a ‘Solutréen’ (Obere Klause III), a Mousterian (Obere Klause II) and finally a level with a single handaxe (Obere Klause I). The OK I handaxe is triangular and flat; no more can be said. The industry from OK II consists of small collections from four niches and from the centre of the chamber; three scrapers in niche A, 14 scrapers in niche B, four scrapers in niche C; 21 tools, including scrapers, a limace and a Kartstein knife plus nine cores in niche E; and 22 scrapers, one denticulate and six cores in the chamber centre. The industry is described as Mousterian by Bosinski on account of the absence of clearly Micoquian bifacial type fossils, although three scrapers in niche B and two tools in niche E bear bifacial edge retouch. The ‘Solutréen’ of OK III is classified as ‘Altmühlian’ by Bosinski, who describes two whole and one fragmentary leaf points, in addition to some 25 tools, mainly scrapers, and two cores. In fact, only the fragment is a convincing leaf point (Appendix 2 Figure A2, 34.i), although the larger of the two claimed whole leaf points is actually incomplete, lacking its tip and a portion of one edge, and bears an asymmetrical basal notch very similar to that on the Mauern leaf point illustrated at Figure A2.26; it is a matter of debate as to whether the piece qualifies as a leaf point in terms of the definition advanced in Chapter 4. These three pieces are the only bifaces in the OK III inventory. It must be concluded that, given the problems of the stratigraphy, it is not advisable to regard either the OK II or OK III industries as distinct from each other or as internally unified. Nevertheless, the cave does furnish evidence in OK III for a small number of leaf points that are typologically and technologically distinctive when compared with the pieces with which they are stratigraphically associated.

Thirdly, the thinness of level G cannot be recruited as evidence for a single interstadial. The thickness of the GKomplex as a whole is 60-100cm; given that the Lower Levels span the period between 118 kya and 74 kya in a thickness of 3m – possibly even longer if OIS 5e is present in the western profile – the thickness of the GKomplex is perfectly reconcilable with a time span of 10,000 years. Finally, Richter recognises in the lithic inventory four discrete episodes of occupation, each commencing with a small ‘pioneer’ occupation and culminating in a larger inventory with the targeting of specific raw materials, as is discussed further below. If the G-Komplex does indeed span 10,000 years, as the radiocarbon dates suggest, this would imply cycles of colonisation and abandonment of the site lasting around 2,500 years – the same order of duration as the Dansgaard-Oeschger warming events apparent in the GRIP ice core oxygen isotope record. In summary, the evidence from the G-Komplex at the Sesselfelsgrotte implies a series of occupations over a period of several thousand years in the context of the cyclical expansion and contraction of the coniferous forest elements in a persistent cool-mosaic landscape. Heidenstein (Figure 6.1, site 5) is a small cave on the southern (right) bank of the Altmühl directly due south of the Sesselfelsgrotte. It was excavated by Birkner (1933) who established no stratigraphy and found a small Upper Palaeolithic industry. However, in the lower part of the

98

The situation at Mittlere Klause is very similar to that at Obere Klause. The chamber consists almost entirely of a series of niches, so Obermaier and Fraunholz’s (1927) stratigraphy must be treated with the same caution as that of the higher cave. Again, an Upper Palaeolithic (‘Magdalenian’) industry overlies a ‘Solutréen’ (Mittlere Klause III), a Mousterian (MK II) and an ‘Acheuléen’ (MK I); again, Bosinski recasts the succession as Altmühlian underlain by a Mousterian with a basal Micoquian, the last of which consists of a single triangular handaxe. The MK II Mousterian amounts to just five scrapers, one Breitklinge (broad blade), 14 notched or edge-damaged pieces and one core. The ‘Altmühlian’ inventory is even smaller and features, according to Bosinski (1967, 158) just a single leaf point fragment (Figure A2.34.ii) and a small flat handaxe in addition to a limace and two other flake tools. Skeletal remains of an adult male Homo sapiens were found in the ‘Solutréen’ horizon, but the burial appears to have been intrusive from the overlying Upper Palaeolithic (Freund 1963, 46).

the industries is entirely speculative and their integrity open to doubt. Obere and Mittlere Klause do suggest a conformity to the pattern identified in Chapter 4 of leaf points occurring in small assemblages indicative of nonresidential occupation, but the Klausennische assemblage is hardly large but still lacks leaf points. It is wise to treat these sites with caution. Fischleitenhöhle, Mühlbach (Figure 6.1, site 10). Excavated in 1918 by Birkner (1918, 1925) this small cave is located in the upper part of the Lower Altmühl, some 17km upstream from the Klausen caves and overlooking the river from its left bank. At the cave entrance Birkner found three superimposed hearths, beneath which was a small Mousterian industry comprising 18 unifacial tools including scrapers, points and a limace, plus one Kartstein knife and several notched pieces (Bosinski 1967, 167). The fauna consists of mammoth, horse, rhino, bison, reindeer, wolf, hyaena and cave bear. It is not clear whether the hearths relate to the Middle Palaeolithic finds since a few Upper Palaeolithic artefacts, including a split based bone point, were found in the upper part of the archaeological horizon. The industry lacks bifaces and cannot be described as Micoquian; its date is not known.

The lowest of the Klausen caves, Untere Klause, is even less informative than the higher two. During the 1860s it was converted into a beer cellar and most of the sediments removed in the process. Excavations in 1960 at the front of cave turned up what Freund (1961) termed a ‘transitional’ leaf point, but which Bosinski (1967, 157) calls a Faustkeilblatt; it is impossible to draw any meaningful conclusions from this single artefact.

The Wellheimer Region There are a further six Middle Palaeolithic sites to the west of the Lower Altmühl, most of which occur close to the Wellheimer Trockental, the dry valley created by the southward migration of the Danube described earlier. The valley itself is rather shallow-sided, although the land to either side rises to over 500m. The Danube-Altmühl interfluve for some 30km to the east, including the course of the Schutter, is more low lying and the relief more rolling.

Finally, the Klausennische has produced the largest inventory of any of the Klausen sites. Obermaier and Wernert (1914, 1929) established a stratigraphy in which a Middle Palaeolithic layer C underlies layer B, which yielded a few Upper Palaeolithic pieces. The Middle Palaeolithic industry is Micoquian in character; there are 6 cores and some 280 retouched tools, among which bifacial pieces are numerically dominant, with 76 Faustkeilblätter, 9 biface knives, 23 Prondnik knives, 69 bifacial scrapers, a few Fäustel and several bifacially worked pieces of uncertain type. Bosinski (1967, 160) also identifies two leaf points, one whole and one fragmentary, although the former is too thick to qualify as a leaf point here and is better described as a small, elongate flat handaxe or as a biface scraper with flat surface retouch. The latter is too incomplete to judge whether or not it is a leaf point. This industry is the type for Bosinski’s ‘Klausennische’ facies of the Micoquian, characterised by a preponderance of Faustkeilblätter. There is no indication in the literature as to any features such as hearths or as to the frequency of waste flakes; all that can be said of the fauna is that mammoth, rhino and horse are present.

Gaimersheim (Figure 6.1, site 11) is not, strictly speaking, in the Wellheimer region but in the centre of the Danube-Altmühl interfluve some 10km northwest of Ingolstadt. Excavations in a brickearth pit in 1942-3 yielded two unifacial convergent scrapers in association with some possibly flaked nodules and with the remains of equids (Müller-Beck 1957). No stratigraphic profile was established. Biesenhard (Figure 6.1, site 12). A single leaf point-like artefact was found on the north bank of the Schutter at the western foot of the Schutterberg, the hill immediately to the east of Wellheim, in 1947 (Bosinski 1967, 154). Being found on the surface it lacks stratigraphic context entirely, although it was found immediately below the entrance to a cave out of which it might have fallen. The artefact is too thick and distally rounded to be regarded as an unequivocal leaf point (Figure A2. 28.ii), but the retouch technique of flat surface removals covering most of both faces, combined with fine edge retouch concentrated on one face, is very similar to that of the

In conclusion, the Klausen caves, despite their historic importance, are not well enough understood to form the basis for sound conclusions as to relationships between site landscape context and lithic-industrial practices in the Middle Palaeolithic of the Altmühl Valley. The dating of

99

As was pointed out in Chapter 4, Müller-Beck (1974) has placed Bohmers’ Layers G' and G (i.e. his Zone 5, Zotz’s Layers F2 and G) in a cold stage immediately prior to Hengelo, and Layer F (his Zone 4, Zotz’s Layer F1) in the Hengelo warm event. It must be pointed out that, for him, this meant a date late in the first Würmian pleniglacial maximum for upper Zone 5, and in the early part of the first interpleniglacial interstadial for Zone 4. He emphasised that Zone 4 displayed evidence of cold conditions at its base, and therefore seems to have accumulated in a period of improving climate. This is generally consistent with the evidence from the Sesselfelsgrotte where the G-Komplex Micoquian industries (which are very similar to both the Mauern ‘Altmühlian’ and ‘Mousterian/Mousteroid late Micoquian’ – see below) relate to early OIS 3. However, the idea that Zone 5 dates to the OIS 4 glacial maximum is doubtful in view of the complete absence of evidence at sites elsewhere in the region for occupation during that period, and in the light of the parallels between the ‘Mousterian/Mousteroid Late Micoquian’ and the GKomplex Micoquian. Early OIS 3 dates for layers F1/F, F2/G' and G seem very likely.

Mauern leaf points. Furthermore, the illustration hints at an asymmetrical basal notch like those on the Mauern leaf point illustrated at Figure A2.26, and on the possible leaf point from Obere Klause III. The piece is made on Plattenhornstein, as is shown by the presence of cortex on both faces, and the attributes it shares with the Mauern leaf points might simply indicate a technique employed to work that material into hafted, elongate tools. Ried (Figure 6.1, site 13). Excavations in 1952 and 1958 at a brickpit near Ried, 2km south of the village of Mauern, each produced a cordiform handaxe from an 8m loess and soil profile (Freund 1963). The artefacts were located in a brown palaeosol at depths of between 4.5m and 7m, overlain by 4m of loess deposits. Freund placed the finds in the Acheulean, while Bosinski (1967) classified them as ‘Jungacheuléen’ and therefore as dating to the penultimate (Riss) glaciation or the early last interglacial. In fact these attributions are made on typological grounds alone; the weakness of soil formation in the profile above the archaeological horizon does not suggest that the OIS 5 interstadials are represented, although the thickness of the loess is considerable. The handaxes are therefore undatable with any precision and might have originated in any warm event between OIS 6 and early OIS 3.

The correlation of the cave sequence with the pollen curves derived from the valley sediments is a matter of some debate; the direct tracing of layers from the valley to the cave is impossible. The question therefore turns on the dating of the pollen successions. That from the 1967 ‘Mauern I’ core (Brande 1975, Bild 2; illustrated here at Figure 6.6) displays an interstadial development of pinespruce-larch forest in pollen zone A2 and B, between 700 and 800cm core depth, in which AP% reaches 75% before declining to 49,970 BP and the paucity of thermophilous taxa in the zone B spectra, rejects the idea that it corresponds to an OIS 5 interstadial and places it instead in the Middle Pleniglacial. However, the infinite date does not rule out an age older than 74 kya for zone B, and thermophiles contribute only 4.27% of

100

Figure 6.5. Plan of the Weinberghöhlen cave system as mapped by Zotz (1955, Bild 3). There are some differences with Bohmers’ plan (1951, Taf. 1), some of which reflect differences in the excavations. Höhle 1 above is Bohmers’ Höhle A (Westhöhle) and Höhle 2 is the western part of Bohmers’ Höhle B (Mittelhöhle), while the latter’s eastern end is Höhle 3 above. Höhle 4 corresponds to Bohmers’ Neben Höhle. Two chambers not excavated by Zotz, and not shown in the plan, were investigated by Bohmers who named them Höhle C (Osthöhle) to the east of Höhle 4, and Höhle D (Untere-Höhle) to its south. On the other hand, Bohmers did not excavate in or name the Saazer Loch.

pollen in the OIS 5a forest spectra at Samerberg (Table 5.2). On balance, the most plausible chronology places the zone B forest in OIS 5a, the cold, wet and open landscapes implied by zone C1 in OIS 4, and the vegetation oscillations of C2 in the early part of OIS 3. Layers G'/F2/G and F/F1 in the cave stratigraphy are therefore best placed early in pollen zone C2, with Layers G'/F2/G possibly corresponding to the boundary between pollen zones C1 and C2, i.e. the OIS 4/3 boundary and early OIS 3.

sites with woodland expansion and contraction on subMilankovich wavelengths in OIS 3. Rohrbach (Figure 6.1, site 15). In 1936 a single Faustkeilblatt was found on the surface near Rohrbach, on the western slopes of the Wellheimer Trockental 5km southwest of the Weinberghöhlen (Bosinski 1967,171). Bosinski places the piece in the Micoquian, but its age is clearly unknowable. Steinerner Rosenkrantz, Mörnsheim (Figure 6.1, site 16). This small cave is near Mörnsheim in the valley of the river Gailach, a small tributary of the Altmühl that flows into the latter from the west. It is separated from the northern portion of the Wellheimer Trockental 5km to the east by a ridge that reaches an altitude of 530m. Excavations in 1924-5 produced a small industry resolved into Mousterian and ‘Solutrean’ components on typological grounds by Mayr-Lenoir (1927). However the sediments defy the construction of a clear stratigraphical sequence, and the artefacts are distributed evenly throughout the deposits (Bosinski 1967, 166). Bosinski details just four artefacts, including a thick convex sidescraper, a bifacial scraper, a biface described as a ‘Halbkeilartiges Werkzeug’ and an unequivocal leaf point (Appendix 2, Figure A2.28.i), in addition to a few blades which he does not describe or illustrate.

The oscillations in AP% in zones C2 and D of the Mauern I curve are not of high amplitude, certainly not sufficient to imply episodes of significant forest development in the Wellheimer region in OIS 3. Comparisons with the 1937 Mauern core (Schütrumpf 1951, Abb. 3) are instructive in this regard. Müller-Beck (1974) argues that a hiatus in this core between 660 and 700cm depth conceals the first Würmian glacial maximum; if this is so then the overlying 260cm, in which AP% oscillates generally between 15% and 50%, with three peaks in excess of 85%, correspond to at least the early part, and possibly the whole, of OIS 3. Oscillations of this amplitude are much more consistent with pulses of well-developed woodland expansion such as are recorded in the Samerberg core; this supports at the level of this specific site the continental-scale pattern revealed in Chapter 4, namely the association of Middle Palaeolithic leaf point

101

The Western Altmühl Valley Hohler Stein, Schambach (Figure 6.1, site 17) is a cave in the valley of the Hirte, a stream that flows into the Altmühl from the north 20km northwest of Dollnstein. The Altmühl in this region enters a steeply-cut valley between two upland plateaux; the site is also close to the source of the Rednitz, and is therefore linked to Kösten in the north by the watercourses described earlier. Gumpert (1952) described a stratigraphy in which Upper Palaeolithic archaeology was found in a loessic layer B beneath a topmost layer A with mediaeval and Neolithic artefacts. Beneath this, horizons with Middle Palaeolithic artefacts were found in layers C, F, G and H. Bosinski (1967, 154-5) places all of these in his ‘Schambach’ facies of the Micoquian, which he defines as rich in Halbkeile (elongate handaxes with a flat ventral surface), small flat handaxes, Prondnik knives and bifacial scrapers but poor in Micoquian handaxes and biface knives, and terms them ‘Schambach IV, III, II and I’ respectively. The inventories are not large; Schambach I amounts to 12 tools plus several notched pieces and three cores. Schambach II consists of 59 tools plus a number of notched pieces but no cores, while Schambach III comprises 52 tools plus notches, denticulates and nine cores. Finally Schambach IV amounts to no more than 28 tools with eight cores and several notches and denticulates; a hearth was found at this level in the cave interior. Throughout, the tools are predominantly bifacially worked and fall into the types on which the Schambach facies is defined; unifacial scrapers, sometimes pointed, are also important components, and Bosinski refers to occasional Levallois flakes and blades. There are no leaf points in the Schambach I and II inventories, but two leaf point-like pieces occur in Schambach III. One is described as a Blattspitzenwechselschaber (pointed foliate alternate scraper) by Bosinski (Appendix 2 Figure A2.37.ii) and cannot be regarded as a leaf point on account of its thickness; the second (Figure A2.37.i), the base of which is elongated in a manner that might be called a tang, is called a Faustkeilblatt by Bosinski but could be an unfinished leaf point. A single fragmentary piece from Schambach IV (Figure A2.37.iii) is a convincing leaf point.

Figure 6.6. Pollen diagram, 1967 core ‘Mauern 1’ After Brande 1975, Bild 2.

The sediments in which the Middle Palaeolithic artefacts occur do not suggest full glacial conditions although woolly rhinoceros, mammoth, reindeer, arctic fox, arctic hare and lemmings are important in the fauna. Cave bear and hyaena are also present. Dating is uncertain, but these factors are certainly consistent with an OIS 3 date and a general contemporaneity with the G-Komplex occupations at the Sesselfelsgrotte. There is no information in the cited publications as to the waste debitage, but the small number of cores present would suggest that it is not present in great quantities. On the other hand, the predominance of bifaces, often on tabular chert, would militate against a relationship between core frequency and waste flake frequency. Nevertheless, the

Figure 6.7. Pollen diagram, 1937 Mauern core. After Schütrumpf 1951, Abb. 3.

102

small size of the assemblages, the presence of cave bear, hyaena and arctic fox, and the paucity of hearths together imply that Middle Palaeolithic occupation of the cave was occasional and ephemeral. The virtual absence of leaf points in these Micoquian assemblages is therefore of interest in that it is in just such occupational contexts that leaf points occur in the Altmühl Valley.

Rosenkrantz and Biesenhard are as small or smaller, and are classified here as leaf point sites essentially on the grounds that they have produced whole, rather than fragmentary examples. It must be admitted that this might not be an adequate basis for the distinction. Schambach poses a problem in that it is a stratified site with small inventories and few or no leaf points on the basis of which Bosinski erected a facies of the Micoquian. It is possible that the reported assemblage sizes would have been larger if the same criteria for the identification of retouched tools as have been applied at the Sesselfelsgrotte, where small, typologically unclassifiable retouched pieces are included, had been adopted.

Breitenfurter Höhle (Figure 6.1, site 18) is located on the north bank of the Altmühl 1km to the west of Höhler Stein. The industry from the cave (Gumpert 1956) bears very close comparison with those from that site. The Middle Palaeolithic occurs at the base of layer D, and amounts to eight tools, of which four are bifaces. One of these is a fragment of a probable leaf point. The fauna is also similar to that of the Höhler Stein.

The key sites are the Sesselfelsgrotte, Mauern and, to a lesser extent, Obernederhöhle. The first of these is a site of the highest importance and value, on the basis of which one might tentatively place the biface-poor industries of Groβes Schulerloch and Mühlbach in OIS 5. The same might be said of the Mousterian levels at Obere and Mittlere Klause, which underlie ‘Altmühlian’ levels, but the assemblages are so small, and their stratigraphical integrity so uncertain, that this might be unwise. The relationship between Middle Palaeolithic collections with and without leaf points in OIS 3 will therefore be explored further by a comparison of the archaeology of the Sesselfelsgrotte G-Komplex with that of Mauern Levels G and F.

Groβe Ofnet, Kleine Ofnet (Figure 6.1, sites 19, 20). The cave sites of Groβe and Kleine Ofnet are situated on the slopes of the Ries crater south of Nordlingen, some 40 km to the west of the Wellheimer Trockental (Bosinski 1967, 162-3). The assemblage from Groβe Ofnet, excavated in 1875 and 1907, is lost; Bosinski reports that an artefact that might possibly have been a leaf point was found at the site, but, even if this were true, the cave stratigraphy is completely unknown. The situation at Kleine Ofnet is a little better. A profile was described by Schmidt (1912; Bosinski 1967) in which a single Middle Palaeolithic piece occurred in layer C; Frickhinger’s subsequent investigation (Frickhinger 1937) referred a few more Middle Palaeolithic pieces to layer D in one section and layer F in a second. In fact the stratigraphic provenance of almost all the finds, many of which are lost, is uncertain; according to Bosinski (1967, 162) only three tools, plus one known to him only from a photograph, can be referred to any particular horizon. Of the pieces of unknown provenance, which number only 11, three might be regarded as leaf points (Figure A2.29); the remainder of the tools are scrapers and bifaces. Allsworth-Jones (1986) refers to 7 bifacial and 3 unifacial leaf points at Kleine Ofnet, and 2 of each from Groβe Ofnet, but the basis of this claim is unclear.

The Sesselfelsgrotte and the Weinberghöhlen It is intended here to address the question of whether the Mauern Layer F and G industries, and at least some of the industries from the G-Komplex at the Sesselfelsgrotte, might properly be regarded as traces of a single mode of engagement with and use of the Altmühl Valley landscape, rather than as culturally and temporally distinct entities. The high probability that these industries are broadly contemporaneous, in that all relate to the earlier part of OIS 3, has been established above. Attention will now be paid to the lithic industries, beginning with an examination of Richter’s analysis of the Sesselfelsgrotte G-Komplex.

Summary It is clear that most of the Middle Palaeolithic sites of the Altmühl Valley lack sufficient precision, with regard to their age and the stratigraphic provenance and integrity of their industries, to form the basis of even a moderately robust interpretation of landscape use during the Middle Palaeolithic of OIS 3. The relation between leaf points and small assemblages generated by short occupations seems, however, to hold, as does the OIS 3 date of leaf point assemblages. It has to be said that Breitenfurt, Mittlere Klause and Obere Klause have produced clearly Micoquian industries in just such occupational contexts in which leaf points are present only in exceedingly small numbers; they are not classified as leaf point sites here. The inventories from the Ofnet caves, Steinerner

The Sesselfelsgrotte G-Komplex The layers of the G-Komplex at the Sesselfelsgrotte, i.e. layers G, H and I, together amount to a thickness of 60100cm. Richter (1997) reports a total of 84,493 artefacts from these horizons, of which he classifies 3859 as retouched tools; there are also 353 cores. The thinness of the G-Komplex militates against clear stratigraphic sequencing of the finds within it, as does the absence of some of the layers, and of some of the sub-units of layer G, from the peripheries of the excavated area. However, Richter defined archaeological units through the

103

horizontal and vertical mapping of artefact scatters and by the identification of artefact clusters marked by similar raw material compositions (Richter 1997, Chapter 4). His method explicitly assumes that each occupation event at the site might be recognised by its specific spectrum of raw materials. It also demands that the units be spatially consistent with features such as hearths. No refitting was carried out, neither was any consideration given to the possibility of post-depositional differential vertical movement of artefacts, or of Middle Palaeolithic occupants of the site ‘scavenging’ discarded lithics from previous occupations and thereby blurring the boundaries between occupation events. Richter’s archaeological units must therefore be regarded as provisional. Nevertheless, a pattern emerges that deserves serious consideration, and his analysis will be accepted as the basis of the following comparison.

if one concentrates analysis on the unifacial flake tools, and that the Micoquian element - i.e. bifaces – varies in importance with the time depth of social occupation of the site. He therefore proposes the term ‘Mousterian with Micoquian Option’ (M.M.O.) to describe the G-Komplex industries. He also describes, on the basis of comparisons between the G-Komplex and Micoquian industries from other sites in Europe, chronological sub-divisions within it. The oldest is the M.M.O.-A, which originates at the end of the first Würmian pleniglacial maximum; it is characterised by Quina reduction techniques (M.M.O.A1) and other non-Levallois techniques (M.M.O.-A2). Larger assemblages contain Bockstein-type bifaces, including Micoquian handaxes and bifacial backed knives. Somewhat younger, but still referable to the early part of OIS 3, is the M.M.O.-B, in which Levallois reduction is much more important. In the M.M.O.-B1 variant centripetal-recurrent Levallois reduction dominates, whereas in the M.M.O.-B2 variant parallelrecurrent techniques dominate. In the Sesselfelsgrotte GKomplex, inventories A11-A08 (the main part of cycle 1, plus cycle 2) correspond to the M.M.O.-A1; inventories A07-A02 (cycle 3 and all except the youngest inventory within cycle 4) are referred to the M.M.O.-B1, while inventory A01 is described as M.M.O.-B2. Richter concludes that the M.M.O.-A2 variant is absent from the Sesselfelsgrotte archaeological succession, but cites Kůlna Layer 7a as an example (Richter 1997, Tab. 9.10).

Richter’s method identified 16 discrete units or inventories (Inventäre) of which 13 are stratified and three are not. The stratified inventories, from lower to higher, are: A13 and A12 (layers I and H); A11 and A10 (sub-layers G5 and G4a); A09 and A08 (sub-layer G4); A07 (sub-layer G3); A06, A05 and A04 (sub-layer G2); A03 and A02 (sub-layers G2 and G1); and A01 (sublayer G1). These are grouped into four cycles. The oldest, cycle 1, comprises inventories A13-A10; cycle 2 incorporates inventories A09 and A08, and cycle 3 A07A04; finally, the youngest, cycle 4, is made up of inventories A03-A01. Each cycle is interpreted as the trace of a period of Middle Palaeolithic settlement of the region during which the Sesselfelsgrotte was occupied several times. Richter shows that, within each cycle, the earliest inventories tend to be small with relatively few bifaces, frequent denticulates, limited use of Levallois reduction and a wide range of raw materials; these inventories (A13 and A12 in cycle 1, A09 in cycle 2, A07 in cycle 3 and A03 in cycle 4) he terms Initialinventäre. Subsequent inventories within each cycle (Konsekutivinventäre) are generally larger with more bifaces, proportionally fewer denticulates, more Levallois reduction and a concentration upon a few raw materials. This he interprets as evidence for pioneer occupations, in which knowledge of the landscape and the location of raw material sources was poor, followed by established occupations characterised by concentration upon a few favoured raw material sources and the use of the Sesselfelsgrotte as an important point of occupation. Abandonment of the region follows. Each cycle in the GKomplex industries corresponds, according to Richter, to one complete occupation succession from pioneer through established to abandonment (Richter 1997, Chapter 8).

The Weinberghöhlen, Mauern Richter’s analysis of the G-Komplex industries, although technically innovative and willing to challenge the German consensus on the Middle Palaeolithic succession, remains firmly rooted in a site-based method in which industries from sites across Europe are compared and temporal industrial successions constructed. The landscape, both as a spatial analytical unit and as the realm in which past human action took place, is entirely absent from his thinking. Nevertheless, he does address the question of Mauern layers G and F (Zones 5 and 4). Although he does not attempt even a rudimentary analysis of the Mauern industries, he places the Layer G/Zone 5 industry in the M.M.O.-B2 variant, and therefore equates it with the G-Komplex inventory A01; the industry from Mauern Layer F/Zone 4 he ascribes to a ‘presumably-later’ variant, the M.M.O.-B3, characterised by a reliance on parallel-recurrent Levallois reduction techniques and the production of relatively few bifaces. Richter’s position therefore generates testable implications. Within the Mauern sequence, the Layer F/Zone 4 industry should feature fewer bifaces than the Layer g/Zone 5 industry; also, Levallois reduction in Layer F/Zone 4 should be exclusively parallel-recurrent, whereas some centripetal-recurrent working ought to be present in the Layer G/Zone 5 industry. Furthermore, the Layer F/Zone 4 industry should possess fewer bifaces and a greater use of parallel-recurrent Levallois reduction

Furthermore, Richter challenges the notion of the Micoquian as a space-time cultural entity distinct from the Mousterian. He observes that most of the inventories, when analysed in terms of bifacial type fossils, could be placed in more than one of Bosinski’s facies, that their classification as Bordian Mousterian variants is possible

104

than any inventory within the Sesselfelsgrotte GKomplex. These implications are tested here.

actually made on Levallois blanks is likely to be higher than the figures recorded here since, in many cases, the application of flat surface retouch obliterates all surface traces of the blank. Certainly, one unifacial leaf point in Mauern Layer F/Zone 4 is clearly made on a convergent Levallois blank (Bohmers 1951, Taf. 27.3); this may also have been true of others.

Comparison of the industries The following comparison uses data collected by the author in the course of examinations of the curated collections held at the Prähistorische Staatsamlung, Munich (Weinberghöhlen, Mauern) and the Institut für Ur- und Frühgeschichte, University of Erlangen (Sesselfelsgrotte) in May of 1997. The following data were collected:

5. The number of pieces in each industry/inventory made on tablets or plaquettes of was recorded. Attribution to this category depended upon the survival of cortex on both faces of the piece. Pieces made on raw material identified as tabular by Richter, but which were clearly made on flakes, were excluded. As with 4. above, the high incidence of flat surface retouch on the whole of both faces is likely to render the number of pieces placed in this category an underestimate of the true figure.

1. The total number of retouched tools present in the curated collections was determined for both the Mauern Layer F/Zone 4 and Layer G/Zone 5 industries, and for each inventory within the Sesselfelsgrotte G-Komplex. The counts do not always conform to those published. For the Sesselfelsgrotte, analysis was restricted to those pieces placed by Richter in his category ‘Ge1’, i.e. typologically classifiable retouched tools. Pieces categorised as ‘Ge2’ and ‘Ge3’ – irregular and unclassifiable retouched pieces larger and smaller than 2cm respectively – are omitted from analysis here, as were a few smaller and less characterisable tools in the Ge1 category. In the case of Mauern, tool counts observed in the course of examination sometimes differed from those published.

The products of Bohmers’ and Zotz’s excavations at Mauern were examined separately, but then amalgamated on the basis of Bohmers’ Layers F, G' and G being the stratigraphic equivalents of Zotz’s F1, F2 and G respectively. Bohmers (1951) drew no archaeological distinction between his Layers G' and G, placing all materials recovered from them in his ‘Mousterian’ category without specifying from which of the two layers any particular pieces derived. Zotz, on the other hand, did draw such a distinction, although he treated the inventories from his Layers F2 and G as culturally identical. In fact Zotz’s stratigraphic attributions are somewhat problematic. Layers F1, F2 and G do not occur together in any profile he describes; F2 is missing from ‘Block I’, while he considers that the uppermost Middle Palaeolithic horizon in the inner part of ‘Höhle 1’ probably corresponds to F1 but might possibly be better placed in F2 (Zotz 1955, 97). The stratigraphic sequence in ‘Profil 2’ from ‘Höhle 2’ is numbered differently, Layers 9, 10 and 11 being equated with F1, F2 and G in the main sequence respectively. Zotz’s correlations are accepted here. He also describes artefacts recovered from the boundary between Layers G and the underlying H, and within H itself; all are excluded from this analysis. Finally, Zotz defined a series of other profiles in the caves that he could not correlate with the main stratigraphy; artefacts recovered from those parts of his excavation are not included in this analysis. The data for Mauern are calculated both for the entire tool assemblage in each layer/zone, and also with the leaf points discounted. For the Sesselfelsgrotte, Richter’s total retouched piece counts (i.e. the sum of published counts for his categories Ge1, Ge2 and Ge3) and his total artefact count (including all retouched pieces, cores and waste flakes) are given for each of the 13 stratified inventories (Richter 1997, Chapter 5 and Tab. 7.5). For Bohmers’ Layers F and G'/G published total artefact counts are cited as reported (Bohmers 1951, Tabs. 2 and 4), although this was not possible for Zotz’s Layers F1 and F2/G since he provides no clear tabulated or listed

2. The incidence of bifacial working was recorded, defined so as to include all examples of bifacial removals intersecting at the same region of an edge. Alternating retouch, where regions of the tool edge are retouched either dorsally or ventrally but not both, are excluded, as is edge damage and trivial edge nibbling. However, ventral thinning, where it intersects with dorsal edge retouch, is included. 3. The number of bifaces in each industry/inventory was recorded. For the purposes of this analysis, pieces on which at least one edge was formed by bifacial retouch, or whose overall form was a product of bifacial working, were categorised as bifaces. This means that, in addition to ‘true’ bifaces such as handaxes, Keilmesser and Faustkeilblätter, scrapers with bifacial edge retouch were also included in this category. It should be noted that, in view of 2. above, the ‘unifacial’ leaf points in the Mauern Layer F/Zone G industry are therefore classified here as bifacially worked but not as bifaces. 4. Pieces made on Levallois blanks were identified and placed in one or other of two categories: centripetal or parallel/convergent. This was achieved by a determination of the directions from which surviving dorsal flake scars were struck. No attempt was made to distinguish between ‘classic’ and recurrent, or between unipolar and bipolar modes of core reduction in either category. It is important to note that the number of pieces

105

Cycle/ Inventory

Bifacial Working

Bifaces

Centripetal Levallois

Parallel and Convergent Levallois

Plaquettes

N

Total Tools

Total Artefacts

CYCLE 1 A13 % A12 % A11 % A10 % SUM %

0 2 8.33% 11 26.19% 30 42.86% 43 31.62%

0 2 8.33% 4 9.52% 21 30.00% 27 19.85%

0 3 12.50% 0 2 2.86% 5 3.68%

0 2 8.33% 3 7.14% 4 5.71% 10 7.35%

0 0 7 16.67% 6 8.57% 14 10.29%

0

0

70

24

45

1584

42

108

3660

70

163

3963

136

316

9277

CYCLE 2 A09 % A08 % SUM %

67 41.61% 90 42.45% 157 42.09%

47 29.19 71 33.49% 118 31.64%

7 4.35% 10 4.72% 17 4.56%

0 2 0.94% 2 0.54%

17 10.56 28 13.21% 45 12.06%

161

326

5861

212

424

8039

373

750

13900

CYCLE 3 A07 % A06 % A05 % A04 % SUM %

22 36.07% 81 28.22% 43 29.45% 34 37.36% 180 30.77%

12 19.67% 58 20.21% 32 14.38% 21 23.08% 123 21.03%

0 13 4.53% 9 6.16% 3 3.30% 25 4.27%

0 23 8.01% 16 10.96% 7 7.69% 46 7.86%

3 4.92% 23 8.01% 7 4.79% 9 9.89% 42 7.18%

61

147

3828

287

741

12777

146

279

7581

91

203

3147

585

1370

27333

CYCLE 4 A03 % A02 % A01 % SUM %

28 35.00% 64 36.57% 91 30.54% 183 33.09%

18 22.50% 40 22.86% 58 19.46% 116 20.98%

2 2.50% 9 5.14% 8 2.68% 19 3.44%

3 3.75% 29 16.57% 14 4.70% 46 8.32%

16 20.00% 28 16.00% 30 10.07% 74 13.38%

80

169

4181

175

412

6102

298

538

17416

553

1119

27699

563

384

66

104

175

1647

3555

78209

34.18%

23.32%

4.01%

6.31%

10.63%

GRAND SUM %

Table 6.1: Analysis of retouched tools in the Sesselfelsgrotte G-Komplex. The 13 stratified inventories are grouped into cycles, and the data are presented per inventory and summated for each cycle; data for all cycles is also summated at the bottom of the table. N=number of pieces examined; cited percentages are calculated with reference to this figure. ‘Total tools’ gives the sum of tools in Richter’s categories Ge1 (classifiable retouched tools, from which all examined pieces were drawn), Ge2 (unclassifiable retouched pieces larger than 2cm) and Ge3 (unclassifiable retouched pieces smaller than 2cm). ‘Total artefacts’ gives all chipped stone pieces, including waste flakes and cores. Note that there are also three unstratified assemblages, X01, X02 and X03, which together contain 181 ‘classifiable’ and 123 ‘unclassifiable’ retouched pieces out of a total of 6284 artefacts. For the G-Komplex as a whole, then, the total number of retouched pieces is 3859 out of a total of 84493 artefacts (Richter 1997).

