Unparalleled Behaviour: Britain and Ireland during the 'Mesolithic' and 'Neolithic' 9781841715360, 9781407319964

This extensive book is organised into three parts. Part one discusses the changing perspectives of the 'Mesolithic&

184 25 34MB

English Pages [717] Year 2003

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Unparalleled Behaviour: Britain and Ireland during the 'Mesolithic' and 'Neolithic'
 9781841715360, 9781407319964

Table of contents :
Front Cover
Title Page
Copyright
Dedication
Opening Quotations
Table of Contents
List of figures
List of tables
Acknowledgements
Abstract
Notes
Preface
PART I
INTRODUCTION: CONSTRUCTING THE PAST
CHAPTER ONE: DEFINING THE ‘NEOLITHIC’
CHAPTER TWO: ECONOMY AND SUBSISTENCE IN THE ‘MESOLITHIC’ AND ‘NEOLITHIC’
PART II
CHAPTER THREE: IDENTIFYING UNPARALLELED BEHAVIOUR
CHAPTER FOUR: MAKING MEGALITHS FROM STONES
CHAPTER FIVE: LIVING WITH THE DEAD, MAKING ANCESTORS
CHAPTER SIX: MOVING AND LIVING AROUND THE LANDSCAPE
PART III
CHAPTER SEVEN: THE AVEBURY REGION
CONCLUSION: UNPARALLELED HUMAN SOCIAL BEHAVIOUR
Appendix 1
Appendix 2
Appendix 3
Appendix 4
Appendix 5
Appendix 6
BIBLIOGRAPHY

Citation preview

BAR 355 2003

Unparalleled Behaviour: Britain and Ireland during the ‘Mesolithic’ and ‘Neolithic’ KING: UNPARALLELED BEHAVIOUR

Martin P. King

BAR British Series 355 2003 B A R

Published in 2016 by BAR Publishing, Oxford BAR British Series 355 Unparalleled Behaviour: Britain and Ireland during the 'Mesolithic' and 'Neolithic' © M P King and the Publisher 2003 The author's moral rights under the 1988 UK Copyright, Designs and Patents Act are hereby expressly asserted. All rights reserved. No part of this work may be copied, reproduced, stored, sold, distributed, scanned, saved in any form of digital format or transmitted in any form digitally, without the written permission of the Publisher.

ISBN 9781841715360 paperback ISBN 9781407319964 e-format DOI https://doi.org/10.30861/9781841715360 A catalogue record for this book is available from the British Library BAR Publishing is the trading name of British Archaeological Reports (Oxford) Ltd. British Archaeological Reports was first incorporated in 1974 to publish the BAR Series, International and British. In 1992 Hadrian Books Ltd became part of the BAR group. This volume was originally published by Archaeopress in conjunction with British Archaeological Reports (Oxford) Ltd / Hadrian Books Ltd, the Series principal publisher, in 2003. This present volume is published by BAR Publishing, 2016.

BAR

PUBLISHING BAR titles are available from:

E MAIL P HONE F AX

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

Dedicated to the work and memory of Vere Gordon Childe (1892-1957)

The past is a foreign country: they do things differently there. Leslie P. Hartley 1953:9

Binarisms are too stable – they tend to smooth over differences and contradictions and end up being no more than a copulation of clichés Ackbar M. Abbas 1996:215

As to the use of the cathedral by local townspeople, it is clear that the building was an integral part of everyday life – and that many of today’s divisions between ‘religious’ and ‘secular’ activities had not yet been invented. Inside the walls of Chartres, people would gossip, exercise their dogs, parakeets and falcons, play ball games, go courting, shelter from the rain, and buy wine and food from merchants who had set up shop on prime sites along the nave; the merchants were only prepared to move when the chapter offered them another equivalent ‘pitch’, this time in the crypt. Some of these activities were frowned upon – we only know about them because of decrees which tried to put a stop to them – but some of them were not, provided they kept their own ‘zones’ of the building. It is clear that the cathedral of Chartres was a social and economic – as well as a religious – centre. The cathedral’s floor has a gentle slope to it, probably to make the place easier to sluice out after pilgrims had turned it into a mass bed-and-breakfast. Christopher Frayling 1996:65-6

iii

Table of Contents List of figures ........................................................................................................................vii List of tables .......................................................................................................................xxiv Acknowledgements ............................................................................................................. xxv Abstract .............................................................................................................................xxvii Notes.................................................................................................................................xxviii Preface ................................................................................................................................xxix