106

Cycle/ Inventory

LAYER F/F1 Bohmers F % minus leaf points % Zotz F1 % minus leaf points % SUM F/F1 % minus leaf points % LAYER G'/F2/G Bohmers G'/G % Zotz F2/G % minus leaf points % SUM G'/F2/G % minus leaf points % GRAND SUM % minus leaf points %

Bifacial Working

Bifaces

Centripetal Levallois

Parallel and Convergent Levallois

Plaquettes

N

Total Tools

Total Artefacts

60 69.77% 26

38 44.19% 6

3 3.49% 3

6 6.98% 5

7 8.14% 2

86

94

319

52

60

259

50.00% 7 77.78% 1

11.53% 7 77.78% 1

5.77% 0 0

9.62% 1 1 1

3.85% 0 0

9

-

-

3

-

-

33.33% 67 70.53% 27

33.33% 45 47.37% 7

3 3.16% 3

33.33% 7 7.37% 6

7 7.37% 2

95

-

-

55

-

-

49.09%

12.73%

5.45%

10.91%

3.64%

15

6

1

11

2

63

70

544

23.81% 11 50.00% 5

9.52% 6 27.27% 0

1.59% 0 0

17.46% 3 13.64% 3

3.17% 1 4.55% 1

22

-

-

16

-

-

31.25% 26

12

1

18.75% 14

6.25% 3

85

-

-

30.59% 20

14.12% 6

1.18% 1

16.47% 14

3.53% 3

79

-

-

25.32%

7.59%

1.27%

17.72%

3.80%

93

57

4

21

10

180

-

-

51.69% 47

31.67% 13

2.22% 4

11.67% 20

5.56% 5

134

-

-

35.07%

9.70%

2.99%

14.92%

3.73%

Table 6.2: Analysis of retouched tools at the Weinberghöhlen, Mauern.. Figures for Bohmers’ and Zotz’s excavations are presented separately and then summated to produce total figures for each of the cultural units, i.e. Bohmers’ Layer F ‘Altmühlian’/Zotz’s Layer F1 ‘Praesolutréen I’, and Bohmers’ Layers G'/G ‘Moustérien’/Zotz’s Layers F2/G ‘Praesolutréen I’. The sum totals for all layers considered are given at the bottom of the table. N=number of pieces examined; cited percentages are calculated with reference to this figure. Figures for ‘Total tools’ and ‘Total artefacts’ are drawn from Bohmers’ 1951, Tabs. 2 and 4. No comparable data can be extracted from Zotz (1955), so no figures are presented in these categories for either for Zotz’s excavations or for the summated totals. Counts both including and excluding the leaf point component of the assemblages are shown, although this does not apply to Bohmers’ Layer G'/G since he recovered no leaf points from those stratigraphical units.

107

indication of the total number of artefacts recovered (Zotz 1955, 96-122). Sums of total tools or artefacts for Bohmers’ and Zotz’s excavations together therefore cannot be given. The results of the analysis are summarised in Tables 6.1 (Sesselfelsgrotte) and 6.2 (Weinberghöhlen, Mauern).

significant portion of the blank total, with parallel and convergent Levallois marginally more common. Including leaf points in the analysis depresses the incidence of Levallois blanks at Mauern slightly; however, the point made earlier that, in all of these industries, the practice of flat bifacial retouch is likely to lead to an underestimation of the true incidence of Levallois blank production must be born in mind. Taking into account the inevitable technological and typological variability one expects to find between sites that has already been noted, and assuming that the technological and typological similarity between the industries can be understood as an index of the social and cultural connectedness of their authors, there is good reason to reach the general conclusion that the people who discarded the Sesselfelsgrotte G-Komplex and Mauern Layer F/F1 lithic materials shared an identical or very closely related socially-transmitted body of lithictechnical knowledge.

The limitations on the value of the results for comparative purposes should be recognised. The two sites are some 60km apart. Schambach is closer to Mauern and might therefore properly be thought a more appropriate site to compare with it; unfortunately Schambach is not nearly so well understood as the Sesselfelsgrotte. The Mauern industries are much smaller than those of the Sesselfelsgrotte, and that might compromise the value of the comparisons between them. On the other hand, the comparison between an intensively occupied central site and an occasionally-visited peripheral site is the point of the exercise. Finally, one should not expect the properties of lithic assemblages discarded at locations whose social roles in the construction of Middle Palaeolithic landscapes of action were very different to conform identically to technological or typological ‘cultural’ templates; on the contrary, one should expect differences based on variations in, for example, access to and use of raw materials, occupation intensity and the range of actions and tasks carried out between the sites. Nevertheless, the data presented here can illuminate certain gross aspects of how stone was worked and therefore highlight the extent to which particular lithic reduction and production practices, as applications of socially transmitted knowledge, might or might not have been shared in common by the people responsible for fabricating the several inventories and assemblages recovered from the Sesselfelsgrotte and the Weinberghöhlen, Mauern.

A closer inspection of the data, including also the industry from Mauern Zone 5 (Layers G'/F2/G), reveals more detailed patterns of similarity and difference. If we measure Richter’s interpretations, discussed above, of the industrial sequences at the two sites against the data, it is immediately obvious that his characterisation of Mauern Zone 5 (i.e. Layers G'/F2/G) as ‘M.M.O.-B2’, and of Zone 4 (Layers F/F1) as ‘M.M.O.-B3’ with few bifaces and an exclusive reliance on parallel and convergent Levallois reduction, is groundless. In fact, the emphasis on parallel and convergent Levallois blanks is greater in Layers G'/F2/G (1.18% centripetal and 16.47% parallelconvergent if leaf points are included; 1.27% and 17.72% respectively if they are not) than in Layers F/F1 (3.16% centripetal and 7.37% parallel-convergent with leaf points; 5.45% and 10.91% respectively without). Equally, both bifacial working and bifaces are more important in Layers F/F1 than in Layers G'/F2/G, whether one compares the two units with leaf points included (70.53% bifacial working and 47.37% bifaces in Layers F/F1, as opposed to 30.59% and 14.12% respectively in Layers G'/F2/G), or with leaf points excluded (49.09% and 12.73% respectively in Layers F/F1, as opposed to 25.32% and 7.59% respectively in Layers G'/F2/G). Clearly, Richter’s construction of a M.M.O-B2 to M.M.O.-B3 industrial succession at Mauern must be rejected. In turn, this calls into question his assertion that the Mauern Zone 4/ Layers F/F1 industry is of a type that does not occur in the Sesselfelsgrotte G-Komplex.

Figure 6.8 affords a first-order comparison between the Sesseslfelsgrotte and Mauern industries. This plot considers only the Zone 4 (Layers F/F1) materials at the latter site since it is in those deposits that the great bulk of the leaf points occur. It is clear that, when leaf points are excluded, the industries from the Sesselfelsgrotte GKomplex and the Mauern Layer F/F1 are closely comparable. The bifacial working is more common at the latter (49.09%) than the former (34.18%), although bifaces comprise a higher proportion of the tools in the G-Komplex (23.32%) than in Mauern Layer F/F1 (12.73%). Nevertheless, the relative frequencies of these two indices are broadly similar in both industries when leaf points are left out of account. Only when the Mauern leaf points are incorporated in the analysis do the industries seem more different in these dimensions.

Comparisons between the data for Mauern and those for the Sesselfelsgrotte are more problematical, for the reasons discussed. However, the analysis does challenge Richter’s view that Mauern Zone 5/Layers G'/F2/G and Sesselfelsgrotte inventory A01, the youngest in the GKomplex, are closely similar and both referable to the M.M.O.-B2. The strong emphasis on parallel-convergent over centripetal Levallois reduction in Mauern Layers G'/F2/G is at odds with the values for Sesselfelsgrotte A01, where 2.68% of the examined tools are made on

The similarity between the Sesselfelsgrotte G-Komplex and Mauern Layer F/F1 is even better expressed when the relative frequency of Levallois blanks is considered. Here the inclusion or exclusion of the leaf points, unsurprisingly, makes less difference. In both industries centripetal Levallois blanks constitute a small but

108

80

Sesselfelsgrotte G-Komplex 70

Mauern F/F1 excluding leaf points 60

Mauern F/F1 including leaf points % Total Tools

50

40

30

20

10

0 Bifacial Working

Bifaces

Centripetal Levallois

Parallel and Convergent Levallois

Plaquette

Figure 6.8. Histogram showing comparison between the lithic industries from the Sesselfelsgrotte G-Komplex and from LayerF (Bohmers) and F1 (Zotz) at the Weinberghöhlen, Mauern.

centripetal Levallois blanks (around twice the proportion found in Mauern Layers G'/F2/G) and just 4.70% on parallel-convergent Levallois blanks (only around one quarter of the value in Mauern Layers G'/F2/G). Again, this challenges the notion that the Mauern Layers F/F1 industry is later than any inventory present in the Sesselfelsgrotte G-Komplex since it undermines the claim that the older Layers G'/F2/G industry at the former site is comparable with the youngest, and only the youngest, inventory at the latter. One must conclude, therefore, that Richter’s scheme for the comparative chronology of the Sesselfelsgrotte and the Weinberghöhlen, Mauern, does not withstand testing against the data presented in this analysis, and that there is no reason to think that the Mauern Layers F/F1 ‘Altmühlian’ leaf point industry is not contemporaneous with some part of the Sesselfelsgrotte G-Komplex.

Layers F/F1 industries at Mauern is unlikely. The same might be said of inventories A12, where, unlike in either unit at Mauern, centripetal Levallois blanks (12.50%) are more frequent than parallel-convergent (8.33%), and A11, from which centripetal Levallois blanks are absent. These inventories together represent, if Richter’s analysis of the occupation of the Sesselfelsgrotte is accepted, the first three of four occupation events in cycle 1, the whole of cycle 2 and the first occupation event of cycle 3. The occupations responsible for the Layers G'/F2/G and Layers F/F1 industries at Mauern are best placed in GKomplex cycle 4 or in the Konsekutivinventäre of cycle 3. This implies in turn that the site’s appropriation into a scheme of movement through and action in the Altmühl Valley during the Middle Palaeolithic of OIS 3 was a feature of landscape occupations of some time depth. Beyond this, attempts to correlate the industries of either Layers F/F1 or Layers G'/F2/G at Mauern with particular inventories in the Sesselfelsgrotte G-Komplex are fraught with danger. Comparisons of the frequency of bifaces and of bifacial working do not help. The values for bifacial working in inventories A06-01 are relatively constant, with a minimum of 28.22% (A06) and a maximum of 37.36% (A04); the figure for Mauern Layers G'/F2/G falls within this range (30.59% with leaf points included, 25.32% without), but bifacial working in Layers F/F1 is significantly higher, both with the leaf points (70.53%) and without (49.09%). On the other hand, bifaces are less

For the reasons discussed earlier, one would not expect to find technological and typological identity between particular archaeological units at the two sites, even if they were strictly contemporaneous and were fabricated and discarded by individuals from a single, synchronic social group. However, the paucity or absence of parallel and convergent Levallois blanks in the G-Komplex inventories A13 (absent), A09 (absent), A08 (0.94%) and A07 (absent) would seem to suggest that their incorporation into the same mode of action in the landscape as is represented by either the Zone 5 or the 109

similarities between their industries (presence of bifacial working, bifaces, centripetal and parallel-convergent Levallois blanks and tabular plaquettes) demonstrated in this investigation suggest strongly that the Mauern industries were produced by peoples equipped with the same body of socially-transmitted body of knapping knowledge as was applied by the various people responsible for the industries of the G-Komplex; the comparison is closest with the inventories of cycles 3 and 4. The evidence therefore favours an interpretation of the Altmühl Valley Middle Palaeolithic in which ‘Micoquian’ lithic-industrial practices from which the production of leaf points was virtually absent, and ‘Altmühlian’ practices in which the production of leaf points was important, were not characteristic of two discrete cultures each recognisable by diagnostic stone tool types and fabrication techniques, but were aspects of the same body of socially transmitted knowledge as to how to act in the landscape. From this point of view, the leaf point industries at Mauern and at least some of the Micoquian inventories at the Sesselfelsgrotte can be seen as specific and knowledgeable applications, in particular social and landscape contexts, of a shared mode of engaging with the world. One might describe them as components of a single settlement and mobility system in which the Sesselfelsgrotte was an intensively occupied social centre at which a wide range of tasks and actions were carried out, and the Weinberghöhlen, Mauern, was an occasionally visited peripheral place, a ‘provisioned place’ (Kuhn 1995) utilised as a leaf point cache. Or, in the terms outlined in Chapter 2, the various acts perpetrated at the two sites can be seen as aspects of the same project for living.

important in both Mauern archaeological units if leaf points are excluded (7.59% in Layers G'/F2/G and 12.73% in Layers F/F1), and in Layers G'/F2/G if leaf points are included (14.12%), than in any of inventories A06-01, where the lowest value is 14.38% (A05) and the remainder fall between 19.46% (A01) and 23.08% (A04). The relative importance of bifaces at the Sesselfelsgrotte might be linked to the greater incidence there of tools made on tablets or plaquettes of Plattenhornstein, a major source of which, Baiersdorf, is just 4km northwest of the site. As with the incidence of bifaces and of bifacial working, so with Levallois blanks; in each of the GKomplex inventories A06-01, parallel-convergent blanks outnumber centripetal blanks by about 2:1. The incidence of centripetal Levallois blanks varies between 2.50% (A03) and 6.16% (A05), whereas parallel-convergent blanks vary between 3.75% (A03) and 16.57% (A02), although this latter figure is something of an outlier; the next highest value is 10.96% (A05). Bearing in mind the expectation of difference between the industries from the two sites, both the values for centripetal and parallelconvergent Levallois blanks, and the ratios between them, in inventories A06-01 are broadly consistent with the industry from Mauern Layers F/F1, where centripetal Levallois blanks account for 3.16% and parallelconvergent blanks account for 7.37% of the examined pieces if leaf points are included (5.45% and 10.91% respectively if the leaf points are excluded). The importance of parallel-convergent Levallois reduction in the industry from Mauern Layers G'/F2/G, where it reaches 16.47% with the leaf points and 17.72% without, is matched in the G-Komplex, as has been pointed out, only in inventory A02.

This view requires some qualification. Chapters 4 and 5 showed that Middle Palaeolithic leaf point sites in Europe occur only in OIS 3, and only in those regions of Europe that experienced pulses of afforestation and deforestation on sub-Milankovich wavelengths comparable with that of the D-O climatic events visible in the GRIP ice core oxygen isotope curve. Evidence has been advanced here to show that the Altmühl Valley experienced just such landscape transformations during the time span in which the Sesselfelsgrotte G-Komplex inventories and the Mauern industries of Layers F/F1 and G'/F2/G were deposited. These archaeological traces therefore refer to a region and a period in which the affordances of the landscape were repeatedly transformed, thereby challenging the viability of socially transmitted knowledge as to how to act in and relate to the material and social world. Furthermore, it has been shown in Chapter 4 that Middle Palaeolithic leaf point sites in Europe lack the archaeological properties associated with frequently or intensively occupied ‘residential’ sites or centres of social settlement, but invariably suggest ephemeral occupation, often in an apparently taskspecific, often workshop, context. That the archaeology of Layers F/F1 and G'/F2/G at the Weinberghöhlen indicates that it was only occasionally occupied has also

Discussion It is worth making several points: that Zotz himself admitted that one leaf point from his Layer G at Mauern might have fallen into that horizon from Layer F1 (Zotz 1955, 131) and that, given his antipathy to Bohmers and his desire to push the ‘Praesolutréen’ back in time, there could therefore be doubt over the provenance of the remaining five leaf points attributed by him to Layers G'/F2/G (Allsworth-Jones 1986, 68); that the high frequency of parallel-convergent Levallois blanks in Mauern Layers G'/F2/G is matched in the Sesselfelsgrotte G-Komplex only in inventory A02; and that the only leaf point present in the G-Komplex, a fragment illustrated here at Figure 6.9, occurs in inventory A01. One can therefore speculate that the Mauern Layers G'/F2/G industry was produced by the same people that produced the G-Komplex inventory A02 at the Sesselfelsgrotte, that there is in truth only one leaf point industry at Mauern, in Layers F/F1, and that it was produced by the same people as made the G-Komplex inventory A01. Of course, it is impossible to demonstrate this. However, given the dating of both the G-Komplex at the Sesselfelsgrotte and of Layers F/F1 and G'/F2/G at Mauern to the early part of OIS 3, the qualitative

110

Figure 6.9. Leaf point fragment from the Sesselfelsgrotte G-Komplex, inventory A01. (After Richter 1997, Taf. 18. 4).

been shown. The clear association between Middle Palaeolithic leaf points and repeated material challenges to projects for living, both in the Altmühl Valley and in Europe as a whole, is therefore evidence that leaf points were aspects of a knowledgeable response to landscape transformations taking place over a few centuries or even decades, a response effected by people whose purpose was to incorporate into their project for living the peripheries of landscapes which, in certain of their aspects, were rapidly becoming unknown. The implications of this for an ecological geography of the European Palaeolithic are discussed in the next Chapter.

certainly or probably referable to OIS 3 that can convincingly be interpreted as more-or-less intensively occupied social centres are the Sesselfelsgrotte and Schambach. The leaf point sites – Obernederhöhle, the Weinberghöhlen, Groβe Ofnet, Kleine Ofnet, Steinerner Rosenkranz and Biesenhard – all have the character of peripheral places at which particular tasks were carried out. It is also noteworthy that the two most important leaf point sites, Obernederhöhle and the Weinberghöhlen, are located in dry valleys, i.e. in topographical features which persisted through climatically-induced transformations in the landscape. Dry valleys would therefore have retained certain key affordances, such as providing favoured seasonal migration routes for ungulates between the higher ground to the north of the Wellheimer Trockental and the lower ground to its south, even as other affordances within the landscape - visibility, the seasonal availability of fruits and other plant foods, herd aggregation or dispersion behaviour and the existence and viability of paths through the landscape - were undergoing rapid and radical change.

Leaf Point Sites in the Wider Altmühl Valley Region This analysis has, for the reasons stated, concentrated on just two sites. It has already been pointed out that the Middle Palaeolithic occupants of Mauern might easily have been connected with a centre of social action at Schambach as well as, or even instead of, that at the Sesselfelsgrotte. The possibility of other, as yet undiscovered central sites of unknown number also having been implicated in the lives of the Mauern leaf point makers must also be borne in mind. The question, therefore, is to assess the degree to which the leaf point sites and the more systematically occupied sites of the Altmühl Valley in OIS 3 might conform to the relationship advanced here for the Sesselfelsgrotte and the Weinberghöhlen.

Certain leaf point sites are close to the Sesselfelsgrotte or Schambach; the Obernederhöhle is some 5km distant from the former, and therefore is highly likely to have been used by individuals whose social base at the time was the Sesselfelsgrotte. The very small industry from the Breitenfurter Höhle, just 1km from Schambach, contains one piece that might be a leaf point, so this site might be thought of as playing a similar role to the Obernederhöhle. However, its industry is so small – just eight artefacts – that one must also consider the possibility that it was occupied just once in an entirely opportunistic manner, for example to shelter from a storm or to tend a wounded individual. Other leaf point sites with very small industries, such as Steinerner Rosenkrantz and the Groβe and Kleine Ofnet caves, should also be thought of in this way. Biesenhard represents just a single leaf point found on the surface.

The first point to make is that the associations between leaf points and small assemblages in ephemerally occupied sites, and between leaf point-poor Micoquian industries and sites with evidence for occupation as a social centre, hold, as discussed earlier in this Chapter. Within the Altmühl Valley sensu stricto, the only sites 111

However, although these sites can scarcely be understood as recognised places in a Middle Palaeolithic landscape of movement and action in the way that Mauern and Obernederhöhle can, they nevertheless support the association between leaf points and action distant from the social centre.

the Altmühl Valley is not sufficiently robust to permit even the kind of analysis performed here on the Sesselfelsgrotte and Mauern industries. At the same time, the leaf point sites in these regions are all characterised by small assemblages with no evidence for persistent or repeated occupation as a centre of social life, and, insofar as they are datable at all, are either probably or possibly referable to OIS 3. To that extent they support the pattern apparent in the Altmühl Valley.

If one considers the wider Upper Danube region, this association continues to hold. Two leaf points (see Appendix 2 Fig. A2.30) are known from the Haldenstein cave (Bosinski 1967, 151) in the Lone Valley, southeastern Baden-Wurttemberg. Unfortunately most of the sediments were removed in 1865, and the stratigraphy subsequently established by Riek (1938) must be regarded as schematic. Nevertheless, their presence in the same tributary valley as the Bockstein, where there are no leaf points, is consistent with the situation in the Altmühl Valley; indeed, Allsworth-Jones (1986, 71) believes the Haldenstein leaf points to be contemporaneous with those from Mauern Layers F/F1.

Conclusions This investigation of the role of leaf points in the use of the Altmühl Valley by Middle Palaeolithic peoples in OIS 3 is necessarily limited by the paucity of wellexcavated, well-understood sites. This should not detract from the robustness of the patterns that do exist. As in the rest of Europe, leaf points are restricted to non-residential contexts and are associated, as far as the dating of the sites permits, with a period in which understood landscapes were undergoing repeated and rapid climatically-induced transformations. The comparison between the Sesselfelsgrotte and the Weinberghöhlen, Mauern, indicates a high probability that, in one or two particular periods likely to be associated with these transformations, the two sites were implicated in a mode of action in the landscape in which leaf points were associated specifically with action distant from the centre of social life. This is also consistent with the Middle Palaeolithic archaeology of the wider Upper Danube region.

Several more leaf point sites are known from the Upper Danube region to the east of Kelheim. Two localities at Zeitlarn, near Regensburg and approximately 30km northeast of the Sesselfelsgrotte, have produced small, undatable surface collections with bifacial leaf points (Figure A2.38) ascribed to the Szeletian on the basis of the presence of endscrapers (Schönweiβ and Werner 1986), although this is open to challenge. The claimed endscrapers are made on flakes or ‘broad blades’ which, to judge from the illustrations, could have been produced by a Levallois technique (Schönweiβ and Werner 1986, Abb. 3, 4-6) and are retouched around their entire periphery; in fact they are very similar to pieces in inventory A01 in the Sesselfelsgrotte G-Komplex and described by Richter as Mikrokratzer (microendscrapers). One piece could also be described as a small handaxe or Fäustel. It is difficult to identify a possible central site, other than the Sesselfelsgrotte, in this region, although Sinzing, a cave site in the Naab Valley close to its confluence with the Danube, is a possible candidate. Unfortunately the site was excavated in the 19th Century and its Middle Palaeolithic artefacts, which include Micoquian bifaces but no leaf points, have scarcely been studied (Bosinski 1967, 171-2). Another surface-collected industry with bifacial and unifacial leaf points and handaxes, but lacking endscrapers, and placed in the Middle Palaeolithic on the basis of typology and condition, has been recovered from Albersdorf (Figure A2.39), 100 km downstream and to the southeast of Zeitlarn and close to the Czech and Austrian borders (Weiβmüller 1995b). Of the 50 tools leaf points are the most common type present and are quite distinct from the handaxes. Less clear is the surface material from nearby Flintsbach-Hardt (Weiβmüller 1995b), where bifaces, including one bifacial leaf point, exhibit a continuum of form (Figure A2.40).

When this pattern is taken together with the wider pattern of European leaf point site distribution in time and space, a crucial scalar dynamic to knowledgeable human action in the Middle Palaeolithic of OIS 3 emerges. Leaf points as an artefactual form are closely coupled with cyclic landscape perturbations on wavelengths of a few millennia; and as a mode of action on the world, they appear to be an aspect of knowledgeable responses to radical changes in landscapes of affordance taking place over much shorter, even biographical time spans, given the evidence from the GRIP ice core oxygen isotope record for increases in global mean annual temperatures of around 8°C in perhaps as little as ten years during OIS 3 (Dansgaard et al 1993), and the rapidity with which afforestation has progressed under favourable conditions (the end of grazing by sheep) in the Altmühl Valley over the last century. Furthermore, leaf points represent a phenomenon that is scarcely, if at all, known in earlier periods of the Palaeolithic, namely the association of a specific tool type with tasks carried out in particular parts of the landscape. What these conclusions might mean for an ecological geography of the European Palaeolithic is discussed now in Chapter 7.

Clearly, the evidence from the regions of the Upper Danube to the west, and more especially to the east, of

112

CHAPTER 7

CONCLUSIONS: THE SCALAR CONVERGENCE OF STRUCTURE AND AGENCY THROUGH THE EUROPEAN PALAEOLITHIC Although this approach remains at the level of the general and the exploratory, it contains one key proposition that, it is suggested, is essential to its future development, and even to the development of a broader and deeper palaeolithic archaeology, namely the dissolution of the dualism between ecology conceived as a set of material relations between organisms as objects on the one hand, and the knowledgeability of the subject on the other. Knowledge mediates and defines the relationship between the acting, behaving organism and the material world in which it is immersed, a relationship which was glossed under the term project for living in Chapter 2 and which Clark regarded as essential for an economic prehistory (Chapter 1). From this perspective knowledgeable subjectivity is explicitly ecological in that it is an irreducible component of the linkage between organisms, including human beings, and the environment that surrounds them. At the same time, knowledge is not a unitary phenomenon but is manifest differently at different scales. Trans-generational knowledge, whether biologically embodied or culturally inscripted, emerges from and defines a mode of engagement with the world, equips individuals with a scheme for life and thereby renders the properties of the world as opportunities and hazards specific to that way of life. Knowledgeable action, on the other hand, is the contextually appropriate and creative application of these structured and structuring principles of engagement by purposeful individual agents to the tasks thrown up by living the project in the world of practical sensed experience. The two realms of knowledge are mutually reproducing and transforming, with the sensitivity of each to changes in the other dependent upon the characteristic process rates of the mechanisms that link them. For human beings, those mechanisms can be brought together under the heading of the social transmission of knowledge. Because the environmental affordances to which knowledgeable action is directed are not independent of the physical properties of the world, changes in the world demand new forms of knowledge. This investigation has aimed to illuminate both the changing sensitivities of socially transmitted knowledge and knowledgeable action to transformation each by the other, and to identify the changing temporal and spatial scales of difference in the world to which each was characteristically coupled, through the course of the European Lower and Middle Palaeolithic. The approach to ecological geography developed here is therefore a hierarchical ecological geography of knowledge.

Introduction This investigation is not so much intended to provide definitive answers as to look at the Palaeolithic in a different way so that new kinds of questions can be asked; or, more specifically, so that old questions asked by McBurney nearly 50 years ago can be posed again and approached from a different angle. His three key observations noted in Chapter 1 are as pertinent as ever. Indeed, he pinpointed the central problems of climatic seasonality and the plains-upland opposition that have been identified here as two crucial constraining and enabling factors in the ecological geography of human settlement in the European Lower and Middle Palaeolithic. What he did not have was any explicit conception of the multiplicity of temporal and spatial scales on which the character of the material world varies, or of the dynamics that link them. Ecological thinking in 1950 remained committed to Clementsian concepts of succession, climax and the plant community as reified superorganism. Watt’s visionary work on the cyclical upgrade-downgrade hypothesis, discussed in Chapter 2, was published just 3 years before The Geographical Study and remained largely unnoticed until the 1970s (O’Neill et al 1986). There is, in any case, no reason to imagine that McBurney would have grasped its significance even if he had been familiar with it - in The Geographical Study McBurney’s gaze is exclusively continental in scope, whereas Watt was looking at small gap dynamics and forest regeneration. One could say that each lay outside the other’s dynamic scale range. The fundamental conclusion to be drawn is that an explicit sensitivity to the question of scale is a necessary prerequisite for the kind of ecological-geographical palaeolithic archaeology towards which McBurney was groping. Our primary objects of study, chipped stone tools, were simultaneously the products of the moment and of the ages, while climates and environments change by the second, the day, the decade and the millennium. The project for an ecological palaeolithic archaeology cannot make much headway if it fails to identify those particular scales, dimensions and aspects of change and persistence in the world that relate to particular scales, dimensions and aspects of past human action and knowledge. The approach developed here is advanced as a first step towards such a project.

113

production as a knowledgeable practice, and its response to raw material constraints, therefore seems to show no net change over an immense period of time or through radical transformations in the material character of the world. By comparison, the Middle Palaeolithic industrial successions documented across the continent and discussed in Chapter 3 reveal a significant increase in the rate of change in knapping practices, and thus of socially transmitted knapping knowledge.

Scales of Structure and Agency This investigation outlines one way in which such an ecological geography of the Palaeolithic might be effected. The process of eastward expansion of populations from western Europe that (as was shown in Chapter 3) began around 200 kya represents an opening up of new landscapes and regions for human occupation, a breaking of the bonds that had previously limited the occupation of Europe to the persistent fine-grained lowland mosaics of the oceanic west. What is more, the occupation of upland landscapes was integrated with the occupation of lowlands. Middle Palaeolithic peoples in Europe therefore constructed lived environments defined in part by a tension between two materially different and spatially discrete aspects, of differing topographical and biotic grain, that no individual could occupy simultaneously. This transformation took place not because the world had fortuitously changed so as to become more hospitable for Neanderthals - there is nothing in the Quaternary record to suggest that Europe in the Upper Pleistocene was in some essential way more humanly habitable than in the Middle Pleistocene - but because the terms of human immersion in the world had changed, with an expansion in the temporal and spatial reach of knowledgeable action without which landscapes characterised by higher levels of distanciation between resources in time and space would have been, as before 200 kya, uninhabitable. Whether one conceives of this as an increase not so much in the length of chaînes opératoires (understood here as linked sequences of acts of all kinds rather than simply the reductive working of stone) as such, but in the length of chaînes opératoires factored into decision making by knowledgeable agents, or in terms of planning depth (Binford 1979), the incorporation into human perspectives of new orders of difference – or, in terms employed in Chapter 2, an expansion in human grain responsiveness - transformed barriers into opportunities. By comparison with the Palaeolithic prior to 200 kya, the subsequent occupation of Europe can be characterised precisely by this incorporation of difference – between uplands and plains, summer and winter, proximity and distance, presence and absence, now and then.

It is therefore proposed that, from around 200,000 years ago, the scale domains of socially transmitted knowledge and of knowledgeable action – or, in sociological terms, structure and agency - underwent a convergence, and that that convergence is correlated with the capacity of the Middle Palaeolithic peoples of Europe to construct new projects for living on time scales over which their Lower Palaeolithic predecessors apparently retained, unchanged, the same stone working and landscape settlement practices. Structure and Agency in the Palaeolithic: Beyond the 200 kya Boundary The Acheulean Before 500 kya Perhaps, in the light of this discussion, we should be promoting the idea of a Lower-Middle Palaeolithic ‘revolution’ as an event. Of course, such a proposition would best be tested by the detailed examination and comparison of settlement distributions, climateenvironment histories and lithic-industrial practices in time slices of, say, 10 or 20 millennia, commencing with the earliest occupation of Europe. This was obviously not possible here and is probably impracticable in principle given the coarse resolutions of the Middle Pleistocene archaeological, palaeoclimatic and palaeoenvironmental records as currently known. However, it is possible to draw general conclusions about the Palaeolithic before 500 kya and, as a consequence of this investigation, much more specific conclusions about the Middle Palaeolithic after 60 kya. In the 1.3 million years or so following the emergence of the handaxe in East Africa some time around 1.6 Mya, handaxe-making hominids colonised the Old World from South Africa to Britain to India. This indicates that, on evolutionary time scales, some aspects of Acheulean projects for living underwent radical transformation, insofar as the colonisation of temperate zones, with summers and winters, and with day lengths varying through the year, represents a shift from the routines of life in the tropics. This implies some expansion in the spatiotemporal reach of knowledgeable action. The same might be said of the emergence in East Africa of the Acheulean itself, associated as it is with a shift from Oldowan occupation of Rift Valley axial watercourses and riverine woodland to the exploitation of marginal volcanic uplands, and with the emergence of the long-

The discussion in Chapter 3 also drew attention to the emergence in the European Middle Palaeolithic of structured change in stone working practices on ecological time scales of a few tens of millennia, and the contrast this represents with the Lower Palaeolithic. Although there is, of course, difference within the European Acheulean, it lacks any pattern that could be described as structured change through time. The relationship between the production of pointed or ovate handaxes and raw material size and shape demonstrated by White (1998) for the British Acheulean holds over a time span of some 250-300,000 years between OIS 13 and OIS 8, and probably longer, and transcends glacialinterglacial climatic cycles. Acheulean handaxe

114

legged, wide ranging obligate biped Homo ergaster (papers in Oliver et al 1994). However, human settlement throughout remained closely coupled to fine-grained landscapes in which all resources occurred in close proximity, so caution must be exercised in judging the import of these developments. Equally, it is difficult to detect any connection between this colonisation process and stone working practices, if the stability in handaxe form suggested by large scale cross-continental comparisons (e.g. Wynn and Tierson 1990) is to be believed. The Schöningen spears (Thieme 1997) reveal another dimension of artefact production which might conceivably have undergone more systematic change through the Lower Palaeolithic than did stone tool fabrication, but the record of wooden artefacts in the Palaeolithic is so vanishingly small as to render all such consideration speculative. One can say, then, that some expansion in the reach of knowledgeable action took place in the Lower Palaeolithic prior to 500 kya, but that the hard evidence for an increase in the rate of change of structured, socially transmitted knowledge in that period is slight.

earlier Middle Palaeolithic structured change through time in the properties of stone tool industries, although apparent on ecological timescales, cannot be unequivocally correlated with environment changes on wavelengths of 20 millennia, such as the warm-cold oscillations of OIS 5 (Turq 1999). The leaf point phenomenon is therefore a powerful indication that, after 60 kya, the social transmission of knowledge had become radically more sensitive to transformation through time – that is, operated at a higher process rate – than in any previous period in the continent’s occupation history. Furthermore, it was shown in Chapter 6 that the settlement of the Altmühl Valley in the millennia after 60 kya entailed an opposition between action at the centre and at the periphery, indicating the implicit presence in the here-and-now of other places, other times and other tasks. The use of the Wellheimer and Ziegel dry valleys, which would certainly have afforded opportunities that persisted through millennial-scale landscape transformations (the Wellheimer Trockental as a favoured seasonal migration route for ungulates, and the much smaller but steep-sided Ziegeltal as a sheltered local environment always liable to support a somewhat different array of plant and animal resources than the narrow flood plain or the plateau top, for example) heightens this sense of the knowledgeability of action stretching over time and space. The restriction of leaf points to peripheral places also indicates that specific practices were developed in the context of rapid and radical changes in the environment as evidenced in the pollen record of Bavaria. A picture therefore emerges of the Altmühl Valley Middle Palaeolithic in OIS 3 in which innovative applications of pre-existing knowledge – the use of bifacial stone working methods to fabricate leaf points, for example – emerged in response to new practical problems, and were institutionalised as received ways of acting in specific concrete contexts. This constitutes a degree of mutual sensitivity between socially transmitted knowledge and knowledgeable action, and a consequent capacity to transform projects for living, which has not been documented in the European Middle Palaeolithic older than OIS 3.

The Late Middle Palaeolithic Chapters 4 and 5 showed that Middle Palaeolithic industries with leaf points as discrete types, rather than as one pole of a continuum of biface form, are found in an OIS 3 space-time zone corresponding to that in which landscapes underwent alternations between open-steppic and woodland vegetation (by which is meant oscillations in the spatial extent and/or number of woodland patches within the fine-grained mammoth steppe mosaic) on D-O wavelengths of around 1-2.5 millennia, and that they invariably possess archaeological properties inconsistent with their formation through use as centres of social action. The examination in Chapter 6 of the Altmühl Valley Middle Palaeolithic confirmed this pattern at a landscape scale. It also showed that the Sesselfelsgrotte and Mauern were very probably complementary places in the early OIS 3 occupation of the Altmühl Valley, with the former site being a central focus of social life and the latter an occasional locus for specific tasks perpetrated in the landscape periphery.

Of course, the incorporation of uplands and plains into landscapes of action, and the enhanced capacity to act in the present in a manner that accounted for other times and other places, emerged in Europe at around 200 kya. There is also some direct evidence for the use of particular points in the landscape for specific purposes prior to 60 kya. Turq (1988) has developed a classification of the open air Middle Palaeolithic of southwestern France that recognises quarry, workshop, mixed activity and temporary stopover sites. To this one might add butchery sites such as Mauran (Farizy et al 1994) and La Borde (Jaubert et al 1990). The dating of these sites is highly problematic, but it is obvious that Middle Palaeolithic peoples prior to 60 kya must have moved and acted in the landscape outside caves and rock shelters, and that some at least of these French open air sites must predate OIS 3.

These patterns are highly significant from the point of view of a hierarchical ecological geography. Transformations in the character of the vegetational landscape are clearly of especial import for hominids. Factors such as primary biomass productivity, biodiversity, landscape visibility, ease and routes of movement, mammalian fauna and their spatial aggregation and mobility patterns are all strongly coupled to the character of plant communities, particularly to the extent of tree cover. The strong association between leaf points and sub-Milankovich wavelengths of transformations in these factors, and thus of landscapes of affordances, stands in stark contrast to the Lower Palaeolithic insensitivity of lithic practices to environment transformation on any scale. Even in the

115

In Bulgaria, Musselievo is 120km to the northeast of Samuilitsa 2; more importantly, it is about 80km north of the major stratified cave site of Bacho Kiro, where two ‘Levalloiso-Mousterian’ horizons occur in layer 12 (Kozlowski 1975; Allsworth-Jones 1986). These industries lack leaf points and bifaces entirely, but are technologically closely similar to the Musselievo leaf point industry which has been described as ‘LevalloisoMousterian with leaf points’. Layer 12 overlies a layer radiocarbon dated to >47,000 BP and underlies an early Aurignacian or ‘Bacho Kirian’, and is probably of an early OIS 3 age close to that of Musselievo. The situation is even less well understood in Greece, although it is apparent that the leaf points of the northwest are known only from open air contexts (e.g. Kokkinopilos and Morfi) and are absent from the nearby rock shelter site at Asprochaliko (Papagianni 1999). In both Bulgaria and Greece, prospection for Palaeolithic sites has been very thin, so more confident conclusions as to relationship in the Balkans as to the landscape relation between sites with and without leaf points in OIS 3 must await the outcome of future research. This is true also of the Adriatic coast of Croatia which, given its position windward of its high-altitude Balkan hinterland, might be expected to have been part of the OIS 3 Middle Palaeolithic leaf point province.

Indeed, open air sites with lithic assemblages like those attributable in cave and rock shelter contexts to the preOIS 3 Ferrassie and Quina Mousterian facies are known in southwestern France (Mellars 1996, 260-2), and La Borde has been placed in an OIS 5 warm stage (Laville 1990). However, most are placed by implication in OIS 3 on the grounds of the presence of cordiform handaxes and their consequent placement in the OIS 3 MTA. The best known workshop site in the region, Lagrave, was certainly a site of cordiform handaxe fabrication (Turq 1988). In short, the degree of systematisation of action at the centre and periphery evident in the Altmühl Valley in OIS 3 cannot be shown to have been present in southwestern France before OIS 3. A similar conclusion can be drawn from the Sesselfelsgrotte ‘Unteren Schichten’, in which a wide range of opportunisticallycollected raw materials are represented in the lithic inventories (Weiβmüller 1995), in sharp contrast to the concentration on the Baiersdorf Plattenhornstein apparent in the G-Komplex. The question therefore arises as to whether this systematic relationship between central places without leaf points and peripheral places with them was a feature of the late Middle Palaeolithic leaf point phenomenon throughout the European leaf point zone, or was peculiar to the Altmühl Valley. In central Germany, Micoquian bifaces are present in the surface-collected inventories from Dickershausen, Kirchberg, Maden, Wichdorf and several localities at Fritzlar, all close to the confluence of the Eder with the Weser in Hesse (Bosinski 1967, 12533). The artefacts are made almost exclusively on tabular Kieselschiefer (siliceous slate) which is locally abundant. A single leaf point fragment, made on quartzite, is known from the Hellen-Ost locality at Fritzlar (Appendix 2, Fig. A2.19). This is of interest in that it is closely similar to those from Rörshain, some 25 km to the south. It appears that leaf points made at Rörshain, a workshop site, were subsequently transported around the landscape over significant distances; no major site that might qualify as a centre of social action is known from this non-karstic region.

The situation in The Crimea is worthy of particular note in that, according to Chabai and Marks (1998), Middle Palaeolithic leaf points, although referable to OIS 3, occur predominantly in ‘base camp’ (e.g. Zaskalnaya V) and ‘short term camp’ (e.g. Starosel’e Level 1) contexts and are rare or absent at sites described as ‘ephemeral stations’. This is of interest in that it follows the distinction in the Altmühl Valley between central and peripheral locations but reverses the role of leaf points. Clearly, an authoritative answer to this particular question would require a detailed study of the landscape context of leaf point rich and poor sites throughout central and eastern Europe, and such a study lies beyond the scope of this investigation. The overall pattern as presented here, however, conforms generally to that of the Altmühl Valley. There is good reason, then, to argue that the fabrication, use and discard of leaf points only in the context of the performance of particular acts in particular places was a widespread feature of the Middle Palaeolithic in central and southeastern Europe after 60 kya.

Further to the east, it is possible to associate some, but not all, leaf point occurrences with major sites. In western Hungary, Jankovich is situated on the eastern escarpment of Öregkő range, 5km south of the Danube and 20km to the northeast of Tata. The latter site has produced over 2000 tools, including some 350 bifaces, but no leaf points (Vértes 1964; Allsworth-Jones 1986). However, the Tata industry is tentatively placed not in OIS 3 but in OIS 5c, since the overlying layer has been U-Th dated to 78,000±5000 years ago, and the underlying layer to underlying to 116,000±5000 years ago (Schwarcz and Skoflek 1982). If this date is correct it would disqualify Tata as a social centre out of which leaf point makers operated in OIS 3. On the other hand, the leaf point workshop at Bohunice, Moravia, is just 35km south of Kůlna, where the upper Micoquian levels at least are of OIS 3 age.

The situation in Iberia and Italy has not been considered thus far, and that in southwestern France only briefly. In respect to the last of these, the tendency for task-specific open air sites such as ‘quarries’ and workshops to be associated with cordiform handaxes suggests modes of action specific to particular points in the landscape in southwestern France during OIS 3 (subject to the caveat that the placement of these site assemblages in OIS 3 relies on an analytical correlation with similar assemblages dated to that period in stratified cave and

116

rock shelter sites (Mellars 1967, 1969, 1986) and, to that extent, bears comparison with the situation to the east, even if no association of cordiform handaxes as a specific form with particular theatres of action is apparent. On the other hand, Kuhn (1995) has argued that the differences between Pontinian Mousterian industries older and younger than 55 kya in Latium – an increase in the use of parallel reduction techniques at the expense of centripetal, and a decline in the application of retouch to flakes – represent economic responses to landscape changes induced by variations in sea level, and do not imply any qualitative change in ‘anticipatory organisation’ within the Middle Palaeolithic of that region.

be a single revolutionary event, but a trajectory of development that can, with the hindsight of the archaeologist, be seen to have been in operation throughout the Lower and Middle Palaeolithic.

The Causes of the Convergence of Scales of Knowledge This all raises, of course, a question that this investigation has so far avoided. What exactly might the mechanisms have been that brought the scale domains of structure and agency closer together, exposing practice-structuring principles of socially transmitted knowledge, or projects for living, to the transforming force of their own practical application by knowledgeable and purposeful agents? From the hierarchical point of view, answers to this question need to explain how the ‘black box’ that links these scale domains of knowledge acquired, metaphorically speaking, a higher characteristic ‘frequency’ and became thereby capable of transmitting to the high level ‘constraining’ structural domain variations in practice to which it had previously been ‘blind’. As was explained in Chapter 2, socially transmitted knowledge is manifest and reproduced only in the perpetration by purposeful and knowledgeable agents of contextually appropriate practical acts underwritten by the project for living into which they are socialised; that is, people learn how to understand the world and how to act in it from other people. This is not to deny that people learn how to knap, for example, through the experience of engagement with stone, or acquire familiarity with the location of resources, or the behaviour of prey species, through active experience of the landscape. Indeed, the concept of the project for living as expounded here explicitly recognises the centrality of the world’s material properties to the ways that people devise to live in it. The point, however, is that the world is never encountered neutrally, but always on particular terms that appropriate it into a way of life and furnish a rationale and a theatre for practical action. The question therefore boils down to: what developments in the mechanisms by which knowledge was socially reproduced and transmitted might have facilitated an acceleration in the rate of fixation or institutionalisation of particular or innovative practices through the Lower and Middle Palaeolithic? Two gross categories of explanation, not necessarily mutually exclusive, are conceivable.

There are, however, other lines of evidence for the emergence of new practices after 60 kya. Convincing evidence for features such as pits, postholes and built fireplaces is absent from virtually the entire Lower and Middle Palaeolithic (Stringer and Gamble 1993, 155-6), but the earliest robust claims for such features, at least in Europe, date to that period. The mammoth bone huts or windbreaks of Molodova, especially site I level 4 (Klein 1969), and possibly Ripiceni-Izvor (Păunescu 1993) reveal a way of constructing personal and domestic space – the erection of, followed by entry into, an enclosure – that differs radically from the that of the earlier Palaeolithic, in which space is typically created by pushing debris outwards from the centre (Kolen 1999). The stone-built hearth in a Mousterian context at Vilas Ruivas, Portugal, is another example of the construction of space on a personal, rather than a landscape, scale that is unknown before 60 kya. Finally, Stiner, through her work on strategies for the procurement of animal tissues as food in the Middle Palaeolithic of Latium (Stiner 1991, 1994), has documented a transition from relatively opportunistic meat acquisition strategies, with an important emphasis on scavenging and on the consumption of old individuals, before 55 kya to the systematic practice of ambush hunting and a concentration on prime age individuals after 55 kya. This also represents a qualitative shift in Middle Palaeolithic practices at some time close to the onset of OIS 3. Evidence from several lines of research therefore supports the contention that, around 60 kya, an array of innovative practices appeared in the Middle Palaeolithic of Europe. The pattern is fragmented rather than unified in that there is no specific new technique or strategy that can be shown to have been practised throughout Middle Palaeolithic Europe after that time. In fact, this hardly weakens the case in that it suggests that local or regional innovations, derived from bodies of social knowledge manifest on local or regional scales and directed to locally or regionally experienced environments, could become fixed in local and regional projects for living. Evidence from both the early Lower and the late Middle Palaeolithic therefore indicates that the scalar convergence of socially transmitted knowledge and knowledgeable action, of structure and agency, was not to

The Evolution of Individual Capacities Evolutionary developments in the biologies of hominid individuals, including their innate cognitive abilities, might have been responsible. The most obvious example, of course, is the emergence and elaboration of language, which has profound implications for both the spatiotemporal reach of knowledgeable action and for the

117

social transmission of knowledge. The capacity of individuals to communicate to each other information about the world permits decision making to be based upon experience beyond the direct experience of any single individual (Gamble 1986, 57; Moore 1981). The capacity to act on the basis not just of the here-and-now but also of other times and other places, already inherent in memory, is raised by an order of magnitude – and in a manner that is explicitly social and which presences the experience of others in the mind of the agent. This is true even if the primary resource afforded by language was the construction and negotiation of interpersonal social relationships (Aiello and Dunbar 1993), and information sharing was merely its incidental by-product. The syntactic and semantic elaboration of language might therefore be a critical component of the expanding reach of knowledgeable action. Whallon (1989) has suggested that the emergence of modern linguistic capacities might have been the enabling factor that underwrote the occupation of highly marginal landscapes such as Siberia in the Upper Palaeolithic, in that the possession of the facilities for categorical representation implicit in language, and for the expression of future contingencies, permitted the construction of networks of mutual rights and obligations between groups by means of which the risks attendant on the periodic occurrence of ‘bad years’ could be spread. Of course, we have no way of knowing the syntactical and semantic properties of Neanderthal and Upper Palaeolithic languages, but Whallon’s ideas do illustrate how language might be implicated in the enhanced predication of action on acts and eventualities perpetrated at other times and other places, including those perpetrated by other people. In treating propositional language and intelligence as adaptive devices for the execution of planning, Parker and Milbraith (1993) make a similar point.

technical act performed by another, in consequence of which they perform a similar act to achieve the same goal, imitative learning also involves the reproduction by the learner of the particular way or ways in which the tool is used by the ‘model’ individual. As Noble and Davidson put it, learning by emulation rather than imitation limits the scope for the diversification of skilled action (Noble and Davidson 1996, 109). Learning based upon true pedagogy and augmented by linguistic instruction whereby the principles of the technical act are described, abstracted and communicated, is likely to be the means by which specific knowledgeable skills are most effectively and faithfully transmitted. Learning through non-linguistic instruction, or through imitation or emulation in the absence of instruction, are progressively more likely to inhibit the social transmission of specific techniques, particularly innovative or unusual techniques that are only rarely observed. The result would be a more restricted social repertoire of knowledgeable skills, less prone to the institutionalisation of particular and innovative practices and therefore more persistent in time. The nub of the problem, of course, is the difficulty of detecting these factors in the Palaeolithic record. The chronology of the origin and elaboration of language remains hotly disputed, with poles represented by Hewes’ long chronology in which gestural language is as old as the genus Homo (Hewes 1973, 1993) and Davidson and Noble’s denial of language prior to the Upper Palaeolithic (Davidson and Noble 1993; Noble and Davidson 1996). Lieberman’s celebrated work on the anatomy of sound production in Neanderthals (Lieberman and Crelin 1971; Lieberman 1984) suggests that their speech must have been slower and more phonetically restricted than that of Homo sapiens, but has been challenged both on methodological grounds and by the Kebara Neanderthal hyoid (Arensburg et al 1989, 1990). With regard to the kinds of social learning processes implied by particular stone working techniques, virtually no work has been performed at all. All one can say is that an emulative mode of learning might be sufficient to account for the simple and repetitive gestures typical of Oldowan knapping, while it is difficult to imagine how the subtle skills involved in Levallois reduction (Schlanger 1996), and in the differences between its variants, could be transmitted without instruction; but these are mere impressions. Still, it is clear that changes through time in the ways that Palaeolithic hominids acquired technical skills from others might have been an important factor in increasing the likelihood that particular ways of acting, including innovative ways, would become fixed in the socially transmitted repertoire of knowledge.

Language is also a possible factor in teaching and learning, which are clearly critical elements in the transmission of knowledge between individuals. Gibson argues that cultural traditions – for the present purposes, socially transmitted knowledge - are predicated upon ‘imitation combined with active teaching’ (Gibson 1993, 132), and cites Boesch’s observation of a chimp mother in the Tai forest demonstrating to her child the proper orientation of a nutcracking hammer, and the child’s imitation of its mother’s demonstration (Boesch 1993), as evidence for the existence of true pedagogy – an interactive relationship between teacher and learner - in primate species other than humans. She also refers to the imitative capacities of chimps exposed to gestural or vocal language by researchers, and points out that behaviours can be acquired by imitation in the absence of instruction in many species, including songbirds. On the other hand, Tomasello (1990; Tomasello et al 1993) and Visalberghi (1993) both doubt the existence of imitation in any living non-human primate, at least as far as the learning of tool making and using skills are concerned, preferring to speak instead of emulation. Whereas emulation involves the observation by one individual of a

Since learning and socialisation are mainly, though not exclusively, features of childhood it is also reasonable to think that an extension in the length of childhood might increase the scope and depth of the individuals’s opportunities to learn. A deceleration in hominid growth rates is certainly documented by the counting of

118

perikymata increments on fossil juvenile teeth. Studies have shown that Australopithecus afarensis and A. africanus both matured at twice the rate of modern children (Stringer and Gamble 1993, 85), while the Nariokotome Homo ergaster boy, dated to around 1.6 Mya, was only about seven years old at death despite being 1.63m tall and displaying a state of tooth eruption equivalent to that of a modern 12 year old. Similar analysis of the upper central incisor of the Gibraltar Neanderthal child, dated to around 50 kya, suggests that it was between three and four years old at death, although the state of tooth eruption is equivalent to that of a modern five year old. At the same time, the temporal bone is at a similar stage of development to that of a modern three year old (Stringer et al 1990), so the acceleration in growth rate by comparison with modern children is rather less marked than in the much earlier African specimens. This sample is far too small to draw general conclusions, but it does support the notion that maturation rates declined, and childhood lengthened, in the course of the Palaeolithic. The period of childhood dependency on the mother, the primary ‘model’ individual from which children learn, might also have been extended by an increase in the birth spacing interval, so that a child could remain in close personal contact with the mother for longer before being displaced by an infant sibling.

chimpanzees (Noble and Davidson 1996, 109). The expression of apparently innate propensities is therefore revealed as dependent upon social practices of child rearing. Space-Time Properties of Social Networks

Note should also be made of Mithen’s ‘modular mind’ model for the evolution of human consciousness (Mithen 1996). This also deals with the increasing integration of difference through the Palaeolithic, although it envisages an evolutionary process of dissolution of the hard-wired psycho-neural boundaries between modes of intelligence such as social, technical and linguistic, leading to a progressively enhanced capacity for ‘lateral thinking’. Mithen has little to say in this model about action in time and space or about social learning, but it is another example of how the incorporation of difference might be a consequence of evolutionary developments in the innate capacities of individuals.

A second category of explanation for the scalar convergence of socially transmitted knowledge and knowledgeable action emphasises the space-time properties of the social networks along which knowledge is transmitted and in the context of which individual action takes place. The study of the characters and properties of social networks and their relevance to palaeolithic archaeology undertaken by Gamble (1999, 42-62) and referred to in Chapter 1 is a useful starting point here. His theme is that, in the Lower and Middle Palaeolithic, social relations were enacted in the context of hierarchically lower orders of small social network, the intimate and the effective, which are constructed and serviced by emotional and material resources respectively without recourse to the representation of persons through material culture symbols (Douglas 1973). Symbolism only becomes an important resource in the Upper Palaeolithic when networks are constructed at the extended level, and relations with relatively large numbers of rarely-encountered individuals are constructed and maintained through the personification of objects and their incorporation into the social realm as things that speak. This permits personhood to be presenced and expressed beyond the contexts of personal proximity upon which intimate and effective relations depend, ‘stretching’ social action in time and space (Gamble 1998). The extent to which Middle Palaeolithic leaf points might reasonably be thought of as symbolic resources is discussed below. The main point to stress here is that Gamble’s views are closely comparable to those expressed here, particularly with respect to the increasing reach of knowledgeable action in time and space.

It should be stressed that these accounts all derive developments in human behavioural flexibility from evolutionary developments in the capacities of individuals, and therefore do not abolish the mutuality between scales of knowledge. They emphasise the dynamic between the behaviour of the individual and the biological nature of species which, as was discussed in Chapter 2, itself represents a store of knowledge and a transmissible project for living. What these models do ignore is the extent to which socially mediated behaviour defines the terms upon which evolution operates. The danger of this is set out by Noble and Davidson, who point out that young chimps reared by human carers who establish human familiar relations with them show a significantly greater propensity for paying shared attention to objects and for imitative learning than young chimps in the wild, and that this is a consequence of the lack of adult attention received by young wild

However, certain criticisms must be raised that are germain to the question of the mechanisms of the social reproduction of knowledge. Firstly, it is inconceivable that sociality in the Lower and Middle Palaeolithic was lived exclusively at the intimate and effective levels. Groups of 20 individuals are not reproductively viable in perpetuity, as Gamble had previously recognised (Gamble 1986, 61). Some kind of immersion of individuals in extended networks must have been a fact of life in the Lower and Middle Palaeolithic, even if limited to occasional chance encounters and the movement of individuals of one or other sex to neighbouring bands at puberty. This raises a crucial point: the extent to which technical knowledge passed not only down the generations, but also between groups. Gamble cites work by Boissevain (1974) and by Turner and Maryanski (1991) which suggests that high density social networks – i.e. networks in which a high proportion of the

119

interpersonal relationships that could exist do in fact exist – promote the standardisation of values and practices, whereas low density networks underwrite high levels of individualism. He also refers to Steele’s simulation of the conductivity of social networks of varying density, which suggests that innovations are more easily transmitted in low density networks (Steele 1994). Since intimate and effective networks are high density in character given the frequent interaction of their members in multiple contexts, while extended networks are of lower density, one has an explanation for the remarkable persistence of technical practices in the Lower and Middle Palaeolithic. Technical action was transmitted and acquired in high density intimate contexts in which the multivalent flows of information between closely interdependent individuals created a resistance to individual innovation and to the successful transmission of any such innovative practices to other members of the network, while the ease by which the practices typical of one group were transmitted to neighbouring groups in the low density extended network ensured that this conservatism was integrated and maintained at a regional scale.

holons at any hierarchical level is therefore much slower than the rate of interaction within them. The example was given in Chapter 2 of the trout and the mole, which share a common scale of life cycle and might live in close spatial proximity, but which are components of distinct ecosystem holons – a stream holon and a terrestrial undersoil holon – because the rate of interactions between the two is low, being manifest only at the hierarchically higher, lower process rate level of the drainage basin ecosystem. Intimate and extended networks in the Lower Palaeolithic can be thought of in this way. The former networks constitute holons defined by the high rate of interaction of their component individuals. But interactions between them, including knowledge transfers, operate very much more slowly in conditions of very low population density. It is certainly the case that there are many more Middle than Lower Palaeolithic sites known in Europe (Gamble 1999, 174). The extent to which this represents differential survival of archaeological traces, particularly in view of the strong tendency for sites older than 200 kya to occur in open air contexts, rather than a real population increase, is uncertain. Equally, the presence of Middle Palaeolithic peoples in central and eastern Europe, from which occupation was virtually absent before 200 kya, complicates any simple extrapolation from numbers of sites to population densities. Nevertheless, a significant increase in population densities after 200 kya, and certainly after 300 kya, is, at the very least, consistent with the archaeological facts. A consequent intensification of social relations in extended networks as encounters between individuals from different intimate networks became a more regular feature of life would therefore have increased the potential for the transmission of knowledge on regional scales.

However, it is possible to mount a contrary argument. A technique practised by a high status individual might be emulated or imitated by others in his or her intimate and effective networks. Technical practices might respond to episodes of environmental transformation that, for example, change the character of stone raw material to which individuals have access, and then become routinised. The inertial resistance of knowledge transmission through intimate networks to innovation would then permit these new practices, once fixed, to be maintained in the social repertoire of technical knowledge thereafter. In short, the contention that the generation of innovative practices and their subsequent institutionalisation in socially transmitted knowledge are excluded by life lived in intimate and effective social networks is open to serious challenge. But if the generation of new technical practices of one kind or another did sometimes occur in intimate social contexts, as it surely must have, why did these practices not spread rapidly through the low density regional extended networks, as the Lower Palaeolithic record of technical stasis shows?

At the same time, the Middle Palaeolithic incorporation of difference and occupation of landscapes more spatially and temporally heterogeneous than those occupied before 200 kya, insofar as it implies the integration of the acts of different individuals or groups of individuals carried out in different parts of the landscape, might also have entailed an increase in the amount of time spent and acts performed in the company of individuals from the same effective network but different intimate networks. This shift from the performance of practical acts in high density intimate to somewhat lower density effective network contexts might itself have loosened the shackles on the individual’s power to do things differently. Increases in population density might therefore have been sufficient to effect both an acceleration in the generation of different practices and an increased likelihood of their being socially transmitted through effective and extended social networks, thereby underwriting the regionalisation of lithic industries in the Middle Palaeolithic (Bosinski 1982; Mellars 1992) and their structured variability through ecological time. One might add that, if Richter (1997) is correct in placing the entire central and eastern

The problem here might lie in the idea that extended networks, as low density networks, readily transmit new or different practices. In situations of very low population density where extended networks are weakly constituted by virtue of the rarity of personal encounters at that level and the absence of symbolic support, the opportunities for mutual transfers of knowledgeable practices between intimate networks are very few. Koestler’s concept of the holon, discussed in Chapter 2, aids an understanding of this. The holon is a scalar entity whose integrity is defined by a process rate boundary in that its components interact at a particular rate, and which integrates its components’ lower-rate interactions with other holons and their components. The rate of interaction between

120

European Micoquian in the period between 60 and 45 kya, then a very significant increase in the Middle Palaeolithic population of Europe might have occurred at around 60 kya. If so, that would, in the terms of this discussion, be consistent with the marked expansion in the reach of knowledgeable action and the sensitivity of socially transmitted knowledge to transformation represented by the Middle Palaeolithic leaf point phenomenon.

sub-ecological timescales and landscape to regional spatial scales. If Middle Palaeolithic populations were less prone to local extinction than their Lower Palaeolithic ancestors then, together with the effects of increased population discussed above, the likelihood of different practices persisting locally or regionally long enough to be visible archaeologically would be enhanced. A possible example of just such a local persistence in specific knapping practices is the use of longitudinal resharpening techniques in the Middle Palaeolithic assemblages of La Cotte de St Brelade, Jersey. Characteristic longitudinal sharpening flakes are found in the lowest archaeological layers H-D in low numbers, increase in frequency through layers C and B before reaching a peak in layer A-3 and then declining in layer 5 (Cornford 1986). Although the technique seems to have been an economising response to declining flint supplies, it is highly idiosyncratic and is known elsewhere only at a very few sites. The timespan represented by these layers is not known with certainty, but it seems to encompass a major cold-warm oscillation (van Vliet-Lanöe 1986) and a series of oscillations in local vegetation (Jones 1986) and therefore must incorporate a time span of several millennia at least. Layers C and D are TL dated to 238±35 kya, placing the industry sometime in the early Middle Palaeolithic.

One problem with this argument is apparent. If new practices could and did arise from social life lived in the intimate networks of the Lower Palaeolithic, but encountered resistance to their transmission to other such social networks in the wider region, why then does the Lower Palaeolithic record not consist of an array of highly particular and localised lithic industries? To answer this question one needs to consider colonisation and migration. These have generally been approached only from the perspective of gross movements of populations into previously unoccupied regions as a trend over the entire Palaeolithic (e.g. Gamble 1993). However, these processes also operate on much smaller spatiotemporal scales. Metapopulation ecology is a branch of ecology that deals with the dynamics that connect local populations of a species with the wider regional and continental population of populations, or metapopulation. It is now established that the typical course of population history in animal species consists of repeated local population extinctions followed by recolonisation of the vacated habitat by individuals derived from neighbouring populations (Hanski, 1999). The regional metapopulation is therefore maintained while local populations are much more ephemeral. The principle here is therefore very close to that of the steady state shifting mosaic (Bormann and Likens, 1979a,b) discussed in Chapters 2 and 3. The pattern of human occupation in Pleistocene Europe is likely to have conformed to this model. No local social group or network in which particular knowledgeable practices were perpetrated and transmitted through time can be expected to have persisted for long, at least in terms of Pleistocene timespans. One can imagine that, 400,000 years ago, three consecutive bad winters might have been sufficient to kill every human being in Kent, which would be recolonised by people from Essex or Sussex 10 years later. Any stone-working practices peculiar to specific, reproducing, local social entities in Kent would disappear with their extinction. What would persist and be visible in the archaeological record is the general body of practices shared throughout the extended social network, which can for these purposes be equated with the metapopulation. One can only speculate as to the frequency of such extinction:recolonisation events, but if they occurred once every 200 years it would be sufficient to negate any local trend towards directed change in lithic technical practices. In this case, socially transmitted knowledge is rendered resistant to transformation in the face of knowledgeable action through the dynamics of population history operative on super-biographical to

One can cite systemic factors other than population size and local extinction that might impact upon the relation between knowledgeable action and socially transmitted knowledge. It is conceivable, for example, that stone tool production and use was, at certain times and places, preferentially practised by one sex. If the other sex left their group of birth at puberty in a system of or analagous to exogamous marriage, then little or no knowledge of knapping techniques would be transmitted within the wider extended network or metapopulation. This would contribute to the isolation of knowledgeable action from socially transmitted knowledge at scales above the local. Mithen's model of Lower Palaeolithic social learning and group size also deserves consideration (Mithen 1994). His proposed correlation between handaxes and steppic environments cannot be accepted; however, the suggestion that the spatial character of the landscape impacted upon hominid aggregation and dispersal patterns, and therefore on the degree of accessibility to the young of model individuals from which they could acquire technical skills, draws attention to precisely the kind of question that a developed ecological geography of the Palaeolithic would need to address. Summary These possible factors behind the changing relationship between scales of knowledge in the Lower and Middle Palaeolithic are precisely that: possibilities. No special claim is made here for any particular factor as the key to the problem. They are not meant to be taken as in any

121

and the on-the-spot production of sharp-edged flakes to cut them, for example, are all acts that unfold over minutes, even seconds. The fact that intimate relationships between two particular individuals can last a lifetime is not the point. After all, the lifelong nature of an individual’s ecological relations with the environment of affordances does not imply, as we have seen, that knowledgeable acts performed in it possess a temporal reach measured in decades. Sociality lived at the intimate level is therefore typified by personal acts of very short duration, but frequently repeated; hierarchically speaking, it is a high process rate, high frequency scale domain.

way mutually exclusive. The Palaeolithic covers a very long time span over a very wide area, and it is to be expected that different combinations of these processes and mechanisms operated at different times and places. Indeed, the real nub of the question might be synergies that emerged from the relations between them. It has already been pointed out that the biological and cognitive-psychological aspects of knowledgeability were necessarily connected by relations of mutual cause and effect, and the ecological and social factors in the acquisition and transmission of knowledge should be added to this web of mutuality. Of course, the mutual relations between life processes operating on different scales, and the need to develop specific understandings of how the linkages between these scale domains play out in time and space, are the central concerns of this study. This consideration of the general principles behind some conceivable linkages is intended to illustrate a hierarchical approach to an ecological geography of the Palaeolithic, and to point the way to possible avenues of future research, rather than to provide answers. One conclusion, however, can be drawn – it is both unnecessary and insufficient to make simplistic connections between Pleistocene behaviour and the biologically-given cognitive powers of individuals. By the same token, developments in behaviour through the Palaeolithic entailed much more than an inexorable evolutionary increase in individual cognitive capacities brought about by a single universal process, natural selection. Insofar as evolutionary biological trends in cognition were implicated in these developments, they flowed from and contributed to matrices of connection between multiple processes on multiple scales that render essentialist dualisms, such as mind and body, knowledge and action, culture and nature, the individual and society and the individual and the species, redundant.