PART I....................................................................................................................... 1 INTRODUCTION: CONSTRUCTING THE PAST .................................................................. 1 Conclusions ............................................................................................................................. 6 CHAPTER ONE: DEFINING THE ‘NEOLITHIC’ ................................................................... 7 Changing perspectives on the ‘Mesolithic’ and ‘Neolithic’.................................................... 7 The ‘Neolithic’ as a technological phenomenon............................................................... 8 The ‘Neolithic’ as an economic phenomenon................................................................... 8 The ‘Neolithic’ as a social phenomenon ........................................................................... 9 The ‘Neolithic’ as defined by megaliths and earthen ‘barrows’ ....................................... 9 The ‘Neolithic’ Discourse ..................................................................................................... 10 The Phenomenological ‘Neolithic’ ....................................................................................... 11 Conclusions ........................................................................................................................... 14 CHAPTER TWO: ECONOMY AND SUBSISTENCE IN THE ‘MESOLITHIC’ AND ‘NEOLITHIC’ ........................................................................................................................... 15 Approaches to Economic Change in the ‘Neolithic’............................................................. 15 ‘Wave of Advance’ Model: Agriculture, Genetics and Linguistics................................ 15 Indigenous and Availability Models ............................................................................... 18 External Factors and Subsistence Change....................................................................... 23 Economic Continuity in Prehistoric Europe.......................................................................... 24 Economic and subsistence patterns in Britain and Ireland.................................................... 24 ‘Mesolithic’ Activities........................................................................................................... 26 ‘Neolithic’ Activities............................................................................................................. 28 Woodland Clearances...................................................................................................... 28 Woodland Regeneration? ................................................................................................ 29 Cereal Domestication ...................................................................................................... 32 Domestic Livestock......................................................................................................... 33 Continued Utilisation of Wild Resources ....................................................................... 35 Human diet: Bone stable isotope analysis............................................................................. 35 Palaeodietary Evidence in Britain and Ireland................................................................ 36 Conclusions ........................................................................................................................... 40

PART II.................................................................................................................... 42 CHAPTER THREE: IDENTIFYING UNPARALLELED BEHAVIOUR .............................. 42 The history of archaeology and standard social theory......................................................... 42 The current use of standard social theory in archaeology..................................................... 44 The issue of Uniformitarianism ...................................................................................... 45 Identifying ritual human behaviour: vague yet familiar ................................................. 47 Identifying the Unparalleled in Prehistoric Britain and Ireland ............................................ 49 A way forward…................................................................................................................... 49 Conclusions ........................................................................................................................... 50 CHAPTER FOUR: MAKING MEGALITHS FROM STONES .............................................. 52 Identifying ‘Tombs’ and ‘Barrows’ in ‘Neolithic’ western and northern Europe ................ 52 Ethnography and the megalithic structures of western and northern Europe........................ 52 iv

‘Megaliths’ in Archaeological Discourse.............................................................................. 55 Constructing a ‘Neolithic’ Ideology...................................................................................... 56 ‘Megaliths’, Subsistence and Symbols.................................................................................. 57 Pre-‘Neolithic’ Use of Stone for Construction across Europe .............................................. 64 Diversity through Continuity ................................................................................................ 77 Continuity in the Use of Large Constructions and Stone...................................................... 77 Conclusions ........................................................................................................................... 82 CHAPTER FIVE: LIVING WITH THE DEAD, MAKING ANCESTORS ............................ 83 Theoretical approaches to the Study of Mortuary Practice ................................................... 83 The Evidence: Human skeletal material and dating.............................................................. 84 Taphonomic Considerations.................................................................................................. 84 The Use of Published Archaeological Data .......................................................................... 87 Treatment and deposition of human skeletal material in the ‘Neolithic’ .............................. 87 Death and Deposition: a many varied thing.................................................................... 87 Location of human skeletal material ............................................................................. 100 Living with the dead...................................................................................................... 116 Animal and human skeletal material: conflicting views ............................................... 129 Human skeletal material in the ‘Mesolithic’ ....................................................................... 131 Treatment and deposition of human skeletal material in the ‘Mesolithic’.................... 134 Location of human skeletal material ............................................................................. 142 Living with the dead...................................................................................................... 142 Mortuary archaeology in the ‘Neolithic’: The Creation of Ancestors ................................ 145 Cooking up the Ancestors!.................................................................................................. 146 Conclusions ......................................................................................................................... 147 CHAPTER SIX: MOVING AND LIVING AROUND THE LANDSCAPE ......................... 152 Residential patterns in the ‘Neolithic’................................................................................. 152 The archaeological identification of mobility ..................................................................... 153 What do we mean by ‘settlement’ and ‘occupation’? ......................................................... 153 Born Dead?: Settlement structures and occupation in Britain and Ireland ......................... 155 ‘Filling in the Gaps’: Artefact scatters in Britain and Ireland............................................. 171 Chronology.................................................................................................................... 172 Taphonomy: Identifying Spatial Entities ...................................................................... 176 Function......................................................................................................................... 181 Order Out of Chaos?: The Future for Artefact Scatters Studies ......................................... 183 Reconstructing Lifetime Movements .................................................................................. 183 Procedures and Methods ............................................................................................... 184 Results and Conclusions ............................................................................................... 187 Occupation and Mobility in the ‘Mesolithic’ ...................................................................... 188 Circulation and Movement of Artefacts in the ‘Mesolithic’ ............................................... 188 Circulation and Movement of Artefacts in the ‘Neolithic’ ................................................. 188 The ‘Axe Trade’ and Lithic Movement and Circulation .............................................. 188 Pottery, Clay and Shell.................................................................................................. 196 Interpreting the movement and circulation of artefacts and raw materials ................... 196 Conclusions: Identifying ‘Residential’ Mobility................................................................. 197