The distinction between intimate and effective networks is necessarily fuzzy. Polygamous mating practices might prevent the crystallisation of intimacy around small nuclear families; the death of an adult might compel the ‘spouse’ or ‘spouses’ to attach themselves to another from within their extended network, thereby fusing intimate networks; these and other events and flows in practical life conspire to blur the boundaries between intimate and effective networks in the small band societies presumably typical of the Lower and Middle Palaeolithic. Nevertheless, some differences in the temporalities of the acts involved in the maintenance of the two orders of network can be supposed. The practical, often cooperative, activities of daily life - hide scraping, food gathering, hunting forays, the dismemberment and transport of carcasses, for example – provide the primary contexts for social relationships at the effective level, although they might also be performed in the presence of former, current or future intimates. They are likely to involve the curation and possibly transport of stone raw materials, the systematic production of blanks as supports for tools and the rejuvenation through retouch of tools used repeatedly to the point of exhaustion. The duration of acts of this nature extends from minutes through hours to perhaps a few days, i.e. overlapping with the duration of acts directed towards intimates but exceeding it by at least an order of magnitude at the upper end of the range. Being practical in nature, these acts are performed frequently, but some particular practices - hide scraping or large game procurement, for example – need not necessarily be performed every day. If individuals preferentially carried out certain tasks in the company of certain others, then the frequency with which material acts affirming effective relationships were performed would be somewhat lower than, but overlapping with, those servicing intimate relations. Effective social networks therefore represent a lower process rate, lower frequency scale domain than intimate networks, but the scalar distance between them is not great, permitting high levels of mutual interaction.

Scale, Society and Signification Social Relations as Scale Domains Both knowledgeable action and socially transmitted knowledge can be understood only in terms of sociality in that the latter refers specifically to a collective, transindividual scale domain and the former deals with directed individual acts whose contexts necessarily include other individuals. From the standpoint of hierarchy theory, this raises the question of the scalar character of social relations and their linkages with scales of knowledge. This requires some assessment of the temporalities of the knowledgeable acts and practices through which Gamble’s orders of sociality are created, maintained and challenged.

It is more difficult to gain an insight into the temporalities of extended networks in the Lower and Middle Palaeolithic, especially in situations where population densities were very low and encounters between strangers rare. Relations of this nature might not have been actively serviced at all, calling into question the extent to which

At the level of the intimate network, emotional acts of familiarity are fundamental. The embrace, the gossip, the sharing of the roasted root vegetable or portion of meat

122

serviced, their low process rates and possible persistence on ecological and sub-ecological timescales might indeed have located them within the dynamic scale range of socially transmitted knowledge. The fundamental scalar boundary here is that between sociality at the extended level and knowledgeable action, a boundary which placed the processes of extended sociality outside the social grain responsiveness of knowledgeable agents and mitigated against the generation of practices directed specifically to that social scale domain. Knowledgeable ways of acting cannot become fixed if they are not practised.

they could be regarded as truly social and precluding the identification of any characteristic temporality of action. It is only when such encounters become sufficiently frequent to be recognised as significant features of practical life – as affordances in a social landscape - that their servicing becomes a practical task. One can only speculate as to the kinds of practices that might have been specifically directed to the servicing of minimal extended relations, although there is no reason to imagine that they necessarily involved acts enduring for days or weeks, such as the temporary fusion of groups, so their temporalities might have been little different from those of effective or even intimate networks. Perhaps the key here is not the duration of the servicing acts but their relatively low frequency of performance. This factor would constitute a process rate boundary between intimate and effective social networks on the one hand and extended networks on the other whereby the transmission of knowledge from the former to the latter would be heavily damped.

The scales of knowledge available for appropriation as resources in the construction of sociality therefore depends upon the scalar distance between the realms of sociality and knowledge in question. ‘Symbolic’ artefacts that signify social personhood are social entities whose production and interpretation are dependent on a mutuality between structured, received ways of doing and seeing on the one hand, and socially directed knowledgeable acts, including acts of interpretation, on the other. Without this mutuality, artefacts as forms and as materialisations of social choices and orientations simply do not exist as resources for knowledgeable social action. Material symbols and signification are absent from the Lower Palaeolithic precisely because no realm of social action operated in a scale domain hierarchically positioned to unify knowledgeable action and socially transmitted technical knowledge.

Scales of Knowledge and the Construction of Palaeolithic Social Relations The Lower Palaeolithic How, then, do the tempos of life in the different orders of network stand in scalar relation to knowledgeable action and socially transmitted knowledge? Intimate and effective networks are maintained by knowledgeable acts and are therefore, in scalar terms, coextensive with knowledgeable action; there is no hierarchical distance or rate process boundary between them. They form a holonic unity, and utilitarian acts are indistinguishable from social acts. On the other hand, socially transmitted knowledge in the Lower Palaeolithic was separated by a scalar chasm from the processes of intimate and effective sociality. Their capacity for mutual transformation cannot, from a hierarchical point of view, have been any greater than that documented in this study between structure and agency in the realm of lithic-technical knowledge. Ways of fabricating stone artefacts – technique and form – were simply not available as social resources at intimate and effective levels in the Lower Palaeolithic because the insensitivity of socially transmitted knowledge to transformation through its creative application by knowledgeable agents precluded the manipulation of technique and form on social time scales. Social action and received technical knowledge lay outside each other’s dynamic scale range. Social persons were in no way constructed through artefact form or style – i.e. material culture – but through the acts to which stone artefacts were directed and in relation to which the artefacts stood only as physical-instrumental extensions of the purposeful hands of their users.

The Middle Palaeolithic, 200–60 kya The Middle Palaeolithic before 60 kya seems little different from the Lower Palaeolithic from this point of view. Lithic industrial practices might show structured change on Milankovich timescales while the temporal reach of knowledgeable action might have expanded beyond that of the Lower Palaeolithic, but it remains difficult to see any evidence for the appropriation of artefact form and fabrication technique as a resource in the construction of social relations. The convergence of socially transmitted knowledge and knowledgeable action after 200 kya does not seem to have been sufficient to effect a dynamic of mutuality between them on social timescales. One point should be made, however. Insofar as the expansion in the temporalities of agency in this period implies an increased reliance on the activities of small sub-groups operating in different parts of the landscape, it also implies a shift relative to the Lower Palaeolithic from activities performed in intimate social contexts to the performance of acts in the contexts of effective networks. The distinction between the intimate and effective orders of sociality acquires greater resonance after 200 kya. The well-documented tendency for raw materials transported over longer distances to occur in Middle Palaeolithic assemblages largely or exclusively as retouched tools, especially scrapers, or as Levallois blanks (Geneste 1988, 1989a,b), is of interest here in that it might be a specific trace of social action in

Very low population densities might, as discussed earlier, have placed extended social networks outside the realm of social knowledge. Where they were practically

123

action performed in effective network contexts. One can speculate that this was heightened yet further by rapid afforestation in that it renders the landscape relatively unknown and therefore unpredictable, one feasible response to which is an increase in foraging activity by task-specific groups in the landscape periphery. Insofar as intimate networks unified children, adults, men and women in relations of emotional familiarity, but effective networks were created and maintained by individuals engaged in particular, relatively long-duration tasks or sets of tasks, then social centres like the Sesselfelsgrotte can be seen as the spatial focus of intimacy, and the distinction between centre and periphery as a distinction between spheres of sociality. The use of leaf points only in effective rather than intimate social settings therefore takes on crucial significance. There is no reason to think that, for whatever task or tasks they were used, Middle Palaeolithic leaf points were functionally superior to Mousterian points or unretouched Levallois points. In fact, the application of flat surface retouch, even when made on flake blanks, makes the Mauern leaf points expensive in a narrow time:energy optimisation sense. Neither can leaf points be explained away as merely incidental outcomes of raw material constraints, economisation and reduction strategies. The Mauern leaf points show no morphological continuity with other bifaces or flakes tools present, and are themselves made variously on Plattensilex or flakes. The same pattern of leaf point formal integrity and distinction from other artefacts, combined with technological variability between leaf points, has already been shown to occur at Rörshain and Musselievo, and in Middle Palaeolithic leaf point assemblages more generally. This can be contrasted with the earlier Middle Palaeolithic association between distance from raw material source and ‘finished’ tools, which represent the technological end points of chaînes opératoires, the earlier stages of which were discarded between the source and the site, rather than morphologically discrete ‘target’ forms.

effective networks. Interestingly, although the distances over which raw material transfers were effected increases relative to the Lower Palaeolithic, implying an intensification of movement and social contacts on regional scales, there is still nothing in the archaeology of this period that appears emblematic of intimate and effective networks within wider extended networks. Leaf Points and the Middle Palaeolithic after 60 kya The key features of the Middle Palaeolithic leaf point phenomenon identified in this investigation – the sensitivity of stone tool fabrication practices to rapid landscape transformations, the systematic opposition of action at the social centre to action in the peripheries of the landscape and the exclusive or nearly exclusive association of leaf points as a specific form with particular places in the landscape periphery – have enormous implications for the construction and negotiation of social life through knowledgeable action. The location of Mauern and the Obernederhöhle in dry valleys, as discussed earlier, indicates a close relationship between the knowledgeable identification of places affording particular opportunities and the performance of specific technical acts at those places. Mauern is particularly instructive. The Wellheimer Trockental would have afforded predictable, seasonal access to migrating ungulates, even in episodes when the landscape was rapidly becoming more wooded, perhaps over just a few decades, and established routines and rhythms of movement and action were coming under challenge. The concentration of leaf points in the Weinberghöhlen, especially in Layers F/F1, is best interpreted in terms of the occasional use of the site for gearing up and as a leaf point cache or store, a provisioning of a place (Kuhn 1995) in a mode of action in the landscape that can be described as logistical (Binford 1980). The absence of hearths or food refuse, the carnivore-generated fauna and the relatively high frequency of cores are all consistent with this. But occasional does not mean unstructured. The caching of leaf points implies the anticipation of future return. The archaeology of Mauern is a record of specific action at a specific place through which multiple scales of knowledgeability were integrated. The journey to the place; the transport of materials; the relation with centres of social action like the Sesselfelsgrotte, the individuals there and the activities in which they were engaged; the provision for future return; all indicate action mediated through knowledgeability on multiple and extensive temporal, spatial and social scales. The temporality that unites the Middle Palaeolithic leaf point phenomenon, however, is the rapid afforestation episodes typical of the OIS 3 regions in which it is represented. In the Altmühl Valley, this is manifest as an awareness of the affordances of dry valleys, and of their persistence through changes in the understood landscape on time scales of one or two individual lifespans.

It is therefore proposed that leaf points make sense only as the defining artefact of action performed in the effective social sphere, the objectification of the solidarity of ‘taskmates’ as opposed to intimates. The Mauern leaf points represent the dissolution of the scalar boundaries between creative knowledgeable action, the servicing of effective social network relations and received bodies of knapping and landscape knowledge; that is, the Mauern leaf points were a symbolic resource in the construction of social life. Presence and absence, centre and periphery, now and then are all embodied in the leaf point form. Their caching in the Wellheimer Trockental presenced the effective network and its component individuals even in the long periods when they were absent, permitting the escape from proximity of which Gamble has written (Gamble 1998). It made the Weinberghöhlen a true place, imbued with social significance and a theatre for the performance of social acts on which the production and reproduction of social being was predicated. The Altmühl Valley leaf points

The Middle Palaeolithic of the Altmühl Valley in OIS 3 therefore represents an intensification of the shift to

124

can, in a certain sense, be thought of as expressions of Wiessner’s emblemic style through which group identity is proclaimed in material objects (Wiessner 1983), except that they referred to within-group, rather than betweengroup, social identity.

practices and scales of heterogeneity in the material world, and enabled the creation of things-as-persons and their appropriation into social life.

From this viewpoint, the use of symbolic resources in the construction of society begins in Europe not with the Upper Palaeolithic, but with the late Middle Palaeolithic. It begins not in the context of the creation of extended social networks between relative strangers, but in the context of the solidification of social relations within effective networks. The apparent absence of leaf points or any other artefacts that might qualify as symbolic at centres of social intimacy is therefore instructive. It suggests that the leaf point makers, either through the resistance of high density networks to innovation or because the short temporalities of intimacy remained outside the dynamic scale range of socially transmitted knowledge, did not bring symbolic resources to bear on interpersonal relationships when immersed in intimate contexts. This is consistent with the paucity of evidence for the symbolic construction of intimate space, such as built hearths and enclosed living space, even in the late Middle Palaeolithic – although the Molodova huts and Vilas Ruivas hearth mentioned earlier suggests that this barrier was not absolutely impermeable. Caution must therefore be exercised in characterising centres of social occupation and action like the Sesselfelsgrotte as centres of intimacy; the intimate and effective social spheres might have been more differentiated than before 60 kya, but must still have overlapped. A wide range of tasks, such as hide scraping and food preparation, that do not necessarily imply the copresence of the closest intimates were certainly carried out at such sites. The penetration of social relationships even at the effective level seems to have been incomplete, although one could speculate as to whether Micoquian bifaces might have been immersed in some particular way with the creation and servicing of effective networks at social centres as opposed to the landscape periphery, and that the association of leaf points with social centres in the Crimean Middle Palaeolithic might reflect the same process.

Scales of Knowledge and Human ‘Modernity’

It is clear that, technical questions such as the shift from flake–based to blade-based stone tool technology notwithstanding, the key feature that differentiates the Upper from the preceding Lower and Middle Palaeolithic for archaeologists is the evidence it furnishes for life lived through symbols and signification (e.g. Mellars 1996; Klein 1999; Leroi-Gourhan 1993; Davidson and Noble 1993; Noble and Davidson 1996; papers in Mellars and Stringer 1989). This is also true of archaeologists who doubt or dispute a recent African origin for all modern people (e.g. Binford 1989), and for those who argue that European Neanderthals attained true modernity quite independently of any immigration by Africans at or around 40 kya (e.g. d’Errico et al 1998; Zilhão and d’Errico 1999a,b). The proliferation in the Aurignacian of tool types, in both stone and hard organic materials, that exhibit high levels of arbitrary and standardised form; the widespread fabrication of objects of personal adornment, and of art objects; the construction of domestic space; the occupation of Siberia and Australia; all are referred to a capacity for representation through signs and symbols that archaic humans lacked, and it is this capacity that is regarded as the hallmark of modernity, of ourselves.

This interpretation of the social significance of Middle Palaeolithic leaf points must be placed within limits. Network theory, and the orders of sociality it proposes, offer a perspective and suggestions for analysis, not ready answers. Neither is it claimed here that Middle Palaeolithic life in Europe after 60 kya was experienced only and entirely through the prism of signification and symbolism, or that the construction of social life in effective network contexts in part through the attribution of meaning to artefacts was universal in that period. But despite these caveats, a general conclusion can be advanced with some confidence: that the hierarchical proximity of the scale domains of socially transmitted knowledge and knowledgeable action in the Middle Palaeolithic after 60 kya underwrote the emergence of new orders of dynamic relations between technical

Although evidence has been forwarded here for the emergence of material culture signification in the late Middle Palaeolithic, the qualitative leap represented by the appearance of the Upper Palaeolithic is not to be denied. The central point here is that this ‘symbolic explosion’ need not be explained by reference to evolutionary biological developments alone; that is, one need not necessarily posit innate abilities such as language and a capacity for representational thought that modern humans possess but archaic humans lacked. Rather, the question is one of the scale domains of knowledge and the nature of the processes that linked them. Upper Palaeolithic ‘jewellery’ is especially instructive in this regard. The presence in Aurignacian sites in southwestern France of both Atlantic and Mediterranean marine shells (Taborin 1990, 1993), and of

This study has quite deliberately avoided the temptation to address the ecological geography of the Upper Palaeolithic. This orientation flows in part from McBurney’s in The Geographical Study, but also from a conviction that any consideration of the Upper Palaeolithic might be sidetracked and confused by an involvement in the problem of the origins of Modern Humans. Some broad observations of the relevance of this investigation for our understanding of human modernity are nevertheless in order at this point.

125

Black Sea shells at Kostenki, southern Russia (Hahn 1977), sometimes indicating transport of these items over distances of 500 km or more, is clear evidence for the movement of things through extended social networks. The high fraction of Aurignacian lithic assemblages made on material derived from over 30km away makes the same point. But the most important implication of jewellery is for the construction of the individual as a social person through material culture. Shell, tooth, bone, ivory and stone beads all bespeak the proclamation of the individual within society. This is the realm of material culture signification termed assertive style by Wiessner (1983). But the materials are transported on regional scales; material culture signification pervades all orders of social life from the individual to the extended. A scalar proximity of socially transmitted knowledge and knowledgeable action is also indicated by the systematic occupation of highly marginal environments, and by a clear correlation between lithic industries and climatic events (e.g. Solutrean-glacial maximum, Magdalenianlate glacial, microliths-terminal glacial/early post glacial). We therefore see the incorporation of person, knowledge, society and the world into the undivided Maussian total social fact (Mauss 1967 [1925]), an aspect of a holonic entity whose dynamics are predicated on received and applied orders of knowledgeability operating at the same or closely linked scale domains. From this point of view, human modernity is not so much a biological state as a system scalar dynamic.

Epilogue It is clear from this study that human ‘modernity’ is not a unidimensional, essential property of Homo sapiens alone, nor is it a basis for erecting a Chinese wall separating two great evolutionary blocs of humanity, the archaic-natural and the modern-cultural. In accepting such a destructive dualism one might as well begin excavating for the obelisk from 2001: A Space Odyssey. Modernity should instead be understood as a region in a continuum of states of knowledgeable sensitivity to and incorporation of difference in the world, a region in which received and applied aspects of knowledge exist in a relation of close mutuality. But these scalar aspects of knowledge had been converging upon each other throughout the Palaeolithic, even to the extent of underwriting the construction of some aspects of social life through material culture signification in the late Middle Palaeolithic. The history of human occupation in Europe before 40 kya is therefore not a story of stasis, but of a trajectory of development continuous with the condition we call ‘modernity’. This is all a long way from McBurney’s questions in The Geographical Study. In moving from fundamental but simple questions about the linkages between topography, annual climate patterns and stone tools to the ecological geography developed here, the emphasis on scale has been the key that opens the door to a unification of ecological and social aspects of past knowledgeable human immersion in the world. Through this unification, wherever and whenever we look in the Palaeolithic past we can see socially-structured and society-structuring agents constituting themselves through knowledgeable action in the world – that is, through ecology.

126

BIBLIOGRAPHY

Absolon, K. 1935. Ottaslawitz, eine neue paläolithische Station in Mähren mit Quartzit Aurignacien. Versuch einer systhematisch-typologischen Bestimmung der Artefakte. Brünn, Polygrafia. Absolon, K. 1947. Palaeolethnologická stavba velikých středoevropských Diluviálních stanic. Příroda 40, 3-7, 25-29. Adam, D.P. and West, G.J. 1983. Temperature and Precipitation Estimates through the Last Glacial Cycle from Clear Lake, California, Pollen Data. Science 219, 168-70. Adamenko, O.M. and Gladilin, V.N. 1990. Korolevo – Un des plus anciens habitats acheuléens-moustériens en Transcarpatie soviétique. L’Anthropologie 94, 689-713. Adams, B. 1998. The Middle to Upper Paleolithic Transition in Central Europe: The record from the Bükk Mountain region. British Archaeological Reports International Series 693. BAR Publishing, Oxford. Aiello, L. and Dunbar, R. 1993. Neocortex size, group size and the evolution of language. Current Anthropology 34, 184-93. Aldhouse-Green, and Pettitt,from P., 1998. Paviland Cave: contextualizing ‘RedRM-110, Lady’. Antiquity 72, 756-72. some workingS.principles hierarchy theory. General Technical the Report United States Department of Agriculture, Rocky Mountain Forest and Range Experiment Station, Fort Collins, Colorado. Allen, T.F.H., O’Neill, R.V. and Hoekstra, T.W. 1984. Interlevel relations in ecological research and management: Alley, R.B. and MacAyeal, D.R. 1994. Ice-rafted debris associated with binge-purge oscillations of the Laurentide Ice Sheet. Paleoceanography 9, 503-11. Allsworth-Jones, P. 1986. The Szeletian and the Transition from Middle to Upper Palaeolithic in Central Europe. Oxford, Clarendon Press. Allsworth-Jones, P. 1990. The Szeletian and the Stratigraphic Succession in Central Europe and Adjacent Areas: Main Trends, Recent Results, and Problems for Resolution. In P. Mellars (Ed) The Emergence of Modern Humans; An Archaeological Perspective, pp 160-242. Edinburgh University Press, Edinburgh. van Andel, T.H. 1998. Paleosols, red sediments, and the Old Stone Age in Greece. Geoarchaeology 13, 361-90. van Andel, T.H. and Tzedakis, P.C. 1996. Palaeolithic Landscapes of Europe and Environs, 150,000-25,000 Years Ago: An Overview. Quaternary Science Reviews 15, 481-500. Andersen, S.T. 1961. Vegetation and its environment in Denmark in the Early Weichselian glacial (last glacial). Danmarks Geologiske Undersøgelse R II 75, 1-175. Anissutkine, N.K. 1990. Industries Moustériennes à Pointes Foliacées du Sud-Ouest de la Partie Européenne de l’URSS. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 31-41. ERAUL 42, Liège. Arensburg, B., Schepartz, L.A., Tillier, A.M. Vandermeersch, B. and Rak, Y. 1990. A reappraisal of the anatomical basis for speech in Middle Palaeolithic hominids. American Journal of Physical Anthropology 83, 137-46. Arensburg, B., Tillier, A.M., Vandermeersch, B., Duday, H., Schepartz, L.A. and Rak, Y. 1989. A Middle Palaeolithic human hyoid bone. Nature 388, 758-60. Ashton, N.M. and Cook, J. 1986. Dating and correlating the French Mousterian. Nature 324, 113.

127

Ashton, N.M. and McNabb, J. 1992. The High Lodge Flint Industries. In N.M. Ashton, J. Cook, S.G. Lewis and J. Rose (Eds) High Lodge: Excavations by G. de G. Sieveking 1962-68 and J. Cook 1988, pp 124-63. London, British Museum Press. Averdieck, F.-R. 1967. Die Vegetationsentwicklung des Eem-Interglazials und der Frühwürm-Interstadiale von Odderade/Schleswig-Holstein. Fundamenta B2; Frühe Menschheit und Umwelt 2, 101-25. Köln. Bailey, G.N. 1981. Concepts, time scales and explanations in economic prehistory. In A Sheridan and G Bailey (Eds) Economic Archaeology, pp 97-117. British Archaeological Reports International Series 9. Oxford, BAR Publishing. Bailey, G.N. (Ed) 1983a. Hunter-Gatherer Economy in Prehistory. Cambridge, Cambridge University Press. Bailey, G.N. 1983b. Hunter-gatherer behaviour in prehistory: problems and perspectives. In G.N. Bailey (Ed) HunterGatherer Economy in Prehistory, pp 2-4. Cambridge, Cambridge University Press. Bailey, G.N. 1983c. Concepts of time in Quaternary prehistory. Annual Review of Anthropology 12, 165-92. Bailey, G.N. 1983d. Economic change in late Pleistocene Cantabria. In G.N. Bailey (Ed) Hunter-Gatherer Economy in Prehistory, pp 96-110. Cambridge, Cambridge University Press. Bailey, G.N. 1997. Klithi: a synthesis. In G.N Bailey (Ed) Klithi: Palaeolithic Settlement and Quaternary Landscapes in Northwest Greece. Vol. 2: Klithi in its Local and Regional Setting, pp 655-77. Cambridge, McDonald Institute for Archaeological Research. Bailey, G.N. 1999. The Palaeolithic archaeology and palaeogeography of Epirus with particular reference to the investigations of the Klithi rockshelter. In G.N. Bailey, E. Adam, E. Panagopoulou, C. Perles and K. Zachos (Eds) The Palaeolithic Archaeology of Greece and Adjacent Areas: Proceedings of the ICOPAG Conference, Ioannina, pp 159-69. British School at Athens Studies 3. Bailey, G.N. 2007. Time perspectives, palimpsests and the archaeology of time. Journal of Anthropological Archaeology 26, 198-223. Bailey, G.N., Papaconstantinou, V. and Sturdy, D. 1992. Asprochaliko and Kokkinopilos: TL dating and reinterpretation of Middle Palaeoliothic sites in Epirus, North-West Greece. Cambridge Archaeological Journal 2, 136-44. Baker, R.R. 1978. The Evolutionary Ecology of Animal Migration. London, Hodder and Stoughton. Bánesz, L. 1960. K otázke listovitých hrotov z Vel’kého Šariša. Archeologické Rozhledy 12, 313-18. Bapty, I. and Yates, T. (Eds) 1990. Archaeology after Structuralism. London, Routledge. Bárta, J. 1960. K problému listovitých hrotov typu Moravany-Dlhá. Slovenská Archeológia 8, Part 2, 295-324. Bárta, J. 1967. Stratigraphische Übersicht der paläolithischen Funde in der Westslowakei. Quartär 18, 57-80. Bastin, B. 1992. Analyse pollinique des sédiments détritique, des coprolithes et des concrétions stalagmitique. In M. Otte (Ed) Recherches aux Grottes de Sclayn Vol. 1: Le Contexte, pp 59-78. Liège, ERAUL 27. Baumann, K.-H., Lackschewitz, K.S., Mangerud, J., Spielhagen, R.F., Wolf-Welling, T.C.W., Henrich, R. and Kassens, H. 1995. Reflection of Scandinavian Ice Sheet fluctuations in Norwegian Sea sediments during the last 150,000 years. Quaternary Research 43, 185-97. de Beaulieu, J.-L. and Reille, M. 1984. A long Upper Pleistocene pollen record from Les Echets, near Lyon, France. Boreas 13, 111-32. de Beaulieu, J.-L. and Reille, M. 1992a. The last climatic cycle at La Grande Pile (Vosges, France): A new pollen profile. Quaternary Science Reviews 11, 431-8.

128

de Beaulieu, J.-L. and Reille, M. 1992b. Long Pleistocene pollen sequences from the Velay Plateau (Massif Central, France). I. Ribains Maar. Vegetation History and Archaeobotany 1, 233-42. Behre, K.-E. 1974. Der Vegetation im Spätpleistozän von Osterwanna/Niedersachsen Makrorestuntersuchungen aus Eem- und Frühweichsel). Geologisches Jahrbuch 49, 3-48.

(Pollen-

und

Behre, K.-E. 1989. Biostratigraphy of the last glacial cycle in Europe. Quaternary Science Reviews 8, 25-44. Behre, K.-E. and Lade, U. 1986. Eine Folge von Eem und 4 Weichsel-Interstadialen in Oerel/Niedersachsen und ihr Vegetationsablauf. Eiszeitalter und Gegenwart 36, 11-36. Behre, K.-E. and van der Plicht, J. 1992. Towards an absolute chronology for the last glacial period in Europe: radiocarbon dates from Oerel, northern Germany. Vegetation History and Archaeobotany 1, 111-17. Bender, M., Sowers, T., Dickson, M.-L., Orchardo, J., Grootes, P., Mayewski, P.A. and Meese, D.A. 1994. Climate correlations between Greenland and Antarctica during the past 100,000 years. Nature 372, 663-6. Bennett, K.D., Tzedakis, P.C. and Willis, K.J. 1991. Quaternary refugia of north European trees. Journal of Biogeography 18, 103-15. Bernard, C. 1856. Introduction à l’étude de la médecine expérimentale. Paris, Garnier-Flammarion (1966). Binford, L.R. 1962. Archaeology as anthropology. American Antiquity 28, 217-25. Binford, L.R. 1965. Archaeological systematics and the study of culture process. American Antiquity 31, 203-10. Binford, L.R. 1968. Post-Pleistocene Adaptations. In S.R. Binford and L.R. Binford (Eds) New Perspectives in Archaeology. Chicago, Aldine. Binford, L.R. 1973. Interassemblage variability: the Mousterian and the ‘functional’ argument. In A.C. Renfrew (Ed) The Explanation of Culture Change, pp 227-54. London , Duckworth Binford, L.R. 1979. Organization and formation processes: Looking at curated technologies. Journal of Anthropological Research 35, 255-73. Binford, L.R. 1980. Willow smoke and dogs’ tails: Hunter-gatherer settlement and archaeological site formation. American Antiquity 45, 4-20. Binford, L.R. 1989. Isolating the transition to cultural adaptations; an organizational approach. In E. Trinkaus (Ed) The Emergence of Modern Humans; Biocultural adaptations in the Later Pleistocene, pp 18-41. Cambridge, Cambridge University Press. Binford, L.R. and Binford, S.R. 1966. A preliminary analysis of functional variability in the Mousterian of Levallois facies. American Anthropologist 68, 238-95. Bintliff, J. 1991. The contribution of an Annaliste/structural history approach to archaeology. In J. Bintliff (Ed) The Annales School and Archaeology, pp 1-33. Leicester, Leicester University Press. Birkner, F. 1916. Die Eiszeitliche Besiedlung des Schulerloches und des unteren Altmühltales. Abhandl.d. Köigl. Bayr. Akad. d. Wissenschaften, Math.-phys. Kl. XXVIII, 5. Abhdlg. Birkner, F. 1918. Bericht über die Höhlenforschung in Bayern im Jahre 1918. Jahrbuch der Bayerische Akademie der Wissenschaft 1918. Birkner, F. 1925. Das Altpaläolithikum in Bayern. Bayerische Vorgeschichtsfreund 5. Birkner, F. 1933. Madeleinewerkstätte bei Neuessing, BA Kelheim, Niederbayern. Bayerische Vorgeschichtsblatt 11S, 55. Bitiri, M. 1967. Paläolithische Blattspitzen in Rumänien. Quartär 18, 139-55.

129

Boesch, C. 1993. Aspects of transmission of tool-use in wild chimpanzees. In K.R. Gibson and T. Ingold (Eds) Tools, Language and Cognition in Human Evolution, pp 171-83. Cambridge, Cambridge University Press. Boesch, C. and Tomasello, M. 1998. Chimpanzee and Human Cultures. Current Anthropology 39, 591-614. Bogutskij, A.B., Sytnik, A.S. and Yamada, M. 1999. Nouveaux sites du Paléolithique inférieur en Ukraine occidentale. Bulletin de la Société Préhistorique Française 96, 71-6. Bohmers, A. 1939. Die Ausgrabungen in den Höhlen von Mauern. Germanien, 151-6. Bohmers, A. 1944. Die Mauerner Höhlen und ihre Bedeutung für die Einteilung der Altsteinzeit. Bericht über die Kieler Tagung 1939 der Forschungs-und Lehrgemeinschaft ‘Das Ahnenerbe’; 65-73. Bohmers, A. 1951. Die Höhlen von Mauern, Teil I: Kulturgeschichte der Altsteinzeitlichen Besiedlung. Groningen, Palaeohistoria 1. Boissevain, J. 1974. Network analysis: a reappraisal. Current Anthropology 20, 392-4. Bond, G., Broecker, W.D., Johnsen, S., McManus, J., Labeyrie, L., Jouzel, J. and Bonani, G. 1993. Correlations between climate records from North Atlantic sediments and Greenland ice. Nature 365, 143-7. Bond, G. and Lotti, R. 1995. Iceberg discharges into the North Atlantic on millennial timescales during the last glaciation. Science 267, 1005-10. Bond, G., Showers, W., Cheseby, M., Lotti, R., Almasi, P., Demenocal, P., Priore, P., Cullen, H., Hajdas, I. and Bonani, G. 1997. A pervasive millennial-scale cycle in North Atlantic Holocene and glacial climates. Science 269, 210-14. Bordes, F. 1950a. Principes d’une méthode d’étude des techniques de débitage et de la typologie du Paléolithique ancien et moyen. L’Anthropologie 54, 19-34. Bordes, F. 1950b. L’évolution buissonante des industries en Europe occidentale: considerations théorique sur le Paléolithique ancien et moyen. L’Anthropologie 54, 393-420. Bordes, F. 1953a. Essai de classification des industries Moustériennes. Bulletin de la Société Préhistorique Française 50, 457-66. Bordes, F. 1953b. Recherches sur les limons quaternaires du bassin de la Seine. Stratigraphie et Archéologie paléolithique. Archives de l’Institute de Paléontologie Humaine, Mémoire No.26. Bordes, F. 1968. The Old Stone Age. London, Weidenfeld and Nicholson. Bordes, F. 1984. Leçons sur le Paléolithique: Vol. 2, Le Paléolithique en Europe. Paris, Centre National de la Recherche Scientifique (Institute de Quaternaire, Université de Bordeaux I, Cahiers du Quaternaire No. 7. Bordes, F., Laville, H. and Paquereau, M.M. 1966. Observations sur le Pléistocène supérieur du gisement de CombeGrenal (Dordogne). Actes de la Société Linnéenne de Bordeaux 103, 3-19. Bormann, F.H. and Likens, G.E. 1979a. Pattern and Process in a Forested Ecosystem. New York, Springer-Verlag. Bormann, F.H. and Likens, G.E. 1979b. Catastrophic disturbance and the steady state in northern hardwood forests. American Science 67, 660-9. Bosinski, G. 1963. Eine mittelpaläolithiische Formengruppe und das Problem ihrer geochronologischen Einordnung. Eiszeitalter und Gegenwart S14. Bosinski, G. 1967. Die Mittelpaläolithischen Funde im Westlichen Mitteleuropa. Cologne, Fundamenta Reihe A/4. Bosinski, G. (Ed) 1969a. Die Bocksteinschmiede in Lonetal (Markung Rammingen, Kreis Ulm). Teil I: Text. Stuttgart, Verlag Müller & Gräff.

130

Bosinski, G. (Ed) 1969b. Die Bocksteinschmiede in Lonetal (Markung Rammingen, Kreis Ulm). Teil II: Tafeln. Stuttgart, Verlag Müller & Gräff. Bosinski, G. 1973. Der Mittelpaläolithische Fundplatz Rörshain, Stadtkreis Schwalmstadt (Hessen). In H.-J. MüllerBeck (Ed) Neue paläolithische und mesolithische Ausgrabungen in der Bundesrepublik Deutschland, pp 27-28. Tübingen. Bosinski, G. 1982. The transition from Lower/Middle Palaeolithic in Northwestern Germany. In A. Ronen (Ed) The Transition from Lower to Middle Palaeolithic and the Origin of Modern Man, pp 165-75. British Archaeological Reports International Series 151. Oxford, BAR Publishing. Bosinski, G. 1995. Stone artefacts of the European Lower Palaeolithic: a short note. In W. Roebroeks and T. van Kolfschoten The Earliest Occupation of Europe, pp 263-8. Leiden, University of Leiden. Bosinski, G., Brunnacker, K., Lanser, K.P., Stephan, S., Urban, B. and Würges, K. 1980. Altpaläolithische Funde von Kärlich, Kreis Mayen-Koblenz (Neuwieder Becken). Archäologisches Korrespondenzblatt 10, 295-314. Bosinski, G., van Kolfschoten, T. and Turner, E. 1988. Miesenheim I. Die Zeit des Homo erectus. Andernach, Andernache Beiträge 2. Bottema, S. 1974. Late Quaternary vegetational history of northwest Greece. Unpublished Doctoral Thesis, University of Gronigen. Bottema, S. 1979. Pollen investigations in Thessaly, Greece. Palaeohistoria 21, 19-40. Bourdieu, P. 1977. Outline of a Theory of Practice. Trans. R. Nice. Cambridge, Cambridge University Press. Bourdieu, P. 1990. The Logic of Practice. Trans. R. Nice. Cambridge, Polity Press. Bozilova, E. and Smit, A. 1982. Palynology of Lake Sucho Ezero from south Rila Mountain (Bulgaria). Fitologija 11, 54-67. Brande, A. 1975. Vegetationsgeschichtliche und pollenstratigraphische Untersuchungen zum Paläolithikum von Mauern und Meilenhofen (Fränkische Alb). Quartär 26, 73-106. Braudel, F. 1949. La Méditerranée et le Monde Méditerranéen à l’époque de Phillipe II. Paris, Libraire A. Colin. Breuil, H. 1925. Stations paléolithique en Transylvanie. Bulletin de la Société des Sciences de Cluj 2, 193-217. Broecker, W.S., Thurber, D.L., Goddard, J., Ku, T.L., Matthews, R.K. and Mesolella, K.J. 1968. Milankovitch hypothesis supported by precise dating of coral reefs and deep-sea sediments. Science 159, 297-300. Brunnacker, K., Farrokh, F. and Sidiriopoulos, D. 1982. Die altquartären Terrassen rechts des Niederrheins. Zeitschrift für Geomorphologie N.F. Suppl. 42, 215-26. Burdukiewicz, M. 1990. Wyniki badań stanowiska dolnopaleolitycznego Trzebnica 2. Šlăske Sprawozdania Archeologiczne 31, 7-24. Burdukiewicz, M. 1991. Badanie osadnictwa dolnopaleolitycznego w Trzebnicy. Šlăske Sprawozdania Archeologiczne 32, 7-19. Burdukiewicz, M. and Winnicki, J. 1989. Nowe materialy paleolitu dolnego z Trzebnicy, woj. Wroclaw. Silesia Antiqua 31, 9-18. Burjachs, F. and Julià, R. 1994. Abrupt Climatic Changes during the Last Glaciation Based on Pollen Analysis of the Abric Romani, Catalonia, Spain. Quaternary Research 42, 308-15. Butzer, K.W. 1982. Archaeology as Human Ecology. Cambridge, Cambridge University Press.

131

Callow, P. 1986. The Lithic Industries. In P.Callow and J.M.Cornford (Eds) La Cotte de St. Brelade 1961-1978: Excavations of C.B M. McBurney, pp 199-362. Norwich, Geo. Callow, P. and Webb, R.E. 1981. The application of multivariate statistical techniques to Middle Palaeolithic assemblages from southwestern France. Revue d’Archéometrie 5, 129-38. Campbell, J.B. 1977. The Upper Palaeolithic of Britain: a Study of Man and Nature in the Late Ice Age. Oxford, Clarendon Press. Campbell, J.B. 1980. Le problème des subdivisions du Paléolithique supérieur brittanique dans son cadre européen. Bulletin de la Société Royale Belge d’Anthropologie et Préhistoire. 91, 39-77. Campen, I. and Hahn, J. 1975. Eine Grabung auf dem Mittelpaläolithischen Fundplatz Rörshain, Gemeinde Schwalmstadt, Kreis Ziegenhain. Fundberichte aus Hessen 13 (1973), 75-85. Cârciumaru, M. 1979. Paysage paléophytogéographique, variations du climat et géochronologie du paléolithique moyen et supérieure de Roumanie (Étude palynologique). Dacia NS XXIII, 21-29. Cârciumaru, M. 1988. L’Environnement et le cadre chronologique du paléolithique moyen en Roumanie. In H Laville (Ed) L’Homme du Néandertal, Vol. 2, L’Environnement, pp 45-54. Liège, ERAUL 29. Cârciumaru, M. 1995. La Transition du Paléolithique Moyen au Paléolithique Supérieur en Roumanie: Contexte Paléoclimatique et Chronologique. In Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, pp 101-3. Carrión, J.S. 1992a. Late Quaternary pollen sequence from Carihuela Cave, southeastern Spain. Review of Palaeobotany and Palynology 71, 37-77. Carrión, J.S. 1992b. A palaeoecological study in the western Mediterranean area. The Upper Pleistocene pollen record from Cova Beneito (Alicante, Spain). Palaeogeography, Palaeoclimatology, Palaeoecology 92, 1-14. Carrión, J.S. and Munuera, M. 1997. Upper Pleistocene palaeoenvironmental change in Eastern Spain: new pollenanalytical data from Cova Beneito (Alicante). Palaeogeography, Palaeoclimatology, Palaeoecology 128, 287-99. Cattelain, P. and Otte, M. 1985. Sondage 1984 au ‘Trou de l’Abime’ à Couvin: état des recherches. Helinium XXV, 123-30. Červinka, J.L. 1927. Pravĕk zemí Českých. Pravĕk 1-2, 33-80. Chabai, V.P. 1998. The History of Crimean Middle Paleolithic Studies. In A.E. Marks and V.P. Chabai (Eds) The Middle Paleolithic of Western Crimea, Vol. 1, pp 1-15. Liège, ERAUL 84. Chabai, V.P. and Marks, A.E. 1998. Preliminary Synthesis: Middle Paleolithic Assemblage Variability in Western Crimea. In A.E. Marks and V.P. Chabai (Eds) The Middle Paleolithic of Western Crimea, Vol. 1, pp 355-67. Liège, ERAUL 84. Chaline, J. and Jerz, H. 1984. Stratotypen des Würm-Glazials. Eiszeitalter und Gegenwart 35, 185-206. Chaoimh, Ú.N. 1998. European Snow Cover and its Influence on Spring and Summer Temperatures. The Geographical Journal 164, 41-54. Chappellaz, J., Barnola, J.M., Raynaud, D., Koretkevich, Y.S. and Lorius, C. 1990. Ice-core record of atmospheric methane over the past 160,000 years. Nature 345,127-31. Chen, F.H., Bloemendal, J., Wang, J.M., Li, J.J. and Oldfield, F. 1997. High resolution multi-proxy climate records from Chinese loess: evidence for rapid climate changes over the last 75 kyr. Palaeogeography, Palaeoclimatology, Palaeoecology 130, 323-35. Chirica, V. 1990. La présence des pointes foliacées dans le Paléolithique Supérieur de la Roumanie. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 163-71. Liège, ERAUL 42.

132

Chirica, V. 1995. Les Pièces Bifaciales et la Transition du Paléolithique Moyen au Paléolithique Supérieur en Roumanie. In Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, 105-9. Chmielewski, W. 1961. Civilisation de Jerzmanowice. Wroclaw-Warsaw-Krakow. Chmielewski, W. 1969. Ensembles Micoquo-prondnikiens en Europe Centrale. Geographica Polonica 17, 371-86. Chmielewski, W. 1972. The Continuity and Discontinuity of the Evolution of Archaeological Cultures in Central and Eastern Europe between the 55th and 25th millenaries BC. In F. Bordes (Ed) Origine de l’Homme Moderne, pp1739. Paris, Unesco. Chmielewski, W. 1977. Chronologia lessów w Muselijewie na podstawie analizy źródeł archeologicznych. Studia Geologica Polonica 52, 97-113. Chmielewski, W., Konecka-Betley, K. and Madeyska, T. 1977. Palaeolithic site Kraków-Zwierziniec I in the light of the investigations carried out in 1972-4. Biuletyn Instytutu Geologicznego 305, Z Badań Czwartorzędu w Polsce 20, 13-30. Chmielewski, W., Kowalski, K., Madeyska-Niklewska, T. and Sych, L. 1967. Wyniki badań osadów jaskini Koziarni w Sąspowie, pow. Olkusz. Folia Quaternaria 26. Clark, J.G.D. 1952. Prehistoric Europe: The Economic Basis. London, Methuen. Clark, J.G.D. 1954. Excavations at Star Carr. Cambridge, Cambridge University Press. Clark, J. G. D. 1972. Star Carr: a Case Study in Bioarchaeology. Addison-Wesley Module in Anthropology No. 10. London, Cummings. Clark, G. A. 1992. Continuity or Replacement? Putting Modern Human Origins in an Evolutionary Context. In H.L. Dibble and P.A. Mellars (Eds) The Middle Paleolithic: Adaptation, Behavior and Variability, pp 183-205. Philadelphia, University of Pennsylvania, University Museum Monographs No. 72. Cliquet, D. 1982. Les industries de Saint-Julien de la Liègue: Le Moustérien à petits bifaces dominants. Mémoire de maîtrise, Paris I. Cliquet, D. 1995. Les industries moustériennes à petits bifaces dominants de Haute-Normandie (France). Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, 127-31. Conard, N.J. 1990. Laminar Lithic Assemblages from the Last Interglacial Complex in Northwestern Europe. Journal of Anthropological Research 46, 243-62. Connell, J.H. and Sousa, W.P. 1983. On the evidence needed to judge ecological stability or persistence. American Nature 121, 789-824. Cook, J., Stringer, C.B., Currant, A.P., Schwartz, H.P. and Wintle, A.G. 1982. A review of the chronology of the European middle pleistocene hominids. Yearbook of Physical Anthropology 25, 19-65. Coope, G.R. 1974. Tibetan species of dung beetle from the late Pleistocene of England. Nature 245, 335-6. Cordy, J.-M. 1988. Apport de la paléozoologie à la paléoécologie et à la chronostratigraphie en Europe du nordoccidental. Actes du Colloque de Liège (December 1986), L’homme de Néanderthal 2 pp55-64. Liège, ERAUL 29. Cordy, J.-M. 1992. Bio- et chrono stratigraphie des dépôts quaternaires à partir des micromammifères. In M. Otte (Ed) Recherches aux Grottes de Sclayn Vol. 1: Le Contexte, pp 79-126. Liége, ERAUL. Cordy, J-M. and Bastin, B. 1992. Synthèse des études paléontologiques. In M. Otte (Ed) Recherches aux Grottes de Sclayn Vol. 1: Le Contexte, pp 153-6. Liége, ERAUL.

133

Cornford, J.M. 1986. Specialised resharpening techniques and evidence of handedness. In P. Callow and J.M. Cornford (Eds) La Cotte de St. Brelade 1961-1978: Excavations of C.B.M. McBurney, pp 337-51. Norwich, Geo. Dakaris, S.I., Higgs, E.S. and Hey, R.W. 1964. The climate, environment and industries of Stone Age Greece: part I. Proceedings of the Prehistoric Society 30, 199-244. Dakaris, S.I., Higgs, E.S. and Hey, R. 1964. The climate, environment and industries of stone age Greece, Part I. Proceedings of the Prehistoric Society 30, 199-244. Dansgaard, W. and Tauber, H. 1969. Glacier oxygen-18 content and Pleistocene ocean temperatures. Science 166, 499502. Dansgaard, W, Johnsen, S.J., Clausen, H.B., Dahl-Jensen, D., Gundestrup, N.S., Hammer, C.U., Hvidberg, C.S., Steffensen, J.P., Sveinbjörnsdottir, A.E., Jouzel, J. and Bond, G. 1993. Evidence for general instability of past climate from a 250-kyr ice-core record. Nature 364, 218-20. Davidson, I. and Noble, W. 1993. Tools and language in human evolution. In K.R. Gibson and T. Ingold (Eds) Tools, Language and Cognition in Human Evolution, pp 363-88. Cambridge, Cambridge University Press. Dawkins, R. 1976. The Selfish Gene. Oxford, Oxford University Press. Dawkins, R. 1982. The Extended Phenotype. Oxford, Oxford University Press. Dawkins, R. 1986. The Blind Watchmaker. London, Longman. Desbrosse, R. and Tavoso, A. 1970. Un gisement moustérien à Blanzy (Saône et Loire). Quartär Bd 21, 1-46. Desbrosse, R. and Texier, P.J. 1973a. La station moustérienne de Blissy-sur-Fley (Saône et Loire). La Physiophile 79, 8-31. Desbrosse, R. and Texier, P.J. 1973b. Les silex de Germolles dans la collection Jeannin. La Physiophile 79, 64-70. Dibble, H.L. 1983. Variability and change in the Middle Paleolithic of western Europe and the near East. In E. Trinkaus (Ed) The Mousterian Legacy, pp53-71. British Archaeological Reports International Series S164. Oxford, BAR Publishing. Dibble, H.L. 1984. Interpreting typological variation of Middle Paleolithic scrapers: function, style or sequence of reduction? Journal of Field Archaeology 11, 431-36. Dibble, H.L. 1988. Typological aspects of reduction and intensity of utilization of lithic resources in the French Mousterian. In H.L.Dibble and A. Montet-White (Eds) Upper Pleistocene Prehistory of Western Eurasia, pp 18198. Philadelphia, University of Pennsylvania, University Museum Monographs No. 54. Dibble, H.L. and Lenoir, M. 1995. The Middle Palaeolithic Site of Combe-Capelle Bas, (France). Philadelphia, University of Pennsylvania. Dibble, H.L. and Rolland, N. 1992. On assemblage variability in the Middle Palaeolithic of Western Europe: history, perspectives and a new synthesis. In H.L. Dibble and P.A. Mellars (Eds) The Middle Paleolithic; Adaptation, Behavior and Variability, pp 1-28. Philadelphia, University of Pennsylvania, University Museum Monographs No. 72. Dobzhansky, T., Ayala, F.J., Stebbins, G.L. and Valentine, J.W. 1977. Evolution. San Francisco, Freeman. Donner, J. 1996. The Early and Middle Weichselian Interstadials in the Central Area of the Scandinavian Glaciations. Quaternary Science Reviews 15, 471-79. Dooley, J.L. and Bowers, M.A. 1998. Demographic responses to habitat fragmentation; experimental tests at the landscape and patch scale. Ecology 79, 969-80. Douglas, M. 1973. Natural Symbols. Harmondsworth, Pelican Books.

134

Drozdowski, E. and Fedorowicz, S. 1987. Stratigraphy of Vistulian glaciogenic deposits and corresponding thermoluminescence dates in the lower Vistula region, northern Poland. Boreas 16, 139-53. Dzambazov, N. 1964. Prochvaniya na paleolitnata i mesolitnata kultura v Bulgariya. Arkheologiya 6, 67-76. Dzambazov, N. 1967. Les pointes bifaciale dans les grottes Samouilitsa I et II et les autres stations du Paléolithique récent en Bulgarie. Quartär 18, 195-9. Dzambazov, N. 1968. Paleolitno nakhodishte pri s. Muselievo, Plevenski okrg. Arkheologiya 10, 54-64. Eglinton, G., Bradshaw, S.A., Rosell, A., Sarnthein, M., Pflaumann, U. and Tiedemann, R. 1992. Molecular record of secular sea surface temperature changes on 100-year timescales for glacial terminations I, II and IV. Nature 356, 423-426. Emiliani, C. 1955. Pleistocene temperatures. Journal of Geology 63, 538-75. Emontspohl, A.-F. 1995. The northwest European vegetation at the beginning of the Weichselian glacial (Brørup and Odderade interstadials) – new data for northern France. Review of Palaeobotany and Palynology 84, 231-42. Erd, K. 1973. Pollenanalytische Gliederung des Pleistozäns der Deutschen Demokratischen Republik. Zietschrift für geologische Wissenschaft 1, 1087-1103. d’Errico, F., Zilhão, J., Julien, M., Baffier, D. and Pelegrin, J. 1998. Neanderthal acculturation in western Europe? A critical review of the evidence and its interpretation. Current Anthropology 39, S1-S44. Evans-Pritchard, E.E. 1940. The Nuer: A Description of the Modes of Livelihood and Political Institutions of a Nilotic People. Oxford, Oxford University Press. Faustova, M.A. 1984. Late Pleistocene Glaciation of European USSR. In A.A. Velichko (Ed: English Translation by H.E. Wright Jr. and C.W. Barnosky) Late Quaternary Environments in the Soviet Union, pp 3-12. Minneapolis, University of Minnesota Press. Farizy, C. 1985. Un habitat du paléolithique moyen à Champlost (Yonne, France). Archäologisches Korrespondenzblatt 15, 405-10. Farizy, C. 1995. Industries charentiennes à influences micoquiennes, l’exemple de l’Est de France. Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, pp 173-78. Farizy, C., David, J. and Jaubert, J. 1994. Hommes et Bisons du Paléolithique Moyen à Mauran (Haute-Garonne). Paris, CNRS (Gallia-Préhostoire Supplement 30). Féblot-Augustins, J. 1993. Mobility strategies in the late Middle Palaeolithic of central Europe and western Europe: elements of stability and variability. Journal of Anthropological Archaeology 12, 211-65. Féblot-Augustins, J. 1997a. La Circulation des Matiéres Premières au Paléolithique. Tome I. Liège, ERAUL 75. Féblot-Augustins, J. 1997b. La Circulation des Matiéres Premières au Paléolithique. Tome II. Liège, ERAUL 75. Filzer, P. 1969. Pollenanalytische Untersuchungen der Höhlen- und Hangschichten. In G. Bosinski (Ed) Die Bocksteinschmiede in Lonetal (Markung Rammingen, Kreis Ulm). Teil I: Text, pp 169-206. Stuttgart, Verlag Müller & Gräff. Flannery, K.V. 1967. Culture history v. cultural process: a debate in American Archaeology. Scientific American 217, 119-22. Flannery, K.V. 1968. Archeological systems theory and early Mesoamerica. In B.J. Meggers (Ed) Anthropological Archeology in the Americas, pp 67-87. Washington DC, Anthropological Society of Washington. Florschütz, F., Menendez Amor, J. and Wijmstra, T.A. 1971. Palynology of a thick Quaternary succession in southern Spain. Paleogeography, Palaeoclimatology, Palaeoecology 10, 233-64.

135

Foley, R.A. 1984. Putting people into perspective: an introduction to community ecology and ecology. In R.A. Foley (Ed) Hominid Evolution and Community Ecology, pp. 1-24. New York, Academic Press. Foley, R.A. 1987. Another unique species. Patterns in human evolutionary ecology. London, Longman. Follieri, M., Magri, D. and Sadori, L. 1988. 250,000-year pollen record from Valle di Castiglione (Roma). Pollen et Spores 30, 329-56. Follieri, M., Magri, D. and Sadori, L. 1989. Pollen stratigraphical synthesis from Valle di Castiglione (Roma). Quaternary International 3/4, 81-4. Forman, R.T.T. 1995. Land Mosaics. The Ecology of Landscapes and Regions. Cambridge, Cambridge University Press. Formozov, A.A. 1958. The cave site Starosele and its place in the Palaeolithic. Materials and Investigations of the Archaeology of the USSR 71 (In Russian). Moscow. Fraipont, J. and Tihon, F. 1889. Explorations scientifiques des cavernes de la vallée de la Méhaigne. Mémoires de l'Académie royale des Sciences, des Lettres et des Beaux-Arts de Belgique XLIII. Fraser, F.C. and King, J.E. 1954. Faunal remains. In J.G.D. Clark (Ed) Excavations at Star Carr, pp 70-95. Cambridge, Cambridge University Press. Freericks, M. 1995. Transition du Paléolithique moyen au Paléolithique supérieur en Allemagne: quelques exemples. In Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, 117-22. Frenzel, B. 1980. Das Klima der letzten Eiszeit in Europa. In H. Oeschger, B. Messerli and M. Svilar (Eds) Das Klima, pp 45-63. Berlin, Springer. Freund, G. 1952. Die Blattspitzen des Paläolithikums in Europa. Bonn, Quartär-Bibliothek 1. Freund, G. 1961. Zur Frage paläolithische Besiedlung der Unteren Klause bei Neu-Essing, Landkreis Kelheim. Germania 39, 1-7. Freund, G. 1963. Die ältere und die mittlere Steinzeit in Bayern. Jahrsbericht der Bayerischen Bodendenkmalpfliege 4, 9-167. Freund, G. 1970. Die Sesselfelsgrotte, eine altsteinzeitliche Siedlungsstelle im unteren Altmühltal. Neue Ausgrabungen in Bayern. Munich. Freund, G. 1984. Die Sesselfelsgrotte im unteren Altmühltal. Führer zu archäologischen Denkmälern in Deutschland 6 (Regensburg, Kelheim, Straubing) Teil 2, 78-89. Freund, G. 1987. Das Paläolithikum der Oberneder-Höhle (Landkreis Kelheim/Donau). Quartär-Bibliothek Band 5. Bonn, Ludwig Röhrscheid. Freund, G. 1998. Sesselfelsgrotte I, Grabungsverlauf und Stratigraphie. Saarbrücken, Quartär-Bibliothek Band 8. Frickhinger, E. 1937. Grabungen in der Kleinen Ofnethöhle bei Holheim. Bayerische Vorgeschichtsblatt 14, 35. Fridrich, J. 1973. Počátky mladopaleolitického osídlení Čech. Archeologické Rozhledy 25, 392-442. Fridrich, J. 1989. Prezletice: A Lower Palaeolithic Site in Central Bohemia (Excavations 1969-1985). Prague, Prague National Museum, Fontes Arch. Prag. 18. Fridrich, J. 1991. The Oldest Palaeolithic Stone Industry from the Beroun Highway Complex. Anthropozoikum 20, 11128.

136

Fridrich, J. and Sklenář, K. 1976. Die paläolithische und mesolithische Höhlenbesiedlung des Böhmischen Karstes. Prague. Gábori, M. 1976. Les Civilisations du Paléolthique moyen entre les Alpes et l’Oural. Budapest. Gábori, V. 1958. Neue paläolithische Funde im Eipel-Tal. Archeologické Rozhledy 10, 55-61. Gábori-Csánk, V. 1956. Beiträge zur Frage des Slowakischen Szeletien. Archaeologiai Értesitö 83, 78-83. Gábori-Csánk, V. 1968. La Station du Paléolithique moyen d’Érd, Hongrie. Budapest. Gábori-Csánk, V. 1970. C14 dates of the Hungarian Palaeolithic. Acta Archaeologica Academiae Scientiarum Hungaricae 22, 3-11. Gábori-Csánk, V. 1974. Nouvelles observations sur le Szélétien. Compte-rendus des communications, Analyse par attributs du matériel dea phases initiales du Paléolithique supérieur européen, 3e réunion, pp 2-7. Paris. Gábori-Csánk, V. 1983. La Grotte Remete ‘Felsö’ (Supérieur) et le ‘Szeletien de Transdanubie’. Acta Archaeologica Academiae Scientiarum Hungaricae 3, 249-85. Gamble, C.S. 1986. The Palaeolithic Settlement of Europe. Cambridge, Cambridge University Press. Gamble, C.S. 1993. Exchange, foraging and local hominid networks. In C. Scarre and F. Healy (Eds) Trade and Exchange in Prehistoric Europe, pp. 35-44. Oxford, Oxbow Monograph 33. Gamble, C.S. 1995a. The Earliest Occupation of Europe: The Environmental Background. In W. Roebroeks and T. van Kolfschoten (Eds) The Earliest Occupation of Europe, pp 279-95. Leiden, University of Leiden. Gamble, C.S. 1995b. Interpretation in the Palaeolithic. In I. Hodder, M. Shanks, A. Alexandri, V. Buchli, J. Carman, J. Last and G. Lucas (Eds) Interpreting Archaeology, Finding Meaning in the Past, pp 87-91. Gamble, C.S. 1996. Making tracks: hominid networks and the evolution of the social landscape. In J. Steele and S.J. Shennan (Eds) The Archaeology of Human Ancestry Power, Sex and Ideology, pp 253-77. Gamble, C.S. 1998. Palaeolithic society and the release from proximity: a network approach to intimate relations. World Archaeology 29, 426-49. Gamble, C.S. 1999. The Palaeolithic Societies of Europe. Cambridge, Cambridge University Press. Gavela, B. 1971. Gradac et Risovača. In G. Novak (Ed) Époque Préhistorique et Protohistorique en Yougoslavie – Rechereches et Résultats, pp 176-77. Belgrade. Geneste, J.-M. 1988. Les industries de la Grotte Vaufrey: technologie du débitage, économie et circulation de la matière première lithique. In J.-P. Rigaud (Ed) La Grotte Vaufrey: Paléoenvironnement, Chronologie, Activités Humaines pp 441-517. Paris, Mémoires de la Société Préhistorique Française. Geneste, J.-M. 1989a. Économie des ressources lithiques dans le Moustérien du sud-ouest de la France. In M. Patou and L.G. Freeman (Eds) Homme de Néandertal, Vol. 6, pp 75-97. Liège, ERAUL 33. Geneste, J.-M. 1989b. Systèmes d’approvisionnement en matières premières au paléolithique moyen et au paléolithique supérieur en Aquitaine. In J.K. Kozlowski (Ed) L’Homme de Néandertal Vol. 8, pp 61-70. Liège, ERAUL 35. Genty, D., Blamart, D., Ouahdi, R., Gilmour, M., Baker, A., Jouzel, J. and Van-Exter, S. 2003. Precise dating of Dansgaard−Oeschger climate oscillations in western Europe from stalagmite data. Nature 421, 833-37. Gibson, J.J. 1979. The ecological approach to visual perception. Boston, Houghton Miflin. Gibson, K.R. 1993. Generative interplay between technical capacities, social relations, imitation and cognition. In K.R. Gibson and T. Ingold (Eds) Tools, Language and Cognition in Human Evolution, pp 131-7. Cambridge, Cambridge University Press.

137

Gibson, K.R. and Ingold, T. (Eds) 1993. Tools, Language and Cognition in Human Evolution. Cambridge, Cambridge University Press. Giddens, A. 1984. The Constitution of Society: Outline of the Theory of Structuration. Cambridge, Polity Press. Gladilin, V.N. 1989. The Korol’evo Palaeolithic Site: Research, Methods, Stratigraphy. Anthropologie 27, 93-103. Gladilin, V.N. and Demidenko, Y.E. 1990. On the Origin of Early Upper Palaeolithic Industries with Leaf Points in the Carpatho-Balkan Region. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 115-124. Liège, ERAUL 42. Gladilin, V.N. and Sitlivyj, V.I. 1990. Ašel Central’noj Evropy. Kiev, Naukova Dumka. Gladilin, V.N., Sitlivyj, V.I. and Tkachenko, V.I. 1995. Les Premières Industries à Pointes Foliacées en Europe Centrale. In Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, 111-16. Glynn, I.M. 1993. The evolution of consciousness: William James’s unresolved problem. Biological Reviews 68, 599616. Godron, M. 1982. L’étude du ‘grain’ de la structure de la végétation: application à quelques exemples méditerranéens. Ecologia Mediterranea 8, 191-5. Golovanova, L.V., Hoffecker, J.F., Kharitinov, V.M. and Romanova, G.P. 1999. Mezmaiskaya Cave: A Neanderthal Occupation in the Northern Caucasus. Current Anthropology 40, 77-86. Goodall, J. 1990. Through a Window: Thirty Years with the Chimpanzees of Gombe. London, Weidenfeld and Nicolson. Gosden, C. 1994. Social Being and Time. Oxford, Blackwell. Gould, S.J. 1977. Ontogeny and Phylogeny. Cambridge MA, Belknap Press of Harvard University Press. Gould, S.J. and Lewontin, R.C. 1979. The spandrels of San Marco and the panglossian paradigm: a critique of the adaptationist programme. Proceedings of the Royal Society of London B205; 581-98. Gould, S.J. and Vrba, E.S. 1982. Exaptation – a missing term in the science of form. Paleobiology 8, 4-15. Gowlett, J.A.J. and Crompton, R.H. 1994. Kariandusi: Acheulean morphology and the question of allometry. The African Archaeological Review 12, 3-42. Grahmann, R. 1955. The Lower Palaeolithic site of Markleeburg and other comparable sites near Leipzig. Transactions of the American Philosophical Society 45, 509-687. Grene, M. 1969. Hierarchy: one word, how many concepts? In L.L. Whyte, A.G. Wilson, and D. Wilson (Eds) Hierarchical Structures, pp 56-8. New York, Elsevier. Grichuk, V.P. 1984. Late Pleistocene Vegetation History. In A.A. Velichko (Ed: English Translation by H.E. Wright Jr. and C.W. Barnosky) Late Quaternary Environments in the Soviet Union, pp 155-78. Minneapolis, University of Minnesota Press. Grootes, P.M., Stuiver, M., White, J.W.C., Johnsen, S. and Jouzel, J. 1993. Comparison of oxygen isotope records from from the GISP2 and GRIP Greenland ice cores. Nature 366, 552-4. Grote, K. 1978. Die Grabung 1977 in der mittelpaläolithischen Freilandstation Salzgitter-Lebenstedt. Archäologisches Korrespondenzblatt 8, 155-62. Grüger, E. 1979. Spätrisz, Risz/Würm und Frühwürm am Samerberg in Oberbayern – ein vegetationgeschichtlicher Beitrag zur Gliederung des Jungpleistozäns. Geologica Bavarica 80, 5-64. Grüger, E. 1989. Palynostratigraphy of the last interglacial/glacial cycle in Germany. Quaternary International 3-4, 6979.

138

Guiot, J., Pons, A., Beaulieu, J.-L. and Reille, M. 1989. A 140,000-year continental climate reconstruction from two European pollen records. Nature 338, 309-13. Gumpert, K. 1952. Höhlengrabung im Schambachtal. Jahresbericht des historische Veröffentlich für Mittelfranken 72, 5. Gumpert, K. 1956. Breitenfurt (Ldkr. Eichstätt). Bayerische Vorgeschichtsblatt 21, 135. Günther, K. 1964. Die altsteinzeitliche Funde der Balve Höhle. Bodenaltertümer Westfalens 8. Guthrie, R.D. 1968. Paleoecology of the large mammal community in interior Alaska during the late Pleistocene. Amer. Midl. Nat. 79, 346-63. Guthrie, R.D. 1984. Mosaics, Allelochemics and Nutrients: An Ecological Theory of Late Pleistocene Megafaunal Extinctions. In P.S. Martin and R.G. Klein (Eds) Quaternary Extinctions: A Prehistoric Revolution, pp 259-98. Tucson, University of Arizona Press. Guthrie, R.D. 1990. Frozen Fauna of the Mammoth Steppe. Chicago, Chicago University Press. Haesarts, P. and Sirakova, S. 1979. Le Paléolithique moyen à pointes foliacées de Mousselievo (Bulgarie). In J.K. Kozlowski (Ed) Middle and Early Upper Palaeolithic in the Balkans, pp 35-63. Prace Archeologiczne 28. Hahn, J. 1977. Aurignacien: Das Ältere Jungpaläolithikum in Mittel- und Osteuropa. Cologne, Fundamenta A9. Hahn, J. 1990. La technologie des pointes bifaciales de Rörshain et leurs relations avec l’Allemagne du Sud. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 79-86. Liège, ERAUL 42. van der Hammen, T. 1971. The Upper Quaternary stratigraphy of the Dinkel valley. Mededelingen Rijks Geologische Dienst NS 22, 55-214. van der Hammen, T., Wijmstra, T.A. and Zagwijn, W.H. 1971. The floral record of the late Cenozoic in Europe. In K.K. Turekian (Ed) The late Cenozoic glacial ages, pp 391-424. New Haven, Yale University Press. Hanski, I. 1999. Metapopulation Ecology. Oxford, Oxford University Press. Harmon, R.S., Glazek, J. and Nowak, K. 1980. site. Nature 284, 132-5.