PART III ................................................................................................................199 CHAPTER SEVEN: THE AVEBURY REGION................................................................... 199 The Avebury Region: A case study..................................................................................... 199 Environment.................................................................................................................. 209 Subsistence.................................................................................................................... 211 Palaeodietary Evidence: Bone stable isotope analysis.................................................. 212 Constructions................................................................................................................. 212 v

Location, treatment and deposition of artefacts and skeletal material ................................ 213 Deposition at contexts prior to constructions................................................................ 226 Deposition at contexts during and after construction.................................................... 228 Deposition at contexts not associated with any current visible constructions .............. 251 Review and Issues ............................................................................................................... 265 Conclusions ......................................................................................................................... 266 CONCLUSION: UNPARALLELED HUMAN SOCIAL BEHAVIOUR.............................. 267 Land use, mobility, construction and deposition................................................................. 267 Issues of mobility and deposition........................................................................................ 268 Seeing a Different Past ........................................................................................................ 271 Moving on… ....................................................................................................................... 272 Final thoughts ...................................................................................................................... 273

APPENDIX 1.........................................................................................................274 APPENDIX 2.........................................................................................................285 APPENDIX 3.........................................................................................................307 APPENDIX 4.........................................................................................................610 APPENDIX 5.........................................................................................................614 APPENDIX 6.........................................................................................................618 BIBLIOGRAPHY..................................................................................................620

vi

List of figures I.1

The local setting of the structures at Barkaer (Jutland) (after Bradley 1998a: figure 2).

4

I.2

The ‘Neolithic’ long ‘houses’ at Barkaer (Jutland) (after Bradley 1998a: figure 1).

5

1.1

The Homeland, by J. Mánes, 1856. The national self-perception created by Czech and Slovak intellectuals referred to ancestors as peaceful farmers, an image captured in this nationalist painting (after Zvelebil 1996a: figure 10.1).

10

1.2

J.M.W. Turner, Stonehenge, c.1825-28 (after Chippindale 1994: figure 80).

13

2.1

Ammerman and Cavalli-Sforza’s ‘wave of advance’ model for the spread of agriculture into Europe. The radiocarbon dates are expressed in years bp (redrawn from Ammerman and Cavalli-Sforza 1971:685). The broken lines represent probable regional variations in the rates of spread (after Thomas 1996c: figure 17.1).

16

Three archaeological patterns for the transition to crop agriculture in Europe. The vertical axis represents the length of time taken from the first introduction of cereals and legumes to the establishment of a predominantly agricultural way of life. The horizontal axis represents the importance of farming relative to hunting and gathering during the transition. I, cereals appeared suddenly in the early ‘Neolithic’, along with domestic livestock and pottery. Although there may be a slight overlap with the local ‘Mesolithic’, communities entirely dependant upon foraging seem to apply to large areas of southeast and central Europe and the Alps. II, there is no clear point at which domestic crops and animals and pottery were first used in the ‘Mesolithic’. Sites used specifically for farming are very rare, and foraging seems to have provided most of the food for several hundred years. In areas such as much of the Mediterranean perimeter and western, eastern and northern Europe, agriculture was of minor importance for up to 1000 years after the first appearance of its components. III, crop agriculture is never wholly successful and was often tried intermittently during prehistory. This pattern predominated in much of north eastern Europe and in many highland and wetland areas (after Dennell 1992: figure 5.3).

19

The agricultural frontier in north western Europe at about 6450 BP (after Bradley 1998a: figure 4).