230

Th/234U dating of travertine from the Bilzingsleben archaeological

Harris, M. 1979. Cultural Materialism: The Struggle for a Science of Culture. New York, Random House. Hays, J.D., Imbrie, J. and Shackleton, N.J. 1976. Variations in the Earth’s orbit; pacemaker of the ice ages. Science 194, 1121-32. Hedges, R.E.M., Housley, R.A., Pettitt, P.B., Bronk Ramsey, C. and Van Klinken, G.J. 1996. Radiocarbon dates from the Oxford AMS system: Archaeometry datelist 21. Archaeometry 38, 181-207. Hedges, R.E.M., Pettitt, P.B., Bronk Ramsey, C. and Van Klinken, G.J. 1998. Radiocarbon dates from the Oxford AMS system: Archaeometry datelist 25. Archaeometry 40, 227-39. Heinrich, H. 1988. Origin and consequences of cyclic ice rafting events in the Northeast Atlantic Ocean during the past 130,000 years. Quaternary Research 29, 142-52. Helle, M., Sønstergaard, E., Coope, G.R. and Rye, N. 1981. Early Weichselian peat at Brumunddal, southeastern Norway. Boreas 10, 369-79. Henry, J.D. and Swan, J.M.A. 1974. Reconstructing forest history from live and dead plant material – an approach to the study of forest succession in southwest New Hampshire. Ecology 55, 772-83. Hewes, G.W. 1973. Primate communication and the gestural origin of language. Current Anthropology 14, 5-24.

139

Hewes, G.W. 1993. A history of speculation on the relation between tools and language. In K.R. Gibson and T. Ingold (Eds) Tools, Language and Cognition in Human Evolution, pp 20-31. Cambridge, Cambridge University Press. Higgs, E.S. 1965. Some recent Old Stone Age discoveries in Epirus. Archaiologikon Deltion 20, Chronika, 361-74. Higgs, E.S. and Jarman, M.R. 1975. Palaeoeconomy. In E.S. Higgs (Ed) Palaeoeconomy, pp 1-7. Cambridge, Cambridge University Press. Higgs, E.S. and Vita-Finzi, C. 1966. The climate, environment and industries of stone age Greece, Part II. Proceedings of the Prehistoric Society 32, 1-29. Higgs, E.S. and Vita-Finzi, C. 1972. Prehistoric economies: a territorial approach. In E.S. Higgs (Ed) Papers in Economic Prehistory, pp 27-36. Cambridge, Cambridge University Press. Higgs, E.S., Vita-Finzi, C., Harris, D.R. and Fagg, A.E. 1967. The climate, emvironment and industries of stone age Greece, Part III. Proceedings of the Prehistoric Society 33, 1-29. Hillebrand, E. 1917. Das Solutréen Ungarns. Prähistorische Zeitschrift 9, 18-21. Hillebrand, E. 1935. Die ältere Steinzeit Ungarns. Archaeologica Hungarica 17, 1-40. Hodder, I.R. 1982a. The Present Past: An Introduction to Anthropology for Archaeologists. London, Batsford. Hodder, I.R. 1982b. Symbols in Action: Ethnoarchaeological Studies of Material Culture. Cambridge, Cambridge University Press. Hodder, I.R. (Ed) 1982c. Symbolic and Structural Archaeology. Cambridge, Cambridge Universtity Press. Hodder, I.R. 1989. ‘This is not an article about material culture as text’. Journal of Anthropological Archaeology 8, 250-69. Honea, K. 1984. Chronometry of the Romanian Middle and Upper Paleolithic: Implications of current radiocarbon dating results. Dacia NS 28, 23-39. Honea, K. 1986. Dating and periodization strategies of the Romanian Middle and Upper Palaeolithic: A retrospective overview and assessment. Pleistocene Perspectives Vol. 1. London, Unwin and Allen. Honea, K. 1990. Recent advances in modern archaeological dating (AMS, ESR, U-Th): First Oxford AMS dates for Mitoc-Malu Galben. Arheologia Moldovei 8, 9-12. Houmark-Nielsen, M. 1989. The last interglacial-glacial cycle in Denmark. Quaternary International 3, 31-39. Houmark-Nielsen, M. and Kolstrup, E. 1981. A radiocarbon dated Weichselian sequence from Sejerø, Denmark. Geologiska Föreningens i Stockholm Förhandlingar 103, 73-78. Hovan, S.A., Rea, D.K., Pisias N.G. and Shackleton, N.J. 1989. A direct link between the China loess and marine O18 records: aeolian flux to the north Pacific. Nature 340, 296-8. Howell, F.C. 1960. European and northwest African Middle Pleistocene hominids. Current Anthropology 1, 195-232. Hublin, J.-J., Ruiz, C.B., Lara, P.M., Fontugne, M. and Reyss, J.-L. 1995. The Mousterian site of Zafarraya (Andalucia, Spain): dating and implications on the palaeolithic peopling processes of Western Europe. Comptes-Rendus de l’Académie des Sciences de Paris t.321, serie IIa, 931-7. Huguenin, G. 1988. Le peuplement humain du bassin supérieur de la Saône au Paléolithique inférieur et moyen. Thèse de l’Ecole pratique des Hautes-Etudes, Dijon. Hülle, W.M. 1977: Die Ilsenhöhle unter Burg Ranis/Thüringen. Stuttgart. Huntley, B. 1996. Quaternary Palaeoecology and Ecology. Quaternary Science Reviews 15, 591-606.

140

Huntley, B. and Birks, H.J.B. 1983. An atlas of past and present pollen maps for Europe; 0-13,000 years ago. Cambridge, Cambridge University Press. Huxtable, J. 1986. The thermoluminescence dates. In P. Callow and J.M. Cornford (Eds) La Cotte de St. Brelade 19611978: Excavations of C.B.M. McBurney, pp 145-9. Norwich, Geo. Imbrie, J., Hays, J.D., Martinson, D.G., McIntyre, A., Mix, A.C., Morley, J.J., Pisias, N.G., Prell, W.L. and Shackleton, N.J. 1984. The orbital theory of Pleistocene climate: support from a revised chronology of the marine δ18O record. In A. Berger, J. Imbrie, J. Hays, G. Kukla and B. Saltzman (Eds) Milankovitch and Climate, pp 269-306. Dordrecht, Reidel. Imbrie, J. and Imbrie, K.P. 1979. Ice Ages: Solving the Mystery. London, MacMillan. Ingold, T. 1980. Hunters, Pastoralists and Ranchers. Cambridge, Cambridge University Press. Ingold, T. 1986a. The Appropriation of Nature Essays on Human Ecology and Social Relations. Manchester, Manchester University Press. Ingold, T. 1986b. Prologue: concerning the hunter and his spear. In T. Ingold The Appropriation of Nature:Essays on Human Ecology and Social Relations, pp 1-15. Manchester, Manchester University Press. Ingold, T. 1986c. The architect and the bee: reflections of the work of animals and men. In T. Ingold The Appropriation of Nature:Essays on Human Ecology and Social Relations, pp 16-39. Manchester, Manchester University Press. Ingold, T. 1988a. Introduction. In T. Ingold (Ed) What is an Animal?, pp 1-16. London, Unwin Hyman. Ingold, T. 1988b. The animal in the study of humanity. In T. Ingold (Ed) What is an Animal?, pp 84-99. London, Unwin Hyman. Ivanova, S. 1979. Cultural Differentiation in the Middle Palaeolithic on the Balkan Peninsula. In J.K. Kozlowski (Ed) Middle and Early Upper Palaeolithic in the Balkans, pp13-33. Prace Archeologiczne 28. Iversen, J. 1954. The late glacial flora of Denmark and its relation to climate and soil. Danmarks Geol. Unders. Sev II 80, 87-119. Jacobi, R.M. 1990. Leaf-points and the British early Upper Palaeolithic. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 271-89. Liège, ERAUL 42. James, W. 1879. Are we automata? Mind 4, 1-22. Janásek, J. and Skutil, J. 1954. Nová Solutréenská stanice v Kvasicích na Moravĕ. Archeologické Rozhledy 6, 729-33. Jansen, E. 1989. The use of stable oxygen and carbon isotope stratigraphy as a dating tool. Quaternary International 1, 151-66. Jarman, M.R. 1972. A territorial model for archaeology: a behavioural and geographical approach. In D.L. Clarke (Ed) Models in Archaeology, pp 705-34. London, Methuen. Jastrzębska-Mamelka, M. 1979. Eemian and Early Vistulian organic sediments at Zgierz-Rudunki near Łódź. Bulletin de la Société de Science et Lettres de Łódź 29, 1-6. Jaubert, J., Lorblanchet, M., Laville, H., Slott-Moller, R., Turq, A. and Brugal, J.-P. (Eds) 1990. Les Chasseurs d’Aurochs de La Borde: un site du Paléolithique moyen (Livernon, Lot). Paris, Maison des Sciences de l’Homme. Johnsen, S.J., Clausen, H.B., Dansgaard, W., Fuhrer, K., Gundestrup, N., Hammer, C.U., Iversen, P., Jouzel, J., Stauffer, B. and Steffensen, J.P. 1992. Irregular glacial interstadials recorded in a new Greenalnd ice core. Nature 359, 407-12. Jones, R.L. 1986. The flora and vegetation of La Cotte de St Brelade and its environs. In P. Callow and J.M. Cornford (Eds) La Cotte de St. Brelade 1961-1978: Excavations of C.B.M. McBurney, pp 99-106. Norwich, Geo.

141

Kadić, O. 1916. Ergibnesse der Erforschung der Szeletahöhle. Mitteilungen aus dem Jahrbuche der Königlich Ungarischen Geologischen Anstalt 22, 161-301. Karlson, R.H. and Cornell, H.V. 1998. Scale-dependent variation in local vs. regional effects on coral species richness. Ecological Monographs 68, 259-74. Kaulich, B., Nadler, M. and Reisch, L. 1978. Führer zu urgeschichtlichen Höhlenfundplätzen des unteren Altmühltales. Erlangen. Kay, M. and Hardy, B.L. 1997. Pointing with a purpose: Crimea Middle Paleolithic projectile points. Presentation delivered to International Colloquium ‘Anatolian Prehistory on the Crossroads of Eurasia and Africa’, Liège, 28th April – 3rd May 1997. King, T. 1996. Quantifying nonlinearity and geometry in time series of climate. Quaternary Science Reviews 15, 24766. Klein, R.G. 1969. The Mousterian of European Russia. Proceedings of the Prehistoric Society 35, 77-111. Klein, R.G. 1999. The Human Career: Human Biological and Cultural Origins. 2nd Edition. Chicago, Universtity of Chicago Press. Klíma, B. 1957. Übersicht über die jüngsten paläolithischen Forschungen in Mähren. Quartär 9, 85-130. Klíma, B. 1969. Nové nálezy na paleolitické stanici u Křepic. Časopis Moravského Muzea 53/54, 31-50. Klíma, B. 1978. Paleolitická stanice u Přestalvk. okr, Přerov. Archeologické rozhledy 30, 5-13. von Koenigswald, W. 1974. Die pleistozäne Fauna der Weinberghöhlen bei Mauern. In W. von Koenigswald, H.-J. Müller-Beck and E. Pressmar Die Archäologie und Paläontologie in der Weinberghöhlen bei Mauern (Bayern), pp 53-106. Tübingen, Archaeologica Venatoria 3. von Koenigswald, W. 1992. Various aspects of migrations in terrestrial mammals in relation to Pleistocene faunas of central Europe. In W. von Koenigswald and L, Werdelin (Eds) Mammalian Migration and Dispersal Events in the European Quaternary, pp 39-47. Courier Forschungsinstitut Senckenberg 153. von Koenigswald, W., Müller-Beck, H.-J. and Pressmar, E. 1974. Die Archäologie und Paläontologie in der Weinberghöhlen bei Mauern (Bayern). Tübingen, Archaeologica Venatoria 3. Koestler, A. 1969. Beyond atomism and holism – the concept of the holon. In A. Koestler and J.R. Smythies (Eds) Beyond Reductionism, pp 192-232. London, Hutchinson. Kolen, J. 1999. Hominids without homes: on the nature of Middle Palaeolithic settlement in Europe. In W. Roebroeks and C. Gamble (Eds) The Middle Palaeolithic Occupation of Europe, pp 139-63. Leiden, University of Leiden. Kolosov, Y. 1990. The Mousterian Leaf Points in Crimea: Classification, Chronology, Origins. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 53-61. Liège, ERAUL 42. Kolosov, Y. 1995. À Propos de Deux Outils de la Culture Moustérienne d’Ak-Kai en Crimée. In Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, 217-20. Kolstrup, E. and Wijmstra, T.A. 1977. A palynological investigation of the Moershoofd, Hengelo and Denekamp interstadials in the Netherlands. Geologie en Mijnbouw 56, 85-102. Kowalski, K. 1967. The Pleistocene extinction of mammals in Europe. In P.S. Martin and H.E. Wright Jr. (Eds) Pleistocene Extinctions: the Search for a Cause, pp 343-64. New Haven, Yale University Press. Kowalski, S. 1969. Nowe Dane do Poznania Kultury Jerzmanowickiej w Polsce. Światowit 30, 177-88. Kozlowski, J.K. 1962. Un gisement paléolithique supérieur à Dzierzysław, district Głubczyce, en Haute-Silésie. Archaeologia Polona 4, 64-76.

142

Kozlowski, J.K. 1975. Badania nad przejściem od środkowego do górnego paleolitu na Bałkanach. Przegląd Archeologiczny 23, 5-48. Kozlowski, J.K. 1983. Le paléolithique supérieur en Pologne. L’Anthropologie 87, 49-82. Kozlowski, J.K. 1988. The transition from Middle to Early Upper Paleolithic in Central Europe and the Balkans. In J.F. Hoffecker and C.A. Wolf (Eds) The Early Upper Paleolithic, pp 193-235. British Archaeological Reports International Series 437. Oxford, BAR Publishing. Kozlowski, J.K. (Ed) 1990a: Feuilles de Pierre: Les Industries à Pointes Foliacées du Paléolithique Supérieure Européen. Liège, ERAUL 42. Kozlowski, J.K. 1990b. Certains aspects techno-morphologiques des pointes foliacées de la fin du Paléolithique moyen et du début du Paléolithique supérieur en Europe centrale. In C. Farizy (Ed) Paléolithique moyen récent et Paléolithique supérieur ancien en Europe, pp 125-33. Mémoires du Musée de Préhistoire d’Ile de France 3. Kozlowski, J.K. 1995. La signification des “outils foliacées”. In Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, 91-99. Kretzoi, M. and Dobosi, V. (Eds) 1990. Vérteszöllös – Man, Site and Culture. Budapest, Akadémiai Kiadó. Kretzoi, M. and Vértes, L. 1965. Upper Biharian (Intermindel) Pebble Industry Occupation Site in Western Hungary. Current Anthropology 6, 74-87. Krzyszkowski, D. 1990. Middle and Late Weichselian stratigraphy and palaeoenvironments in central Poland. Boreas 19, 333-50. Kröger, K., van den Bogaard, P., Bittman, F. and Turner, E. 1988. Der Fundplatz Kärlich-Seeufer. Neue Untersuchungen zum Altpaläolithikum im Rheinland. Jahrbuch des Römisch-Germanischen Zentralmuseums Mainz 35, 111-35. Kuhn, S.L. 1995. Mousterian Lithic Technology: An Ecological Perspective. Princeton, Princeton University Press. Kukla, G., An, Z.S., Melice, J.L., Gavin, J. and Xiao, J.L. 1990. Magnetic susceptibility record of Chinese loess. Transactions of the Royal Society of Edinburgh: Earth Sciences 81, 263-88. Kulakovskaya, L. 1990. Leaf-shaped Implements in the ‘Eastern Micoquian’ of the European Part of the URSS. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 43-48. Liège, ERAUL. Kupryjanowicz, M. 1991. Eemian, Early and Late Vistulian and Holocene vegetation in the region of Machnacz peatbog near Białystok. Acta Palaeobotanica 31, 215-25. Larsen, E., Gulliksen, S., Lauritzen, S.-E., Lie, R., Løvlie, R. and Mangerud, J. 1987. Cave stratigraphy in western Norway: multiple Weichselian glaciations and interstadial vertebrate faunas. Boreas 16, 267-92. Laville, H. 1975. Climatologie et Chronologie du Paléolithique en Périgord: étude sédimentologique de dêpots en grottes et sous abris. Marseilles, Université de Provence, Études Quaternaires, Mémoire No. 4. Laville, H. 1990. Le remplissage de l’aven: charactéristiques et signification. In J. Jaubert, M. Lorblanchet, H. Laville, R. Slott-Moller, A. Turq and J.-P. Brugal (Eds) Les Chasseurs d’Aurochs de La Borde: un site du Paléolithique moyen (Livernon, Lot), pp 23-32. Paris, Maison des Sciences de l’Homme. Laville, H., Rigaud, J.-P. and Sackett, J.R. 1980. Rock Shelters of the Perigord. London, Academic Press. Legge, A.J. and Rowley-Conwy, P.A. 1989. Some preliminary Results of a Re-examination of the Star Carr Fauna. In C. Bonsall (Ed) The Mesolithic in Europe, pp 225-30. Edinburgh, John Donald. Lehman, S. 1993. Ice sheets, wayward winds and seachange. Nature 365, 108-10.

143

Leroi-Gourhan, A. 1993. Gesture and Speech. Trans. A.B. Berger. Cambridge Mass., MIT Press. [1964. Le Geste et la Parole. Paris, Albin Michel.] Leroi-Gourhan, Arl. 1988. Le passage Moustérien-Chatelperronien à Arcy-sur-Cure. Bulletin de la Société Préhistorique Française 85, 102-4. Leroi-Gourhan, Arl. 1997. Chauds et Froids de 60,000 à 15,000 BP. Bulletin de la Societé Préhistorique Française 94, 151-60. Leroy, S.A.G., Giralt, S., Francus, P. and Seret, G. 1996. The high sensitivity of the palynological record in the Vico maar lacustrine sequence (Latium, Italy) highlights the climatic gradient through Europe for the last 90 ka. Quaternary Science Reviews 15, 189-201. Leroyer, C. 1986. Les gisements Castelperroniens de Quinçay et de Saint-Césaire: quelques comparaisons préliminaires des études palynologiques. In B. Vandermeersch (Ed) Préhistoire de Poitou-Charentes: Problèmes Actuel, pp 12534. Paris, Editions du Comité des Travaux Historiques et Scientifiques. Lévi-Strauss, C. 1955. Tristes Tropiques. Paris, Librairie Plon. Lewontin, R.C. 1966. Is nature probable or capricious? Bioscience 16, 25-7. Lewontin, R.C. 1978. Adaptation. Scientific American 239, 157-69. Lewontin, R.C. 1993. The Doctrine of DNA: Biology as Ideology. London, Penguin. Lieberman, P. 1984. The Biology and Evolution of Language. Cambridge MA., Harvard University Press. Lieberman, P. and Crelin, E.S. 1971. On the speech of Neanderthal man. Linguistic Enquiry 11, 203-22. Little, M.G., Schneider, R.R., Kroon, D., Price, B., Bickert, T. and Wefer, G. 1997. Rapid palaeoceanographic changes in the Benguela Upwelling System for the last 160,000 years as indicated by abundances of planktonic foraminifera. Palaeogeography, Palaeoclimatology, Palaeoecology 130, 135-61. Lloyd, D. and Rossi, E.L. 1993. Biological rhythms as organization and information. Biological Reviews 68, 563-77). Lloyd, D. and Stupfel, M. 1991. The occurrence and function of ultradian rhythms. Biological Reviews 66, 275-99. de Loë, A. 1906. Fouilles dans la terrasse du ‘Trou del’Abime’ à Couvin (province de Namur). Bulletin Musées Royaux d’Art et Histoire VI(1), 6-7. Lorius, C., Jouzel, J., Ritz, C., Merlivat, L., Barkov, N.I., Korotkevich, Y.S. and Kotlyakov, V.M. 1985. A 150,000 year climatic record from Antarctic Ice. Nature 316, 591-6. Lowe, J.J. and Walker, M.J.C. 1997. Reconstructing Quaternary Environments, 2nd Edition. Harlow, Addison Wesley Longman. Ložek, V. 1996. Pleistocene Paleoenvironments. In J. Svoboda, V. Ložek and E. Vlček Hunters between East and West: The Paleolithic of Moravia, pp 15-36. New York, Plenum Press. Lukács, G. 1980. The Ontology of Social Being: Labour. London, Merlin Press. Luttrop, A. 1970. Mittelpaläolithische Funde von Rörshain, Kreis Ziegenhain (Hessen). In K. Gripp, R. Schütrumpf and H. Schwabedissen (Eds) Frühe Menschheit und Umwelt Part I. Archäologisches Beiträge, pp131-3. CologneVienna. Luttrop, A. and Bosinski, G. 1967. Rörshain, Kreis Ziegenhain. Fundberichte aus Hessen 7, 13-17. Mai, D.H. 1988. Einige exotische Gehölze in den Interglazialfloren der Mitteldeutschen Florenregion.

144

Mai, D.H., Mania, D., Nötzold, T., Toepfer, V., Vlček, E. and Heinrich, W.D. 1983. Bilzingsleben II. Berlin, Veröff. Landesmus. Vorgesch. Halle 36. Malez, M. 1976. Excavation of the villafranchian site Šandala I near Pula (Jugoslavia). In K. Valoch (Ed) Les premiers industries de l’Europe, pp 104-23. Nice, IX Congress UISPP, Colloque VIII. Mamakowa, K. 1989. Late Middle Polish Glaciation, Eemian and Early Vistulian vegetation at Imbramowice near Wrocław and the pollen stratigraphy of this part of the Pleistocene in Poland. Acta Palaeobotanica 29, 11-176. Mamakowa, K., Mook, W. and Środoń, A. 1975. Late Pleistocene flora at Kąty (Pieniny Mts., West Carpathians). Acta Palaeobotanica XVI, 147-172. Mangerud, J., Gulliksae, S., Larsen, E., Longvar, O., Miller, G.H., Sejrup, H.-P. and Sønstegaard, E. 1981. A Middle Weichselian ice-free period in Western Norway: the Ålesund Interstadial. Boreas 10, 447-62. Mania, D. 1975. Bilzingsleben (Thüringen): eine neue altpaläolithische Fundstelle mit Knochenreste des Homo erectus. Archäologisches Korrespondenzblatt 5, 263-72. Mania, D. 1983. Zur Jagd des Homo erectus von Bilzingsleben. Ethnographische-archäologische Zeitschrift 24, 32637. Mania, D. 1984. Zur Geochronologie des Mittelpleistozäns und einiger paläolithischen Fundstellen im Saale- und mittleren Elbegebiet. Arbeits- und Forschungsberichte Dresden 27/28, 13-58. Mania, D. 1988. Le Paléolithique ancien et moyen de la région de la Saale et l’Elbe, Allemagne de l’Est. L’Anthropologie 92, 1051-92. Mania, D. 1990. Auf den Spuren des Urmenschen. Die Funde auf der Steinrinne bei Bilzingsleben. Berlin. Mania, D. and Baumann, W. 1980. Neufunde des Acheuléen von Markkleeburg bei Leipzig (DDR). Anthropologie 18, 237-48. Mania, D. and Toepfer, V. 1973. Königsaue: Gliederung, Ökologie und mittelpaläolithische Funde der letzten Eiszeit. Berlin, Deutscher Verlag der Wissenschaften. Mardones, M. and Jalut, G. 1983. La Tourbière de Biscaye (Alt. 409m, Hautes Pyrénées): Approche paléoécologique des 45.000 dernières années. Pollen et Spores XXV/2, 163-212. Marks, A.E. and Chabai, V.P. (Eds) 1998. The Middle Paleolithic of Western Crimea, Vol. 1. Liège, ERAUL 84. Marks, A.E. and Monegal, K. 1998. Starosele 1993-1995: The Lithic Artefacts. In A.E. Marks and V.P. Chabai (Eds) The Middle Paleolithic of Western Crimea, Vol. 1, pp 117-65. Liège, ERAUL 84. Martinson, D.G., Pisias, N.G., Hays, J.D., Imbrie, J., Moore, T.C. and Shackleton, N.J. 1987. Age dating and the orbital theory of the ice ages: development of a high resolution 0-300,000 year chronostratigraphy. Quaternary Research 27, 1-29. Marx, K. 1969. The Eighteenth Brumaire of Louis Bonaparte. In K. Marx and F. Engels Selected Works Vol. 1, pp 394474. Moscow, Progress Publishers. Maryanski, A.R. 1993. The elementary forms of the first protohuman society: an ecological/social network approach. Advances in Human Ecology 2, 215-41. Mauss, M. 1967. The Gift. Trans. I. Cunnison. London, Norton. [1925: Essai sur le don, forme archaïque de l’échange.] Mayhew, D.F. 1977. New information on British Quaternary lagomorphs and larger rodents. 10th International Quaternary Association Congress, Abstracts. p 296. Mayr-Lenoir, A. 1927. Paläolithvorkommnisse im mittleren Altmühltal. Nachrichten der Deutsche Anthropologische Geschicht 2, 30.

145

MacArthur, R.H. and Levins, R. 1964. Competition, habitat selection and character displacement in a patchy environment. Proceedings of the National Academy of Science (USA) 51, 1207-10. McBurney, C.M.B. 1948. The Stone Tools of Neanderthals. Unpublished PhD Dissertation, University of Cambridge. McBurney, C.M.B. 1950. The Geographical Study of the Older Palaeolithic Stages in Europe. Proceedings of the Prehistoric Society 11, 163-83. McGrew, W.C. 1992. Chimpanzee Material Culture: Implications for Human Evolution. Cambridge, Cambridge University Press. McManus, J., Oppo, D., Cullen, J. and Healy, S. 2003. Marine Isotope Stage 11 (MIS 11): Analog for Holocene and future climate? American Geophysical Union Geophysical Monograph 137, 69-85. Mertens, S.B. 1996. The Middle Paleolithic of Romania. Current Anthropology 37, 515-21. Mellars, P.A. 1967. The Mousterian Succession in South-west France. Unpublished PhD Dissertation, University of Cambridge. Mellars, P.A. 1969. The chronology of Mousterian industries in the Périgord region of south-west France. Proceedings of the Prehistoric Society 35, 134-71. Mellars, P.A. 1986. A new chronology for the French Mousterian period. Nature 322, 410-11. Mellars, P.A. 1992. Technological change in the Mousterian of southwest France. In H.L. Dibble and P.A. Mellars (Eds) The Middle Paleolithic: Adaptation, Behavior and Variability, pp 29-43. Philadelphia, University of Pennsylvania, University Museum Monographs No. 72. Mellars, P.A. 1996. The Neanderthal Legacy. Princeton, Princeton University Press. Mellars, P.A. and Grün, R. 1991. A comparison of the electron spin resonance and thermoluminescence dating methods: the results of ESR dating at Le Moustier (France). Cambridge Archaeological Journal 1, 269-76. Mellars, P.A. and Stringer, C. (Eds) 1989. The Human Revolution: Behavioural and Biological Perspectives in the Origins of Modern Humans. Edinburgh, Edinburgh University Press. Menke, B. and Tynni, R. 1984. Das Eeminterglazial und das Weichselfrühglazial von Rederstall/Dithmarschen und ihre Bedeutung für die mitteleuropäische Jungpleistzän-Gliederung. Geologisches Jahrbuch A76, 120. Mertens, S.B. 1996. The Middle Paleolithic of Romania. Current Anthropology 37, 515-21. Meyer, H-H. and Kottmeyer, C. 1989. Die atmosphärische Zirkulation in Europa im Hochglazial der Weichsel-Eiszeit – abgeleitet von Paläowind-Indikatoren und Modellsimulationen. Eiszeitalter und Gegenwart 39,10-18. Meyer, R.K.F. and Schmidt-Kaler, H. 1991. Wanderungen in die Erdgeschichte: Durchs Urdonautal nach Eichstätt. Munich, Naturpark Altmühltal. Milardo, R.M. 1992. Comparative methods for delineating social networks. Journal of Social and Personal Relationships 9, 447-61. Mithen, S. 1994. Technology and society during the Middle Pleistocene: hominid group size, social learning and industrial variability. Cambridge Archaeological Journal 4, 3-32. Mithen, S. 1996. The Prehistory of the Mind: A Search for the Origins of Art, Religion and Science. London, Thames and Hudson. Mix, A.C. 1987. The oxygen isotope record of glaciation. In W.F. Ruddiman and H.E. Wright (Eds) The Geology of North America, Vol. K3, North America and Adjacent Oceans During the Last Deglaciation, pp. 111-35. Boulder Colorado, Geological Society of America.

146

Mix, A.C. and Ruddiman, W.F. 1985. Structure and timing of the last deglaciation; oxygen isotope evidence. Quaternary Science Reviews 4, 59-108. Movius, H.L. 1948. The Lower Palaeolithic cultures of southern and eastern Asia. Transactions of the American Philosophical Society 38, 329-420. Movius, H.L. 1949. Lower Paleolithic archaeology in southern and eastern Asia. Studies in Physical Anthropology 1, 17-81. Movius, H.L. 1955. Palaeolithic archaeology in southern and eastern Asia, exclusive of India. Cahiers d’Histoire Mondiale 2, 257-82. Müller-Beck, H-J. 1957. Das Obere Altpaläolithikum in Suddeutschland, 1. Habelt, Bonn. Müller-Beck, H-J. 1966. Paleohunters in America: Origins and Diffussion. Science 152, 1191-210. Müller-Beck, H-J. 1974. Die pleistozänen Sedimente und ihre archäologischen Einschlüsse in dem Weinberghöhlen bei Mauern. In W. von Koenigswald, H.-J. Müller-Beck and E. Pressmar Die Archäologie und Paläontologie in der Weinberghöhlen bei Mauern (Bayern), pp 17-31. Tübingen, Archaeologica Venatoria 3. Müller-Beck, H-J. 1988. The Ecosystem of the “Middle Paleolithic” (Late Lower Paleolithic) in the Upper Danube Region.: A Stepping-Stone to the Upper Paleolithic. In H.L. Dibble and A. Montet-White (Eds) Upper Pleistocene Prehistory of Western Eurasia, pp 233-54. Philadelphia, University of Pennsylvania, University Museum Monographs No. 54. Musil, R. and Valoch, K. 1968. Stránská Skála: its Meaning for Pleistocene studies. Current Anthropology 9, 534-9. Mussi, M. 1995. The earliest occupation of Europe: Italy. In W. Roebroeks and T. van Kolfschoten (Eds) The Earliest Occupation of Europe, pp 27-49. Leiden, University of Leiden. Neustupný, E. and Neustupný, J. 1960. Nástin pravĕkých dĕjin Československa. Sborník Národního Musea v Praze 14, 3-5, 95-216. Nicolăescu-Plopşor, C.S. 1957. Le Paléolithique dans la République populaire Roumanie à la lumière des dernières recherches. Dacia NS 1, 1-60. Nicolăuescu-Plopşor, C.S. and Bassa, B. 1957. Cercetări asupra paleoliticului timpuriu. Materiale 3, 281-5, 291. Nicolăuescu-Plopşor, C.S. and Covacs, E. 1959. Cercetările paleolitice din regiunea Baia Mare. Materiale 6, 33-42. Nicolăuescu-Plopşor, C.S., Păunescu, A., Bolomey, A., Pop, I. and Zaharia, N.N. 1959. Paport preliminar asupra certetărilor paleolitice din anul 1956. Materiale 5, 15-43. Nicolăuescu-Plopşor. C.S., Păunescu, A. and Mogosanu, F. 1966. Le Paléolithique de Ceahlău. Dacia NS 10, 5-116. Nicolis, G. 1989. Physics of far-from-equilibrium systems and self-organization. In P. Davis (Ed) The New Physics. Cambridge, Cambridge University Press. Noble, W. and Davidson, I. 1996. Human Evolution, Language and Mind: A Psychological and Archaeological Enquiry. Cambridge, Cambridge University Press. Norrdahl, K. 1995. Population cycles in northern small mammals. Biological Reviews 70, 621-37. Nötzold, T. 1983. Charophytenfruktifikationen von Bilzingsleben. In D.H. Mai, Mania, D., Nötzold, T., Toepfer, V., Vlček, E. and Heinrich, W.D. (Eds) Bilzingsleben II, pp 41-44. Berlin, Veröff. Landesmus. Vorgesch. Halle 36. Obermaier. H. and Fraunholz, J. 1926. Eine Mammutdarstellung aus Süddeutschland. IPEK. Obermaier, H. and Fraunholz, J. 1927. Der skulptierte Rengeweihstab aus der Mittleren Klausenhöhle bei Essing (Niederbayern). IPEK.