20

The three-stage availability model of the transition to farming (after Zvelebil 1996b: figure 18.1).

21

The transition to farming along the southern rim of the Baltic in terms of the three-stage availability model. Dates calibrated BC (after Zvelebil and Lillie 2000: figure 3.9).

21

Forager-farmer contacts expected during the earlier part of the availability phase, when cooperation prevails over competition. The example is drawn from the LBK and the Ertebølle (after Zvelebil 1996b: figure 18.5).

22

Forager-farmer interaction in the east Baltic, c. 5000-4000 BC (after Zvelebil and Lillie 2000: figure 3.10).

22

Distribution of all stump diameter groupings at Hightown (Merseyside), 1994 (after Clapham et al. 1997: figure 3).

30

Part of Sweet Track. Railway site. Somerset Levels (after Hillam et al. 1990).

30

Provisional map of woodland types 5000 years ago for Britain and Ireland (after Bennett 1989: figure 1).

31

The field systems at Rathlackan (Co. Mayo) and Roughaun Hill (Co. Clare) (after Cooney 2000a: figure 2.7).

34

2.12

The layout of Céide Fields (Co. Mayo) (after Cooney 2000a: figure 2.2).

34

2.13

Faunal assemblages from southern Britain: Ratio of pigs:cattle:sheep (after Thomas 1999a: figure 2.5).

35

The ‘Mesolithic’/‘Neolithic’ transition (after Thomas 1999a: figure 2.1).

40

4.1

A Merina tomb, Madagascar (after Bloch 1977: figure 7).

54

4.2

Photograph of an above average Merina tomb, Madagascar (after Bloch 1977: plate 3b).

54

4.3

A Tandroy tomb from Madagascar (after Parker Pearson and Ramilisonina 1998a: figure 3).

55

2.2

2.3 2.4 2.5 2.6

2.7 2.8 2.9 2.10 2.11

2.14

vii

4.4

The ‘Mesolithic’ (left) and ‘Neolithic’ (right) burials at Dragsholm (Zealand) (after Bradley 1998a: figure 7).

57

The plan of a Linear Pottery Culture long house at Olszanica (Poland) (left), compared with that of Kilham long barrow (East Riding of Yorkshire) (right) (after Bradley 1998a: figure 13).

59

The distribution of long houses (vertical shading), long mounds and megalithic tombs (horizontal shading) in north western Europe (after Bradley 1998a: figure 12).

59

Synthesis of the development of funerary structures (from simple grave to gallery graves, through mounds with cists and passage graves) and succession of the regional pottery styles (after Boujot and Cassen 1993: figure 3).

60

4.8

General plan of Horslip long mound (Wiltshire) (after Ashbee et al. 1979: figure 2).

61

4.9

General plan of South Street long mound (Wiltshire). Contours are at 20 centimetre intervals. Roman numerals are ditch cuttings (after Ashbee et al. 1979: figure 23).

62

4.10

Plan of Beckhampton Road (Bishops Cannings 76) (after Barrett et al. 1991: figure 2.10).

63

4.11

The Bay of Quiberon region (Morbihan, France), showing locations of selected ‘Mesolithic’ sites (after Schulting 1996: figure 1).

65

4.12

Plan of Téviec (Morbihan) (note the cenotaph?) (after Schulting 1996: figure 2).

65

4.13

Grave K at Téviec (Morbihan) (after Schulting 1996: figure 4).

66

4.14

Plan of Hoëdic (Morbihan, France) (after Schulting 1996: figure 3).

66

4.15

Grave K at Hoëdic (Morbihan) with red deer antler structure (after Schulting 1996: figure 11).

67

4.16

Outline plans of Tumulus-St-Michel (Morbihan, France) (after Bradley 1998a: figure 17).

69

4.17

The tertre tumulaire of Le Manio I (Morbihan, France) (after Scarre 1998: figure 3).

69

4.18

Long mounds of the Carmac/Locmariaquer area (Morbihan, France). Note the presence in each of these monuments of an initial circular or oval core (after Scarre 2002a: figure 11).

70

Compilation of areas excavated in the late 1950s and in the 1980s at Poças de São Bento (Baixo-Alentejo, Portugal) 1, the extent of the shell midden: 2, colouring left by post-holes; 3, pits; 4, large stones. Roman numerals indicate human skeletons (after Larsson 1996: figure 8).

70

4.20

Colombres (Asturias, Spain): Burial and associated features (after Clark 1983: figure 3.14).

71

4.21

Roc del Migdia (Cataluña, Spain): ‘Megalithic’ burial (after Bahn 1989a: figure 1).