147

Obermaier, H. and Wernert, P. 1914. Paläolithbeiträge aus Nordbayern. Mitteilung der Anthropologischen Gesellschaft in Wien 44, 44-63. Obermaier, H. and Wernert, P. 1929. Alt-Paläolithikum mit Blatt-typen. Mitteilung der Anthropologischen Gesellschaft in Wien 59, 293-310. Oliva, M. 1979. Die Herkunft des Szeletien im Lichte neue Funde von Jezeřany. Mitteilungen der Anthropologischen Gesellschaft in Wien 64, 45-78. Oliva, M. 1981. Die Bohunicien-Station bei Podoli (Bez. Brno-Land) und ihre Stellung im beginnenden Jungpaläolithikum. Časopis Moravského Muzea 66, 7-45. Oliva, M. 1989. Paleolit. In L. Belcredi (Ed) Archeologické lokality a nálezy okresu Brno-venkov, pp 8-32. Brno, Okresní Muzeum. Oliver, J.S., Sikes, N.E. and Stewart, K.M. (Eds) 1994. Early Hominid Behavioral Ecology. London, Academic Press. O’Neill, R.V., DeAngelis, D.L., Waide, J.B. and Allen, T.F.H. 1986. A Hierarchical Concept of Ecosystems. Princeton, Princeton University Press. Oppo, D.W., McManus, J.F. and Cullen, J.L. 1998. Abrupt Climate Events 500,000 to 340,000 Years Ago: Evidence from Subpolar North Atlantic Sediments. Science 279, 1335-8. Otte, M. 1976. L’Aurignacien en Belgique. In J.K. Kozlowski (Ed) L’Aurignacien en Europe, pp 144-63. Nice, Colloque XVI, UISPP IX Congrés. Otte, M. 1977. Données générales sur le Paléolithique supérieur ancien de Belgique. L’Anthropologie 81, 235-72. Otte, M. 1984. Paléolithique supérieur en Belgique. In D. Cahen and P. Haesarts (Eds) Peuples Chasseurs de la Belgique Préhistorique dans leur cadre Naturel, pp 157-79. Brussels, Institut Royal des Sciences Naturelles. Papagianni, D. 1999. Beyond ‘flint scatters’ and ‘findspots’: assessing the potential for compiling a synthesis of the Greek Middle Palaeolithic surface data. In G.N. Bailey, E. Adam, E. Panagopoulou, C. Perles and K. Zachos (Eds) The Palaeolithic Archaeology of Greece and Adjacent Areas: Proceedings of the ICOPAG Conference, Ioannina, pp 130-36. British School at Athens Studies 3. Paquereau, M.M. 1975. Le Würmien ancien en Périgord: Étude palynologique. Quaternaria 18, 67-159. Parker, S.T. and Milbraith, C. 1993. Higher intelligence, propositional language and culture as adaptations for planning. In K.R. Gibson and T. Ingold (Eds) Tools, Language and Cognition in Human Evolution, pp 314-33. Cambridge, Cambridge University Press. Patience, A.J. and Kroon, D. 1991. Oxygen isotope chronostratigraphy. In P.L.Smart and P.D. Frances (Eds) Quaternary Dating Methods – a User’s Guide, pp 199-228. Cambridge, Quaternary Research Association Technical Guide 4. Patten, B.C. and Odum, E.P. 1981. The cybernetic nature of ecosystems. American Naturalist 118, 886-95. Păunescu, A. 1965. Sur la succession des habitats paléolithiques et postpaléolithiques de Ripiceni-Izvor. Dacia N.S. IX, 5-32. Păunescu, A. 1988a. Chronologie du Paléolithique Moyen en Roumanie dans le contexte de l’Europe Centre-Orientale et Meridionale. In H.P. Schwarcz (Ed) L’Homme de Néandertal Vol. 1, pp 73-80. Liège, ERAUL 28. Păunescu, A. 1988b. Le Passage du Paléolithique Moyen au Paléolithique Supérieur entre les Carpates et le Prut. In J.K. Kozlowski (Ed) L’Homme de Néandertal Vol. 8, pp 133-47. Liège, ERAUL 35. Păunescu, A. 1989. Structures d’Habitat Moustériennes Mises au Jour dans l’Établissement de Ripiceni-Izvor (Roumanie) et Quelques Considerations Concernant le Type d’Habitat Paléolithique Moyen de l’Est des Carpates. In M. Patou and L.H. Freeman (Eds) L’Homme de Néandertal Vol. 6, pp 127-143. Liège, ERAUL 33.

148

Păunescu, A. 1993. Ripiceni-Izvor, paleolitic şi mezolitic: studiu monografic. Bucharest, Editura Acadamiei Romne. Penck, A. and Brückner, E. 1909. Die Alpen in Eiszeitalter. Leipzig, Tachnitz. Peterson, G., Allen, C.R. and Holling, C.S. 1998. Ecological Resilience, Diversity and Scale. Ecosystems 1, 6-18. Petit, J.R., Jouzel, J., Raynaud, D., Barkov, N.I., Barnola, J.-M., Basile, I., Bender, M., Chapellaz, J., Davis, M., Delaygue, G., Delmotte, M., Kotlyakov, V.M., Legrand, M., Lipenkov, V.Y., Lorius, C., Pépin, L., Ritz, C., Saltzman, E. and Stievenard, M. 1999. Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature 399, 429-36. Pisias, N.G., Martinson, D.G., Moore, T.C. Jr., Shackleton, N.J., Prell, W., Hays, J. and Boden, G. 1984. High resolution stratigraphic correlation of benthic oxygen isotopic records spanning the last 300,000 years. Marine Geology 56, 119-36. Pons, A., Guiot, G., de Beaulieu, J.-L. and Reille, M. 1992. Recent Contributions to the Climatology of the Last Glacial-Interglacial Cycle based on French Pollen Sequences. Quaternary Science Reviews 11, 439-48. Pons, A. and Reille, M. 1988. The Holocene and Upper Pleistocene pollen record from Padul (Granada, Spain): A new study. Palaeogeography, Palaeoclimatology, Palaeoecology 66, 243-63. Poulianos, A. 1982. The Cave of the Petralonian Archanthropinae: a Guide to the Science behind the Excavations. Athens, Library of the Anthropological Association of Greece. Prentice, M.L. and Denton, G.H. 1988. The deep-sea oxygen record, the global ice sheet system and hominid evolution. In F.E. Grine (Ed) Evolutionary history of the “robust” australopithecines, pp 383-403. New York, Aldine de Gruyter. Prošek, F. 1951. Výzkum Jeskynĕ Dzeravé Skály v Malých Karpatech. Archeologické Rozhledy 3, 293-8. Prošek, F. 1953. Szeletien na Slovensku. Slovenská Archeología 1, 133-94. de Puydt, M. and Lohest, M. 1885-6. Exploration à la grotte de Spy. Annales de la Société Geologique de Belge XIII, 34-9. de Puydt, M. and Lohest, M. 1886. L’homme contemporain du Mammouth à Spy (Namur). Fédération Archéologique et Historique de Belgique, Annales et Compte Rendus des Travaux du Congrés 2nd Session, Namur 1886, 205-40. Raposo, L. and Santonja, M. 1995. The earliest occupation of Europe: the Iberian Peninsula. In W. Roebroeks and T. van Kolfschoten (Eds) The Earliest Occupation of Europe, pp. 7-25. Leiden, University of Leiden. Rasmussen, T.L., van Weering, T.C.E. and Labeyrie, L. 1997. Climatic instability, ice sheets and ocean dynamics at high northern latitudes during the last glacial period (58-10 Ka BP). Quaternary Science Reviews 16, 71-80. Raymo, M.E., Ganley, K., Carter, S., Oppo, D.W. and McManus, J. 1998. Millennial-scale climate instability during the early Pleistocene epoch. Nature 392, 699-702. Raynal, J-P., Magoga, L., Fatima-Zohra, S-A. and Geraards, D. 1995. The earliest occupation of Atlantic Morocco: the Casablanca evidence. In W. Roebroeks and T. van Kolfschoten (Eds) The Earliest Occupation of Europe, pp. 25562. Leiden, University of Leiden. Reed, E.S. 1988a. The affordances of the animate environment: social science from the ecological point of view. In T. Ingold (Ed) What is an Animal?, pp 10-26. London, Unwin Hyman. Reed, E.S. 1988b. James J. Gibson and the Psychology of Perception. New Haven, Yale University Press. Reille, M. and de Beaulieu, J.-L. 1990. Pollen analysis of a long upper Pleistocene continental sequence in a Velay maar (Massif Central, France). Palaeogeography, Palaeoclimatology, Palaeoecology 80, 35-48.

149

Reille, M., Guiot, J. and de Beaulieu, J.-L. 1992. The Montaigu Event: An Abrupt Climatic Change During the Early Würm in Europe. In G.J. Kukla and E. Went (Eds) Start of a Glacial, pp 85-95. NATO ASI Series 13. Berlin, Springer. Reisch, L. 1985. Die Molluskenreste aus der Schicht O der Sesselfelsgrotte. Eine chronologisch wichtige Fauna mit Fagotia acicularis (Férussac, 1823). Quartär 35/36, 35-54. Renfrew, C. 1973. Before Civilization: The Radiocarbon Revolution and Prehistoric Europe. London, Cape. Richter, J. 1997. Sesselfelsgrotte III. Der G-Schichten-Komplex der Sesselfelsgrotte. Zum Verständnis des Micoquien. Saarbrücken, Quartär-Bibliothek Band 7. Riek, G. 1934. Die Eiszeitjägerstation am Vogelherd. Band 1: Die Kulturen. Tübingen. Riek, G. 1938. Feststellung des Hochsolutréen in Württemberg. Forschungen und Fortschritte 14, 146-8. Ringer, Á. 1983. Bábonyien: eine mittelpaläolithische Blattwerkzeugindustrie in Nordostungarn. Dissertationes Archaeologicae: Régészeti Dolgozatok, Series 2, 11. Budapest, Eötvös Lorand University. Ringer, Á. 1995. Les industries à pièces foliacées en Europe Centrale: Proposition de synthèse. Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, 15-18. Rink, W.J., Lee, H.-K., Rees-Jones, J. and Goodger, K.A. 1998. Electron Spin Resonance (ESR) and Mass Spectrometric U-Series (MSUS) Dating of Teeth in Crimean Paleolithic Sites: Starosele, Kabazi II, and Kabazi V. In A.E. Marks and V.P. Chabai (Eds) The Middle Paleolithic of Western Crimea, Vol. 1, pp 323-40. Liège, ERAUL 84. Roberts, M.B., Gamble, C.S. and Bridgland, D.R. 1995. The earliest occupation of Europe: the British Isles. In W. Roebroeks and T. van Kolfschoten (Eds) The Earliest Occupation of Europe, pp 165-191. Leiden, University of Leiden. Roe, D.A. 1968. British Lower and Middle Palaeolithic Handaxe Groups. Proceedings of the Prehistoric Society 34, 182. Roe, D.A. 1981. The Lower and Middle Palaeolithic Periods in Britain. London, Routledge and Kegan Paul. Roebroeks, W., Conard, N. and van Kolfschoten, T. 1992. Dense forests, cold steppes, and the Palaeolithic settlement of northern Europe. Current Anthropology 33, 551-86. Roebroeks, W. and van Kolfschoten, T. (Eds) 1995a. The Earliest Occupation of Europe. Leiden, University of Leiden. Roebroeks, W. and van Kolfschoten, T. 1995b. The earliest occupation of Europe: a reappraisal of artefactual and chronological evidence. In W. Roebroeks and T. van Kolfschoten (Eds) The Earliest Occupation of Europe, pp 297-315. Leiden, University of Leiden. Rolland, N. 1990. Existe-t’il un Moustérien oriental? In J.K. Kozlowsli (Ed) Feuilles de Pierre, pp 95-112. Liège, ERAUL 42. Romme, W.H. 1982. Fire and landscape diversity in subalpine forests of Yellowstone National Park. Ecological Monographs 52, 199-221. Rose, S., Lewontin, R.C. and Kamin, L.J. 1984. Not in our genes: biology, ideology and human nature. London, Pantheon. Rosu, L. 1966. Les Découvertes de Ocna Sibiului et quelques problèmes du Moustérien tardif en Roumanie. Actes du VIIe Congrès International des Sciences Préhistoriques et Protohistoriques, Prague, 265-9. Rossignol-Strick, M. and Planchais, N. 1989. Climate patterns revealed by pollen and oxygen isotope records of a Tyrrhenian Sea core. Nature 342, 413-6.

150

Roucoux, K.H., N.J. Shackleton, L. de Abreu, J. Schönfeld & P.C. Tzedakis. 2001. Combined marine proxy and pollen analyses reveal rapid Iberian vegetation response to North Atlantic millennial-scale climate oscillations. Quaternary Research 56, 128-32. Rousseau, D-D., Puisségur, J-J. and Lautridou, J.P. 1990. Biogeography of the Pleistocene pleniglacial malacofaunas in Europe: stratigraphic and climatic implications. Palaeogeography, Palaeoclimatology, Palaeoecology 80, 7-23. Ruddiman, W.F. and McIntyre, A. 1976. Northeast Atlantic paleoclimatic changes over the last 600,000 years. Geological Society of America Memoir 145, 199-214. Ruddiman, W.F. and Raymo, M.E. 1988. Northern hemisphere climate regimes during the past 3Ma: possible tectonic connections. Philosophical Transactions of the Royal Society of London B318, 1-20. Saád, A. and Hellebrandt, M. 1974. A Diósgyőr-Tapolca barlang feltárása. A Miskolci Herman Ottó Múzeum Közlemények 13, 10-15. Sancetta, C., Imbrie, J. and Kipp, N.G. 1973. Climatic record of the past 130,000 years in north Atlantic deep sea core V23-82: correlation with the terrestrial record. Quaternary Research 3, 110-6. Schlanger, N. 1996. Understanding Levallois: lithic technology and cognitive archaeology. Cambridge Archaeological Journal 6, 231-54. Schmidt, R.R. 1912. Die diluviale Vorzeit Deutschlands. Stuttgart. Schönweiβ, W. and Werner, H.-J. 1986. Ein Fundplatz des Szeletien in Zeitlarn bei Regensburg. Archäologisches Korrespondenzblatt 16, 7-12. Schultz, H., von Rad, U. and Erlenkeuser, H. 1998. Correlation between Arabian Sea and Greenland climate oscillations of the past 110,000 years. Nature 393, 54-7. Schütrumpf, R von 1951. Die Pollenanalytische Datierung der Altsteinzeitlichen Funde. Palaeohistoria 1, 10-20. Schwarcz, H.P. and Skoflek, I. 1982. New dates for the Tata, Hungary, archaeological site. Nature 295, 590-1. Šercelj, A. 1966. Pelodne analize Pleistocenskih in Holocenskih sedimentov Ljubljanskega Barje. Slovenska Akademia Znanosti in Umetnosti Classis IV Dissertationes 9, 431-72. Shackleton, N.J. 1967. Oxygen isotope analyses and Pleistocene temperatures re-assessed. Nature 215, 15-17. Shackleton, N.J. 1969. The last interglacial in the marine and terrestrial records. Proceedings of the Royal Society of London (B)174, 135-54. Shackleton, N.J. 1977. The oxygen isotope stratigraphic record of the Late Pleistocene. Philosophical Transactions of the Royal Society of London B280, 169-182. Shackleton, N.J. 1987. Oxygen isotopes, ice volume and sea level. Quaternary Science Reviews 6, 183-90. Shackleton, N.J., Imbrie, J. and Hall, M.A. 1983. Oxygen and carbon isotope record of East Pacific core V19-30: implications for the formation of deep water in the late Pleistocene North Atlantic. Earth and Planetary Science Letters 65, 233-44. Shackleton, N.J., Imbrie, J. and Pisias, N.G. 1988. The evolution of oceanic oxygen-isotope variability in the North Atlantic over the past three million years. Philosophical Transactions of the Royal Society of London B318, 67988. Shackleton, N.J. and Opdyke, N.D. 1973. Oxygen isotope and palaeomagnetic stratigraphy of equatorial Pacific core V28-238. Quaternary Research 3, 39-55. Shanks, M. and Tilley, C. 1987. Social Theory and Archaeology. Cambridge, Polity Press.

151

Shanks, M. and Tilley, C. 1992. Reconstructing Archaeology: Theory and Practice (2nd Edn). London, Routledge. Shen, G. and Yokoyama, Y. 1986. T-230/U-234 Dating of Petralona Speleothems. Anthropos 11, 23-32. Siman, K. 1990. Considerations on the “Szeletian Unity”. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 189-198. Liège, ERAUL 42. Simon, H.A. 1962. The architecture of complexity. Proceedings of the American Philosophical Society 106, 467-82. Simon, H.A. 1969. The Sciences of the Artificial. Cambridge MA, MIT Press. Simon, H.A. 1973. The organization of complex systems. In H.H. Pattee (Ed) Hierarchy Theory, pp 3-27. New York, Braziller. Sirakov, N. 1979. Analytical Study on Blanks from the Cave Samuilitsa II (an Example of Application of Attribute Correlation Method). In J.K. Kozlowski (Ed) Middle and Early Upper Palaeolithic in the Balkans. Prace Archeologiczne 28, 65-76. Sirakova, S. 1990. The leafpoints of Muselievo.In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 63-78. Liège, ERAUL 42. Société des Amis du Musée National de Préhistoire et de la Recherche Archéologique 1995. Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale. Sparks, B.W. and West, R.G. 1972. The Ice Age of Britain. London, Methuen. Steele, J. 1994. Communication networks and dispersal patterns in human evolution: a simple simulation model. World Archaeology 26, 126-43. Steward, J.H. 1953. Evolution and process. In A.L. Kroeber (Ed) Anthropology Today, pp 313-26. Chicago, University of Chicago Press. Steward, J.H. 1955. Theory of Culture Change. Urbana, University of Illinois Press. Stiner, M.C. 1991. A taphonomic perspective on the origins of the faunal remains of Grotte Guattari (Latium, Italy). Current Anthropology 32, 103-117. Stiner, M.C. 1994. Honor Among Thieves: A Zooarchaeological Study of Neandertal Ecology. Princeton, Princeton University Press. Stokes, S., Haynes, G., Thomas, D.S.G., Horrocks, J.L., Higginson, M. and Malifa, M. 1998. Punctuated aridity in southern Africa during the last glacial cycle: the chronology of linear dune construction in the northeastern Kalahari. Palaeogeography, Palaeoclimarology, Palaeoecology 137(3-4), 305-22. Straus, L. and Otte, M. 1995. Stone Age Wallonia (Southern Belgium). Current Anthropology 36, 851-4. Stringer, C.B. 1974. A multivariate study of the Petralona skull. Journal of Human Evolution 3, 397-404. Stringer, C.B., Dean, M.C. and Martin, R.D. 1990. A comparative study of cranial and dental development within a recent British sample and among Neanderthals. In C.J. DeRousseau (Ed) Primate Life History and Evolution, pp 115-52. New York, Wiley-Liss. Stringer, C.B. and Gamble, C.S. 1993. In Search of the Neanderthals. London, Thames and Hudson. Svoboda, J. 1990. The Bohunician. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 199-211. Liège, ERAUL 42. Svoboda, J. 1991. Neue Erkentnisse zur Pekárna Höhle im Mährischen Karst. Archäologisches Korrespondenzblatt 21, 39-43. Svoboda, J., Czudek, T., Havlíček, P., Ložek, V., Macoun, J., Přichystal, A., Svobodov, H. and Vlček, E. 1994. Paleolit Moravy a Slezska. Brno, Institute of Archaeology.

152

Svoboda, J., Ložek, V. and Vlček, E. 1996. Hunters between East and West: the Paleolithic of Moravia. New York, Plenum. Taborin, Y. 1990. Les prémices de la parure. In C. Farizy (Ed) Paléolithique Moyen Récent et Paléolithique Supérieur Ancien en Europe, pp 335-44. Nemours, Mémoires de la Musée de Préhistoire de l’Ile de France. Taborin, Y. 1993. Shells of the French Aurignacian and Perigordian. In H. Knecht, A. Pike-Tay and R. White (Eds) Before Lascaux: The Complex Record of the Early Upper Paleolithic, pp 211-29. Boca Raton, CRC Press. Tansley, A.G. 1935: The use and abuse of vegetational concepts and terms. Ecology 16, 284-307. Taute, W. 1966. Bie Mittel-bis Jungpaläolithische Stratigraphie im Hohlen Stein bei Böhmfeld (Bayern) und die Frage einer Höhlenbären-Schädeldeposition. Actes du VIIe Congrès International des Sciences Préhistoriques et Protohistoriques, Prague, pp 308-11. Taylor, L.C., Hammer, C.U., Alley, R.B., Clausen, L.B., Dahl-Jensen, D., Gow, A.J., Gundestrup, L.S., Kipfstuhl, J., Moore. J.C. and Waddington, E.D. 1993. Electrical conductivity measurements from the GISP2 and GRIP Greenland ice cores. Nature 366, 549-52. Teunissen, D. and Teunissen-van Oorschot, H. 1974. Eine interstadiale Torfschicht bei Nijmegen (Niederlande) und deren Bedeutung für die Erklärung der dortigen Landschaftsmorphologie. Geologie en Mijnbouw 53, 393-400. Thieme, H. 1997. Lower Palaeolithic hunting spears from Germany. Nature 385, 807-10. Thieme, H., Maier, R. and Urban, R. 1992. Neue Erkenntnisse zum urgeschichtlichen Siedlungsgeschehen. Archäologie Deutschland 8, 26-30. Thieme, H. and Mania, D. 1993. ‘Schöningen 12’ – ein mittelpleistozänes Interglazialvorkommen im Nordharzvorland mit paläolithischen Funden. Ethnographische-Archäologische Zeitschrift 34, 610-19. Thompson, L.G.,Yao, T., Davis, M.E., Henderson, K.A., Mosley-Thompson, E., Lin, P-N., Beer, J., Synal, H-A., ColeDai, J. and Bolzan, J.F. 1997. Tropical Climate Instability: The Last Glacial Cycle from a Qinghai-Tibetan Ice Core. Science 276, 1821-5. Thouveny, N., de Beaulieu, J-L., Bonifay, E., Creer, K.M., Guiot, J., Icole, M., Johnsen, S., Jouzel, J., Reille, M., Williams, T. and Williamson, D. 1994. Climate variations in Europe over the past 140 kyr deduced from rock magnetism. Nature 371, 503-6. Tilley, C. 1991. Material Culture and Text: The Art of Ambiguity. London, Routledge. Tilley, C. 1994. A Phenomenology of Landscape. Oxford, Berg. Tode, A., Preul, F., Richter, K., Selle, W., Pfaffenberg, K., Kleinschmidt, A., Guenther, E., Müller, A. and Schwartz, W. 1953. Die Untersuchung der paläolithischen Freilandstation von Salzgitter-Lebenstedt. Eiszeitalter und Gegenwart 3, 144-220. Toepfer, V. 1968. Das Clactonien im Saale-Mittelgebiet. Jahresschrift mitteldeutsche Vorgeschicht 57, 1-26. Toepfer, V. 1970. Stratigraphie und Ökologie des Paläolithikums. Petermanns Geograph. Mitt. Ergänzungsheft 274, 329-422. Tomasello, M. 1990. Questions regarding imitation, ‘language’, and cultural transmission in apes and monkeys. In S.T.Parker and K.R. Gibson (Eds) ‘Language’ and Intelligence in Monkeys and Apes: Comparative Developmental Perspectives, pp 274-311. Cambridge, Cambridge University Press. Tomasello, M., Kruger, A.C. and and Ratner, H.H. 1993. Cultural learning. Behavioral and Brain Sciences 16, 495-552. Trinkaus, E. and Thompson, D.D. 1987. Femoral diaphyseal and histomorphometric age determinations for the Shanidar 3, 4, 5 and 6 Neandertals and Neandertal longevity. American Journal of Physical Anthropology 72, 1239.

153

Tuffreau, A. and Antoine, P. 1995. The earliest occupation of Europe: continental north western Europe. In W. Roebroeks and T. van Kolfschoten (Eds) The Earliest Occupation of Europe, pp 147-63. Leiden, University of Leiden. Turner, E. (Ed) 1995. Miesenheim I. Excavations at a Lower Palaeolithic Site in the Central Rhineland of Germany. Mainz, Monogr. Römisch-Germanisches Zentralmuseums Mainz. Turner, J.H. and Maryanski, A. 1991. Network analysis. In J.H. Turner (Ed) The Structure of Sociological Theory, pp 540-72. Belmont, Wadsworth. Turq, A. 1988. Le Paléolithique inférieur et moyen en Haut-Agenais: état des recherches. Revue de l’Agenais 115, 83112. Turq, A. 1999. Reflections on the Middle Palaeolithic of the Aquitaine Basin. In W. Roebroeks and C.S. Gamble (Eds) The Middle Palaeolithic Occupation of Europe, pp 107-20. Leiden, University of Leiden. Tylor, E.B. 1871. Primitive Culture. London, John Murray. Tzedakis, P.C. 1993. Long-term tree populations in northwest Greece through multiple Quaternary climatic cycles. Nature 364, 437-40. Tzedakis, P.C. 1994. Vegetation change through glacial-interglacial cycles: A long pollen sequence perspective. Philosophical Transactions of the Royal Society of London B345, 403-32. Ulrix-Closset, M. 1973. Le Moustérien à retouche bifaciale de la Grotte du Docteur à Huccorgne (Province de Liège). Helinium XIII, 209-34. Ulrix-Closset, M. 1975. Le Paléolithique Moyen dans le Bassin Mosan en Belgique. Wettern, Edition Universa. Ulrix-Closset, M. 1995. Le Moustérien récent à pointes foliacées en Belgique. Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, 201-5. Ulrix-Closset, M., Otte, M. and Cattelain, P. 1988. Le “Trou de l’Abime”, Couvin (Province de Namur, Belgique). In M. Otte (Ed) L’Homme de Néandertal Vol. 8, pp 225-39. Liège, ERAUL 35. Valladas, H., Geneste, J.-M., Joron, J.-L. and Chadelle, J.-P. 1986. Thermoluminescence dating of Le Moustier (Dordogne, France). Nature 322, 452-54. Valoch, K. 1960. K otázce Předmagdalénského osídlení jeskyní Adlerovy a Křížovy na Říčách u Brna. Časopis Moravského Muzea 45, 5-20. Valoch, K. 1965a. Altsteinzeitliche Funde aus Brno und Umgebung. Časopis Moravského Muzea 50, 21-30. Valoch, K. 1965b. Die Altsteinzeitlichen Gegehungen der Höhle Pod hradem. Anthropos 18, 93-106. Valoch, K. 1966. Die altertümlichen Blattspitzenindustrien von Jezeřany (Südmähren). Časopis Moravského Muzea 51, 5-60. Valoch, K. 1968a. Evolution of the palaeolithic in central and eastern Europe. Current Anthropology 9, 351-68. Valoch, K. 1968b. Eine jungpaläolithische Station in Brno-Kohoutovice. Sborník Prací Filosofické Fakulty Brněnské University 17, 63-80. Valoch, K. 1968c. Gisement du Pléistocène supérieur à Mamaia sur la côte de la Mer Noire. La Préhistoire, pp 465-72. Paris, CNRS. Valoch, K. 1970. Die Bedeutung paläolithische Funde für die Stratigraphie des Pleistozäns. Časopis pro Mineralogii a Geoloii 15, 261-9.

154

Valoch, K. 1973. Neslovice, eine bedeutende Oberflächenfundstelle des Szeletiens in Mähren. Časopis Moravského Muzea 58, 5-76. Valoch, K. 1976a. Neue mittelpaläolithische Industrien in Südmähren. Anthropologie 14, 15-64. Valoch, K. 1976b. Die altsteinzeitliche Fundstelle in Brno-Bohunice. Prague, Academia. Valoch, K. 1977. Neue alt- und mittel paläolithische Funde aus der Umgebung von Brno. Anthropozoikum 11, 93-112. Valoch, K. 1981. Beitrag zur Kenntnis des Pavloviens. Archeologické Rozhledy 33, 279-98. Valoch, K. 1982. Neue paläolithische Funde von Brno-Bohunice. Časopis Moravského Muzea 67, 31-48. Valoch, K. (Ed) 1988. Die Erforschung der Kůlna Höhle 1961-1976. Anthropos 24. Brno, Moravské Muzeum. Valoch, K. 1990a. Le Szélétien en Moravie. In J.K. Kozlowski (Ed) Feuilles de Pierre, pp 213-221. Liège, ERAUL 42. Valoch, K. 1990. La Morave il y à 40,000 ans. In C. Farizy (Ed) Paléolithique Moyen Récent et Paléolithique Supérieur Ancien en Europe, pp 115-124. Nemours, Mémoires du Musée de Préhistoire d’Ile de France No 3, Valoch, K. 1995. La Variabilité Typologique du Paléolithique Moyen de la Grotte de Kůlna en Moravie. Paleo Supplément No.1, Actes du Colloque de Miskolc, Les Industries à Pointes Foliacées d’Europe Centrale, 73-77. Valoch, K. and Zeman, A. 1979. Mušov. In V. Šibrava (Ed) Quaternary Glaciations in the Northern hemisphere, 6th Session, pp 62-4. Prague, Ostrava, Guide to Excursions. Velichko, A.A. 1984. Introduction. In A.A. Velichko (Ed) Late Quaternary Environments in the Soviet Union, pp xxiiixxvii. Trans. H.E. Wright Jr. and C.W. Barnosky. Minneapolis, University of Minnesota Press. Vereshchagin, N.K. 1977. Berelekh “cemetery” of mammoths. In O.A. Skarlato (Ed) Mammoth Fauna of the Russian Plain and Eastern Siberia, pp 5-50. Leningrad, Nauka. Vértes, L. 1951. Mezoliticheskie nakhodki na vershine gory Kőporos pri g. Eger. Acta Archaeologica Academiae Scientiarum Hungaricae 1, 153-90. Vértes, L. 1955a. Paläolithischen Kulturen des Würm I/II-Interstadials in Ungarn. Acta Archaeologica Academiae Scientiarum Hungaricae V, 261-77. Vértes, L. 1955b. Neuere Ausgrabungen und Paläolithische Funde in der Höhle von Istállóskő. Acta Archaeologica Academiae Scientiarum Hungaricae V, 111-31. Vértes, L. 1959. Das Moustérien in Ungarn. Eiszeitalter und Gegenwart 10, 21-40. Vértes, L. 1962-3. Einige Angaben des ungarischen Szeletiens. Arheološki Vestnik 8-14, 167-95. Vértes, L. 1964. Tata, eine Mittelpaläolithische Travertinsiedlung in Ungarn. Budapest. Vértes, L. 1965. Az őskőkor és az átmeneti kőkor emlékei magyarorszáagon. Budapest. Vértes, L. 1968. Szeleta-Symposium in Ungarn, 4-11 September 1966. Quartär 19, 381-90. Visalberghi, E. 1993. Capuchin monkeys: A window into tool use in apes and humans. In K.R. Gibson and T. Ingold (Eds) Tools, Language and Cognition in Human Evolution, pp 138-50. Cambridge, Cambridge University Press. Vishnyatsky, L.B. 1999. The Paleolithic of Central Asia. Journal of World Prehistory 13, 69-122. Vita-Finzi, C. 1978. Archaeological Sites in their Setting. London, Thames and Hudson. Vita-Finzi, C., Higgs, E.S., Sturdy, D.A., Harris, J., Legge, A.J. and Tippett, H. 1970. Prehistoric economy in the Mount Carmel area of Palestine: Site catchment analysis. Proceedings of the Prehistoric Society 36, 1-37.

155

van Vliet-Lanoë, B. 1986. Micromorphology. In P. Callow and J.M. Cornford (Eds) La Cotte de St. Brelade 1961-1978: Excavations of C.B.M. McBurney, pp 91-96. Norwich, Geo. van Vliet-Lanoë, B. 1989. Dynamics and extent of the Weichselian permafrost in Western Europe (Substage 5e to Stage 1). Quaternary International 3-4, 109-13. Vogel, J.C. and Waterbolk, H.T. 1972. Groningen Radio-carbon Dates X. Radiocarbon 14.1. de Waal, F. 1982. Chimpanzee Politics. London, Cape. de Waal, F. 1991. Peacemaking among Primates. London, Penguin. Watson, P., LeBlanc, S. and Redman, C. 1971. Explanation in Archaeology: An Explicitly Scientific Approach. New York, Columbia University Press. Watt, A.S. 1947. Pattern and process in the plant community. Journal of Ecology 35, 1-22. Watts, W.A. 1985. A long pollen record from Laghi di Monticchio, southern Italy: a preliminary account. Journal of the Geological Society 142, 491-9. Watts, W.A., Allen, J.R.M. and Huntley, B. 1996. Vegetation History and Palaeoclimate of the last glacial period at Lago Grande di Monticchio, Southern Italy. Quaternary Science Reviews 15, 133-53. Weber, T. 1977. Clactonienfunde von Memleben, kr. Nebra. Ausgrabungen u. Funde 22, 195-99. Welten, M. 1981. Verdrängung und Vernichtung der anspruchsvollen Gehölze am Beginn der letzten Eiszeit und die Korrelation der Frühwürm-Interstadiale in Mittel- und Nordeuropa. Eiszeitalter und Gegenwart 31, 182-202. Welten, M. 1982. Pollenanalytische Untersuchungen im Jüngeren Quartär des nördlichen Alpenvorlandes der Schweiz. Beiträge zur Geologischen Karte der Schweiz NF 156. Bern. Weiβmüller, W. 1995a. Sesselfelsgrotte II. Die Silexartefakte der Unteren Schichten der Sesselfelsgrotte. Ein Betrag zum Problem des Moustérien. Saarbrücken, Quartär-Bibliothek Band 6. Weiβmüller, W. 1995b. Drei Fundstellen mit Blattformen aus dem südostbayern Donauraum. Quartär 45/46, 99-134. Weiβmüller, W. 1996. Die Hornsteinlagerstätte von Flintsbach-Hardt: Geologie und Archäologie. Archäologische Denkmäler im Landkreis Deggendorf, Heft 10. West, R.G. 1977. Flora and Fauna. Early and Middle Devensian flora and vegetation. Philosophical Transactions of the Royal Society of London B280, 229-46. Wetzel, R. 1958. Die Bocksteinschmiede. Teil I. Stuttgart. Whallon, R. 1989. Elements of cultural change in the later Paleolithic. In P.A. Mellars and C.B. Stringer (Eds) The Human Revolution: Behavioural and Biological Perspectives on the Origins of Modern Humans, pp 432-54. Edinburgh, Edinburgh University Press. Wiessner, P. 1983. Style and social information in Kalahari San projectile points. American Antiquity 48, 253-76. White, L.A. 1949. The Science of Culture. New York, Farrar Straus. White, L.A. 1959. The Evolution of Culture. New York, McGraw-Hill. White, M.J. 1998. On the Significance of Acheulean Biface Variability in Southern Britain. Proceedings of the Prehistoric Society 64, 15-44. Wijmstra, T.A. 1969. Palynology of the first 30 metres of a 120m deep section in Northern Greece. Acta Botanica Neerlandica 18, 311-27.

156

Willis, K.J. 1994. The vegetational history of the Balkans. Quaternary Science Reviews 13, 769-88. Woillard, G.M. 1978. Grande Pile peat bog: a continuous pollen record for the last 140,000 years. Quaternary Research 9, 1-21. Woillard, G.M. 1979. The last interglacial-glacial cycle at Grande Pile in Northeastern France. Bulletin Société belge de Géologie 88, 51-69. Woodward, F.I. 1987. Climate and Plant Distribution. Cambridge, Cambridge University Press. Woodward, F.I. and Sheehy, J.E. 1983. Principles and Measurements in Environmental Biology. London, Butterworths. Würges, K. 1986. Artefakte aus den ältesten Quartär-Sedimenten (Schichten A-C) der Tongrube Kärlich, Kreis MayenKoblenz/Neuwieder Becken. Archäologische Korrespondenzblatt 16, 1-6. Würges, K. 1991. Neue altpaläolithische Funde aus der Tongrube Kärlich, Kreis Mayen-Koblenz/Neuwieder Becken. Archäologische Korrespondenzblatt 21, 449-55. Wymer, J. 1999. The Lower Palaeolithic Occupation of Britain, Vol. 1. Salisbury, Wessex Archaeology and English Heritage. Wynn, T. and Tierson, F. 1990. Regional Comparison of the Shapes of Later Acheulean Handaxes. American Anthropologist 92, 73-84. Zagwijn, W.H. 1961. Vegetation, climate and radiocarbon datings in the Late Pleistocene of the Netherlands. Part I: Eemian and Early Weichselian. Mededelingen Geologisches Stichting (NS) 4, 15-45. Zagwijn, W.H. 1974. Vegetation, climate and radiocarbon datings in the Late Pleistocene of The Netherlands, Part II: Middle Weichselian. Mededelingen Rijks Geologische Dienst NS 25, 101-11. Zagwijn, W.H. 1989. Vegetation and climate during warmer intervals in the Late Pleistocene of western and central Europe. Quaternary International 3/4, 57-67. Zagwijn, W.H. 1996. An analysis of Eemian climate in western and central Europe. Quaternary Science Reviews 15, 451-69. Zedler, P.H., Gautier, C.R. and McMaster, G.S. 1983. Vegetation change in response to extreme events: the effect of a short interval between fires in California chapparal and coastal shrub. Ecology 64, 809-18. Zilhão, J. and d’Errico, F. 1999a. Reply to Straus, Mellars and Otte. Current Anthropology 40, 355-64. Zilhão, J. and d’Errico, F. 1999b. The chronology and Taphonomy of the Earliest Aurignacian and its Implications for the Understanding of Neandertal Extinction. Journal of World Prehistory 13, 1-68. Zotz, L. 1951. Altsteinzeitkunde Mitteleuropas. Stuttgart. Zotz, L. 1955. Das Paläolithikum in den Weinberhöhlen bei Mauern. Bonn, Ludwig Röhrscheid. Zotz, L. 1959. Kösten, ein Werkplatz des Praesolutréen in Oberfranken. Bonn, Ludwig Röhrscheid.

157

158

APPENDIX 1

POLLEN DIAGRAMS

159

Figure A1. 1. Watten. (After Emontspohl 1995, Figure 3.)

Figure A1.2.a. Scladina Coupe 6-7, Carré F. (After Bastin 1992, Figure 1.)

Figure A1.2.b. Scladina Coupe C-D, Carré 4 (below) and Coupe F-G, Carré 8 (above). (After Bastin 1992, Figure 2.)

160

Figure A1.3. La Grande Pile XX. (After de Beaulieu and Reille 1992, Figure 1.)

161

Figure A1.4.a. Les Echets G, 14.5-29m depth (‘Pleniwürm’). (After de Beaulieu and Reille 1984, Figure 4.)

162

Figure A1.4.b. Les Echets G, 29-34m depth (early Würm). (After de Beaulieu and Reille 1984, Figure 3.)

163

Figure A1.4.c. Les Echets G, 34-39m depth (late Riss and Eemian). (After de Beaulieu and Reille 1984, Figure 2.)

164

Figure A1.5. Ioannina 249 AP/NAP% (above), arboreal taxa % (below). (After Tzedakis 1994, Figures 4(b) and 5.)

165

Figure A1.6.a. Padul 2. (After Pons and Reille 1988, Figure 3.)

Figure A1.6.b. Padul 3. (After Pons and Reille 1988, Figure 2.)

166

Figure A1.7. Gołkow. (After Mamakowa 1989, Figure 35.)

Figure A1.8. Zgierz-Rudunki. (After Mamakowa 1989, Figure 35.)

Figure A1.9. Kąty, Profile II. (After Mamakowa et al 1975, Fig. 5.)

167

Figure A1. 10.a. Samerberg ‘Kernbohrungsprofil’, 4.2 -12.8m depth. (After Grüger 1979, Beilage 3.)

168

Figure A1.10.b. Samerberg ‘Kernbohrungsprofil’, 12.2 - 19.6m depth. (After Grüger 1979, Beilage 3.)

169

Figure A1.10.c. Samerberg ‘Kernbohrungsprofil’, 19.0 – 22.4m depth. (After Grüger 1979, Beilage 3.)

170

Figure A1.11.a. Oerel OE61, 16.3 – 19.0m (Eemian and earliest Weichsel). (After Behre and Lade 1986, Tafel 1.)

Figure A1.11.b. Oerel OE61, 13.3 – 15.3m (‘Herning’ stadial and Brørup interstadial). (After Behre and Lade 1986, Tafel 2.)

171

Figure A1.11.c. Oerel OE61, 5.20 – 7.60m (‘Rederstall’ stadial and Odderade interstadial). (After Behre and Lade 1986, Tafel 2.)

Figure A1.11.d. Oerel OE61: Above, 2.70 – 3.00m (Glinde interstadial); below, 3.95 – 4.73m (Oerel interstadial). (After Behre and Lade 1986, Tafel 2.)

172

Figure A1.12.a. Lac Du Bouchet D, Lower Part. (After Reille and de Beaulieu 1990, Figure 3.)

173

Figure A1.12.b. Lac du Bouchet D, Upper Part. (After Reille and de Beaulieu 1990, Figure 2.)

174

Figure A1.13. Machnacz MII. (After Kupryjanowicz 1991, Figure 3.)

Figure A1.14. Kittlitz. (After Erd 1973, Abb 8.)

175

Figure A1.15. Sulzberg. (After Welten 1982, Dia. 44.)

176

APPENDIX 2

ILLUSTRATIONS OF STONE ARTEFACTS

177

Figure A2.1. 1,2,3 and 4, bifacially retouched tools from Blanzy; 1 and 3, ventrally thinned tools; 2, backed biface knife; 4, planoconvex bifacial scraper. 5 and 6, biface prondniks from Germolles. Pieces shown to scale. (After Farizy 1995, Fig. 4.)

178

Figure A2.2. Bifacially retouched tools from Frettes; 1 can be classified as a leafpoint. Pieces shown to scale. (After Farizy 1995, Figure 5.)

179

Figure A2.3. Bifacially retouched tools from Trou de l’Abime, Couvin. Pieces shown actual size. Illustrations by T. Hopkinson.

180

i.

ii.

iii.

Figure A2.4. Couvin: i. ventrally retouched dejeté scraper; ii. limace on broad flake with partial ventral thinning; iii: distal fragment of a unifacial Mousterian Point with dorsal surface retouch. Pieces shown actual size. Illustrations by T. Hopkinson, except ii. after Ulrix-Closset 1995, Planche 2.2.

181

Figure A2. 5. Foliate bifaces from Spy, ‘niveau moyen’. Pieces shown actual size. (After Ulrix-Closset 1975, Figs. 132 and 133.)

182

Figure A2.6. Foliate bifaces from Spy, ‘niveau moyen’. Pieces shown actual size. (After Ulrix-Closset 1975, Figs. 134 and 129.)

183

Figure A2.7. Foliate bifaces from Spy, ‘niveau moyen’. Pieces shown actual size. (After Ulrix-Closset 1975, Figs. 131 and 130.)

184

Figure A2.8. Foliate bifaces from Goyet. Pieces shown actual size. (After Ulrix-Closset 1975, Figs. 192, 190, 186.)

185

Figure A2.9. Bifaces from Grotte du Docteur. Pieces shown to scale. (After Ulrix-Closset 1975, Figs. 300, 304, 301, 302,303.)

186

Figure A2.10. Trou du Diable. Piece shown actual size. (After Ulrix-Closset 1975, Fig. 29.)

Figure A2.11. Grottes d’Engis. Pieces shown actual size. (After Ulrix-Closset 1975, Fig. 409.)

187

Figure A2.12. Engihoul. Pieces shown actual size. (After Ulrix-Closset 1975, Figs. 377, 378.)

188

Figure A2.13. Grotte du Mont Falhize. Piece shown actual size. (After Ulrix-Closset 1975, Fig. 433.)

Figure A2.14. Liège-Sainte-Walburge. Piece shown actual size. (After Ulrix-Closset 1975, Fig. 454.)

189

Figure A2.15. Salzgitter-Lebenstedt. Pieces shown to scale. (After Allsworth-Jones 1986, Fig. 4.)

190

Figure A2.16. Bifacial leaf point, Ranis. Piece shown actual size. (After Allsworth-Jones 1986, Fig. 6.)

191

i.

ii.

iii.

Figure A2.17. Points from Ranis. i. bifacial leaf point; ii., iii. Jerzmanowice points. Pieces shown to scale. (After Allsworth-Jones 1986, Fig 7.2, 3, 1.)

192

i.

ii.

Figure A2.18. Leaf points from surface finds in northern Germany. i. Osterwald (after Bosinski 1967, Taf. 125.12); ii. Graste (after Bosinski 1967, Taf 127.8). Pieces shown actual size.

193

Figure A2.19. Leaf point fragment from Fritzlar. Piece shown actual size. (After Bosinski 1967, Taf. 96.5.)

194

Figure A2.20. Leaf point fragments from Rörshain. Pieces shown to scale. (After Bosinski 1967, Taf. 104.)

195

Figure A2.21. Handaxes from Rörshain. Pieces shown to scale. (After Bosinski 1967, Taf 102.)

196

Figure A2.22. Leaf points from Kösten. Pieces shown to scale. (After Zotz 1959, Figs. 73a, 75, 80, 84.)

197

Figure A2.23. Bifacial leaf points from Mauern, Bohmers’ excavation. Pieces shown actual size. (After Bohmers 1951, Taf. 25.)

198

i.

ii.

Figure A2.24. Leaf points from Mauern, Bohmers’ excavation: i. bifacial fragment; ii. unifacial. Pieces shown actual size. (After Bohmers 1951, Taf. 29.2 and Taf 27.2.)

199

i.

ii.

Figure A2.25. Bifaces from Mauern, Bohmers’ excavation. i. Handaxe; ii. Faustkeil. Pieces shown to scale. (After Bohmers 1951, 35/36.1, Taf. 37.1.)

200

Figure A2.26. Bifacial leaf points from Mauern, Layer F1 of Zotz’s excavation. Pieces shown to scale. (After Zotz 1955, Bild 43, 46.)

201

i

ii.

Figure A2.27. Bifacial leaf points from Mauern, Layer G of Zotz’s excavation: i. bifacial leaf point; ii. unifacial leaf point. Pieces shown actual size. (After Zotz 1955, Bild 49, 47.4.)

202

i.

ii.

Figure A2.28. Possible Altmühlian leaf points from i. Mörnsheim and ii. Biesenhard. Pieces shown actual size. (After Bosinski 1967, Taf 123.1, 127.12.)

i.

ii.

iii. Figure A2.29. Leaf points from Kleine Ofnet. i. unifacial; ii., iii. bifacial. Pieces shown to scale. (After Bosinski 1967, Taf. 126.1, 4, 6.)

203

Figure A2.30. Leaf points from Haldenstein. Pieces shown to scale. (After Bosinski 1967, Taf. 123.6, 7.)

Figure A2.31. Handaxes from the Bocksteinschmiede. Pieces shown to scale. (After Bosinski 1969b, Taf 20.1 and 23.2.)

204

i.

ii.

Figure A2.32. Bifaces from the Bocksteinschmiede. i. Faustkeilblatt; ii. biface knife. Pieces shown actual size. (After Bosinski 1969b, Taf. 63.6 and 84.6.)

205

Figure A2.33. Leaf point-like bifaces from the Bocksteinschmiede. Pieces shown to scale. (After Bosinski 1969b, Taf. 93.3, 5; Taf 94.3.)

206

i.

ii.

Figure A2.34. Leaf point fragments from the Klausen Caves; i. Obere Klause; ii. Mittlere Klause. Pieces shown actual size. (After Bosinski 1967, Taf. 124.3, 8.)

207

Figure A2.35. Leaf points from Obernederhöhle. Pieces shown to scale. (After Freund 1968, Abb. 29.6, 35.1, 36.1, 52.2.)

208

i.

ii.

Figure A2.36. Bifaces from Obernederhöhle; i. handaxe; ii. Keilmesser. Pieces shown to scale. (After Freund 1968, Abb 37, 39.)

209

i.

ii.

iii.

Figure A2.37. Hohlen Stein, Schambach; i., ii.; leaf point-like artefacts; iii. leaf point fragment. Pieces shown actual size. (After Bosinski 1967, Taf 89.2, 5; Taf 91.1.)

210

Figure A2.38. Leaf points from Zeitlarn. Pieces shown to scale. (After Schönweiβ and Werner 1986, Abb. 4.)

211

Figure A2.39. Leaf points and Blattformen from Albersdorf. Pieces shown actual size. (After Weiβmüller 1995b, Abb. 15.)

212

Figure A2.40. Bifaces from Flintsbach-Hardt. Pieces shown actual size. (After Weiβmüller 1995b, Abb.24.)

213

i.

ii.

Figure A2.41. Jerzmanowice points from i. Buchberghöhle and ii. Pottenstein. Pieces shown actual size. (After Bosinski 1967, Taf. 125.7 and 127.11.)

214

Figure A2.42. Leaf point-like biface from Bohunice. Piece shown actual size. (After Allsworth-Jones 1986, Fig. 33.1.)

Figure A2.43. Bifacial leaf points from Líšeň. Pieces shown to scale. (After Svoboda 1990, Fig. 7.)

215

Figure A2.44. Jerzmanowice points from Líšeň. Pieces shown to scale. (After Svoboda 1990, Fig. 8.)

216

Figure A2.45: Small handaxes from Kůlna. Pieces shown actual size. (After Svoboda et al 1996, Fig. 4.4.4, 8.)

Figure A2.46. leaf point-like bifaces from Kůlna. Pieces shown to scale. (After Valoch 1995, Figs. 3, 4.)

217

Figure A2.47. Jerzmanowice points from Kent’s Cavern. Pieces shown actual size. (After Jacobi 1990, Fig. 1.)

218

Figure A2.48. Bifacial leaf points from Neslovice. Pieces shown actual size. (After Allsworth-Jones 1986, Fig. 41.1, 2.)

Figure A2.49. Bifacial leaf points from Vedrovice V. Pieces shown actual size. (After Valoch 1990b as reproduced in Mellars 1996 Fig. 13.16.)

219

Figure A2.50. Jerzmanowice points, Nietoperzowa. Pieces shown to scale. (After Allsworth-Jones 1986, Fig. 42.)

220

Figure A2.51. Bifacial leaf points, Mamutowa. Pieces shown actual size. (After Allsworth-Jones 1986, Fig. 47.1, 3.)

221

i.

ii. Figure A2.52. Zwierzyniec; i. bifacial leaf point fragment; ii. Upper Palaeolithic tools. Pieces shown to scale. (After Allsworth-Jones 1986, Figs. 35.1 and 36.)

222

i.

ii.

Figure A2.53. Moravany-Dlha; i. bifacial leaf point; ii. unifacial leaf point. Pieces shown actual size. (After Allsworth-Jones 1986, Fig. 37.4, 1.)

223

Figure A2.54. Bifacial leaf points, Jankovich. Pieces shown to scale. (After Allsworth-Jones 1986, Fig. 24.)

224

Figure A2.55. Bifacial leaf points, Szeleta Cave Lower Industry. Pieces shown to scale. (After Allsworth-Jones 1986, Figs 12.1, 4 and 17.1, 6.)

225

Figure A2,56. Bifaces, Subalyuk. Pieces shown actual size. (After Allsworth-Jones 1986, Fig. 21.1, 2.)

226

i.

ii.

Figure A2.57. Korolevo Va; i. foliate bifaces; ii. leaf points. Pieces shown to scale. (After Gladilin et al 1995, Figs. 8, 9.)

227

ii. i.

iii.

Figure A2.58. Foliate bifaces from i. Bouzdoujany (after Anissutkine 1990, Fig.2); ii. Rikhta (after Kulakovskaya 1990, Fig. 2.1); iii. Velikij Glubotchok I (after Bogutskij et al 1999). All pieces shown to scale.

228

Figure A2.59. Foliate bifaces, Stinka I Lower; the piece at right centre is a leaf point. Pieces shown to scale. (After Anissutkine 1990, Fig. 1.)

229

Figure A2.60. Leaf points, Stinka I Upper. Pieces shown to scale. (After Anissutkine 1990, Fig. 3.)

Figure A2.61. Foliate bifaces, Zhitomir. Pieces shown to scale. (After Kulakovskaya 1990, Fig. 2.2, 3.)

230

i.

ii.

Figure A2.62. Ripiceni-Izvor. i. flat handaxes (after Păunescu 1993, Fig. 70.2, 4); ii. bifacial scrapers (after Păunescu 1993, Fig. 69.8, 9). Pieces shown to scale.

231

Figure A2.63. Leaf points, Ripiceni Izvor Level IV. Pieces shown to scale. (After Păunescu 1993, Fig. 74.)

232

Figure A2.64. Foliate bifaces, Ripiceni-Izvor Level V. Pieces shown to scale. (After Păunescu 1993, Fig. 79.)

233

Figure A2.65. Leaf points, Musselievo. Pieces shown to scale. (After Sirakova 1990, Plate II.)

234

Figure A2.66. Leaf points, Musselievo. Pieces shown to scale. (After Sirakova 1990, Plate IV.)

235

Figure A2.67. Leaf points, Musselievo. Pieces shown to scale. (After Sirakova 1990, Plate V.)

236

Figure A2.68. Leaf points, Samuilitsa. (After Allsworth-Jones 1986, Fig. 8.)

237

Figure A2.69. Leaf points, Kokkinopilos. Pieces shown actual size. (After Gamble 1986, Fig. 5.10.b, c.)

Figure A2.70. Leaf point, Morfi. Piece shown actual size. (After Gamble 1986, Fig. 5.10.d.)

238

Figure A2.71. Foliates, Starosele. The piece on the upper right is a unifacial leaf point. Pieces shown to scale. (After Kolosov 1990, Plate 5.)

239

Figure A2.72. Foliate handaxes, Starosele. Pieces shown to scale. (After Marks and Monegal 1998, Fig. 7.17.c-f.)

240

Figure A2.73. Leaf points, Zaskalnaya VI. Pieces shown to scale. (After Kolosov 1990, Plate 1.)

241

Figure A2.74. Biface knives, Zaskalnaya V and VI. Pieces shown to scale. (After Kolosov 1995, Fig. 2.)

242

Figure A2.75. Bifaces, Kabazi II. Pieces shown to scale. (After Marks and Chabai 1998, Fig. 10-5.)

243

Figure A2.76. Jerzmanowice points and pointes à face plane, Kostienki-Tel’manskaya. (After Allsworth-Jones 1986, Fig. 45.)

244

Figure A2.77. Bifaces, Mezmaiskaya. Pieces shown to scale. (After Golovanova et al 1999.)

245

246

APPENDIX 3

TABLES OF SITES

247

SITE

TYPOLOGICAL LEAF POINTS?

PROPOSED DATE

REFERENCE(S)

ILLUSTRATION

Saint-Julien de la Liègue

NO

Unknown

Cliquet 1982, 1995

-

Villegats

NO

Unknown

Cliquet 1995

-

Champlost

NO

OIS 4/3: 65-45 Kya (ESR)

Farizy 1985, 1995

-

Bissy-sur-Fley

YES

OIS 5 or 3

Desbrosse and Texier 1973a

-

Blanzy

YES

OIS 5 or 3

Desbrosse and Tavoso 1970

Fig. A.2.1

Germolles

YES

OIS 5 or 3

Desbrosse and Texier 1973b

Fig. A.2.1

Frettes

YES

OIS 5 or 3

Huguenin 1988

Fig. A2.2

Goyet

NO

OIS 5 or 3

Ulrix-Closset 1975

Fig. A2.8

Trou du Diable

NO

unknown

Ulrix-Closset 1975

Fig. A2.10

Engihoul

NO

unknown

''

Fig. A2.12

Grotte du Mont Falhize

NO

unknown

''

Fig. A2.13

Liège-SainteWalburge

YES

unknown

''

Fig A2.14

Trou Magrite

NO

unknown

''

-

Ramioulle

NO

OIS 5 or 3

''

-

Trou de Sureau

NO

unknown

Ulrix-Closset 1968, 1975

-

Grotte de l’Hermitage

NO

unknown

Ulrix-Closset 1975

-

SalzgitterLebenstedt, Layer L

NO

OIS 6, 5 or 3 C14 dates 55,600±900, 43,000 BP

Fritzlar

YES

Dickershausen

N. FRANCE

BELGIUM

N. & C. GERMANY Tode et al 1953; Bosinski 1963; Grote 1978

Fig. A2.15

Unknown

Bosinski 1967

Fig. A2.19

NO

Unknown

''

-

Kirchberg

NO

Unknown

''

-

Maden

NO

Unknown

''

-

Wichdorf

NO

Unknown

''

-

Untere Klause

NO

unknown

Mittlere Klause

YES

OIS 5 or 3

''

Fig. A2.34.ii

Obere Klause

YES

OIS 5 or 3

''

Fig. A2.34.i

Klausennische

NO

OIS 5 or 3

''

-

Hohler Stein, Schambach

NO

OIS 5 or 3

S. GERMANY Bosinski 1967; Allsworth-Jones 1986

Taute 1966; Bosinski 1967

-

Fig. A2.37

Table A3.1. Sites with Middle Palaeolithic bifaces, some of which grade toward or into leaf point form (see continuations).

248

SITE

TYPOLOGICAL LEAF POINTS?

PROPOSED DATE

REFERENCE(S)

ILLUSTRATION

Breitenfurter Höhle

YES

Unknown

Bosinski 1967

-

Sesselfelsgrotte, Level G

NO

Early OIS 3

Richter 1997

-

Ve Vratech

NO

Unknown

Fridrich and Sklenář 1976

-

Sloupová

NO

Unknown

''

-

Chlupáčova sluj

NO

Unknown

''

-

Pekárna

NO

Unknown

Svoboda 1991; Svoboda et al 1996

-

Drátenická

NO

Unknown

Svoboda et al 1996

-

Výpustek

NO

Unknown

Valoch 1965a; Svoboda et al 1996

-

Predmostí

NO

OIS 4

Valoch 1981; Svoboda et al 1996

-

Ciemna

NO

OIS 5 or 3

Chmielewski 1969; Allsworth-Jones 1986

-

Piekary

NO

OIS 5 or 3

NO

Unknown

Valoch 1970

Subalyuk

NO

OIS 5 or 3

Kretzoi 1968; Allsworth-Jones 1986

Fig. A2.56

Tata

NO

OIS 5: 116-78 Kya (U-Th)

Vértes 1964; Allsworth-Jones 1986

-

Büdöspest

YES

OIS 4/3

Gábori-Csank 1970; Allsworth-Jones 1986

-

Rikhta

YES

OIS 5 or 3

Zhitomir

YES

Unknown

Stinka 1 Lower and Upper

YES

Lower OIS 4/3; Upper OIS 3

Antonovka II

YES

Unknown

Kulakovskaya 1990

-

Velikij Glubotchok I

YES

Younger than 212 kya (magnetic susceptibility)

Bogutskij et al 1999

Fig. A2.58.iii

S. GERMANY CONT.

BOHEMIA

MORAVIA

S. POLAND

''

-

SLOVAKIA Zamarovce

-

HUNGARY

UKRAINE Kulakovskaya 1990 '' Anissutkine 1990

Table A3.1 continued.

249

Fig. A2.58.ii Fig. A2.61 Fig. A2.59, 60

SITE

TYPOLOGICAL LEAF POINTS?

PROPOSED DATE

REFERENCE(S)

ILLUSTRATION

Boutechty

YES

OIS 4/3

Bouzdoujany I

YES

OIS4/3

''

Fig. A2.58.i

Mitoc La Sărături

YES

OIS 3

Nicolăescu-Plopşor et al 1959; Allsworth-Jones 1986

-

Mitoc Valea Izvorului

YES

OIS 3

Bitiri 1967; Allsworth-Jones 1986

-

Remetea-Şomoş

NO

Unknown

Boineşti

NO

Unknown

Nicolăuescu-Plopşor and Covacs 1959; Allsworth-Jones 1986

-

Ocna-Sibiului

NO

Unknown

Rosu 1966; Allsworth-Jones 1986

-

Kiik-Koba

NO

Unknown

Klein 1969

-

Kabazi II

YES

OIS 4/3 70-45,000 BP (C14)

Khotylevo

YES

Unknown

UKRAINE CONT. Anissutkine 1990

-

ROMANIA

''

-

CRIMEA

Marks and Chabai 1998; Rink et al 1998

Fig. A2.75

Golovanova et al 1999

Fig. A2.77

S.W. RUSSIA

Kulakovskaya 1990

Table A3.1 continued.

250

-

SITE

PROPOSED DATE

REFERENCE(S)

ILLUSTRATION

S. GERMANY Groβe Ofnet

Unknown

Freund 1952; Bosinski 1967

-

Kleine Ofnet

Unknown

''

Fig. A2.29

Steinerner Rosenkrantz

Unknown

''

Fig. A2.28.i

Biesenhard

Unknown

''

Fig. A2.28.ii

Haldenstein

OIS 3

Bosinski 1967; Riek 1938

Fig. A2.30

Zeitlarn

Unknown

Schönweiβ and Werner 1986

Fig. A2.38

Albersdorf

Unknown

Weiβmüller 1995b

Fig. A2.39

Flintsbach-Hardt

Unknown

Weiβmüller 1995b

Fig. A2.40

OIS 4 or 3

Gábori-Csánk 1983; Allsworth-Jones 1990

-

OIS 5 or 3

Valoch 1968c; Allsworth-Jones 1986

-

Unknown

Gavela 1971; Allsworth-Jones 1986

-

Uncertain; OIS 5-3?

Dakaris et al 1964: Higgs 1965; Higgs and Vita-Finzi 1966; Bailey et al 1992; van Andel 1998

Fig. A2.69

Gamble 1986

Fig. A2.70

HUNGARY Máriaremete

ROMANIA Mamaia SERBIA Risovača GREECE Kokkinopilos β

Morfi Galatas

''

Aiya

''

CRIMEA Ak-Kaya III

Uncertain; OIS 3?

Sary-Kaya

Uncertain; OIS 3?

Prolom II

Uncertain; OIS 3?

Kolosov 1977; Chabai 1998 ''

-

Marks and Chabai 1998

-

Klein 1969

-

S.W. RUSSIA Il’skaya

OIS 3: 37,200±1800 BP, 40,800±1200 BP (C14)

Table A3.2. Sites with unequivocal Middle Palaeolithic leaf points that are formally distinct from other biface types present.

251

SITE

>5 LEAF POINTS?

JERZMANOWICE POINTS PRESENT?

REFERENCE(S)

ILLUSTRATION

Lužná-Hlaváčov

YES

NO

Lužná-Krásná Dolina

YES

NO

''

-

Ondratice

NO

YES

Svoboda 1990

-

Líšeň

YES

YES

Svoboda 1990

Figs. A2.43, 44

Mohelno

YES

NO

Svoboda et al 1996

-

Pod hradem

YES

NO

Valoch 1965b

-

Dzeravá Skalá

YES

NO

Prošek 1951

-

Vlčkovce

NO

NO

Bárta 1967

-

Vel'ký Šariš

NO

NO

Bánesz 1960

-

Kechnec

YES

NO

Bánesz 1959, 1976

-

Tibava

NO

NO

Bánesz 1976

-

DiósgyőrTapolca

YES

NO

Saád and Hellebrandt 1974

-

Herman Ottó Rock Shelter

NO

NO

Vértes 1955a

-

Alsószentgyörgy

YES

NO

Vértes 1965

-

Eger Kőporostető

YES

NO

Vértes 1951

-

Hont-Csitár

YES

NO

V.Gábori 1958

-

SajóbábonyMéhészetö

YES

NO

Ringer 1983

-

SajóbábonyKövesoldal

YES

NO

''

-

Kánástetö

YES

NO

''

-

Szabadkatetö

YES

NO

''

-

Mályi-Öreghegy

YES

NO

''

-

BOHEMIA Fridrich 1973

-

MORAVIA

SLOVAKIA

HUNGARY

-

ROMANIA Valea Chichereului

YES

NO

Breuil 1925

-

Iosăşel

YES

NO

Nicolăescu-Plopşor and Bassa 1957

-

Kamen

NO

NO

M.Gábori 1976; Ivanova 1979

-

Visoko Brdo

NO

NO

Gábori 1976

-

BOSNIA

Table A3.3. Sites with more than one leaf point but which might be either Middle or Upper Palaeolithic.

252

SITE

REFERENCE(S)

ILLUSTRATION

>5 LEAF POINTS?

JERZMANOWICE POINTS PRESENT?

Kent’s Cavern

YES

YES

Hyaena Den

NO

YES

''

-

Badger Hole

NO

YES

''

-

Uphill Quarry

NO

YES

''

-

Beedings

YES

YES

''

-

Paviland

YES

YES

''

-

Bramford Road

YES

YES

''

-

Robin Hood Cave

YES

YES

''

-

Soldiers Hole

NO

NO

''

-

YES

YES

Ulrix-Closset 1975

Neslovice

YES

NO

Valoch 1973

Fig. A2.48

Otaslavice

YES

NO

Absolon 1935

-

Kvasice Nový Dvůr

YES

NO

Janàsek and Skutil 1954

-

Jezeřany 1 and 2

YES

NO

Valoch 1966

-

Přestalvky

YES

NO

Klíma 1978

-

Předmostí

YES

NO

Valoch 1981

-

Kohoutovice

YES

NO

Valoch 1968b

-

Křepice

YES

NO

Klíma 1969

-

Vedrovice V

NO

NO

Valoch 1976a

Židlochovice

YES

NO

Oliva 1989

-

Ořechov 1 and 2

YES

NO

Valoch 1973

-

Želešice 1

YES

NO

Valoch 1973

-

Mamutowa

YES

YES

S. Kowalski 1969

Koziarnia

NO

YES

Chmielewski et al 1967

-

Dzierzysław 1

YES

NO

Kozlowski 1962

-

Zwierzyniec

NO

NO

Chmielewski et al 1977

Balla

YES

NO

Vértes 1962-3

-

Puskasporos

YES

NO

Vértes 1955a

-

YES

YES

Klein 1969; Allsworth- Jones 1986

-

GREAT BRITAIN: Campbell 1977, 1980; Jacobi 1990

Fig. A2.47

BELGIUM Goyet

-

MORAVIA

Fig. A2.49

POLAND Fig. A2.51

Fig. A2.52

HUNGARY

S.W. RUSSIA KostienkiStreletskaya

Table A3.4: Sites with more than one Early Upper Palaeolithic leaf point.

253