72

4.22

Plan of Dunford Bridge Site B (South Yorkshire). The dashed line marks the limits of the main flint distribution. The blackened area marks the hearth (after Radley et al. 1974: figure 4).

72

Plan of Dunford Bridge Site A (South Yorkshire). The dashed line marks the limits of the main flint distribution. The stippled area was rich in burnt flints and charcoal fragments (after Radley et al. 1974: figure 5).

73

Barsalloch: Plan of the excavated area with key to grid squares (after Cormack 1970: figure 2).

73

Culverwell (Dorset): Plan of the floor and hearth 4, Area A and B (part) (after Palmer 1999: figure 6(1)).

75

Culverwell (Dorset): Plan of the floor and features in Areas A and B (after Palmer 1999: figure 6(2)).

75

Culverwell (Dorset): Plan of the floor in part of Trench 3 indicating a straight line of limestone slabs which, possibly, could have provided support for the footings of a superstructure, such as a shelter or ‘hut’ (after Palmer 1999: figure 10).

76

4.28

Culverwell (Dorset): Parts of Area A and B (after Palmer 1999: figure 6(5)).

76

4.29

Culverwell (Dorset): View of part of the floor in Area A with two straight lines of limestone slabs next to a square gap in the floor which could possibly mark the position of a superstructure (after Palmer 1999: plate 7).

77

Lussa Wood 1 (Argyll and Bute): a - rings; b - cobbled patch (after J. Mercer 1981: figure 4).

78

4.5

4.6 4.7

4.19

4.23

4.24 4.25 4.26 4.27

4.30

viii

4.31

March Hill Carr (West Yorkshire), Trench A, Site 1 excavations in 1994 showing hearth 1 and 2 and the refit patterns (after Spikins et al. 2002: figure 2).

78

Plan and profile of feature F41 at Staosnaig (Argyll and Bute) (after Mithen and Finlay 2000: figure 5.2.35).

79

Feature F41 at Staosnaig (Argyll and Bute), following removal of fill showing stone lining (after Mithen and Finlay 2000: figure 5.2.37).

79

Plan of the initial phase of the construction at Street House (Cleveland) (after Vyner 1984: figure 2).

80

4.35

Plan of the final form of the cairn at Street House (Cleveland) (after Vyner 1984: figure 7).

80

4.36

‘Tomb’ No. 4 at Carrowmore (Co. Sligo) (after Burenhult 2001: page 19).

80

4.37

The round dolmen at Skredsvik (Bohuslän) (after Tilley 1996a: figure 4.23).

81

Distribution of radiocarbon dated human skeletal material from British caves (after Chamberlain 1996: figure 1).

85

Map generated from Online Database showing the distribution of all contexts with radiocarbon obtained from samples identified as human skeletal material (excluding loose teeth).

86

An archaeological view of the processes that human and animal skeletal material can undergo (after Kinnes 1992b: figure 2.5.3).

88

Suggested structure, use and two phases of decay of a three-point support raised platform (after Scott 1992: figure 8.7).

90

Human corpse found in a rural woodland area after being exposed for 22 days (after Haglund et al. 1989: figure 4).

91

Human corpse found in a rural woodland area after being exposed for 2½ months (after Haglund et al. 1989: figure 5).

91

Anatomical distribution of human remains from Hambledon Hill (Dorset), shown as a percentage of the total bones present (after Evans 2000: figure 7.02).

93

Graph showing the anatomical distribution of both articulated and disarticulated cut human bone, displayed as a percentage of the total number of cut bone (after Evans 2000: figure 7.10).

93

Anatomical distribution of cut marked cattle bone from Hambledon Hill (Dorset), expressed as a percentage of the total number of bones recorded (after Evans 2000: figure 7.21).

93

Fine cut marks appearing under magnification, parallel to the main cut marks noted during recording (human skull - 682, 725 + 1037, HH77) (after Evans 2000: figure 7.14).

94

Patterns of parallel cut marks on a human mandible, seen under magnification using the stereo microscope (1916, HH76) (after Evans 2000: figure 7.15).

94

Scanning electron microscope of a cut mark illustrating the weathered ‘rounded’ appearance of the cuts, not clearly visible through the naked eye (human scapula - 2919, ST78) (after Evans 2000: figure 7.16).

95

5.13

Well defined cut marks on a cattle metacarpal (1470, HH75) (after Evans 2000: figure 7.23).

95

5.14

Well defined cuts on a cattle skull (1613, HH75) (after Evans 2000: figure 7.24).

96

5.15

Cuts across the distal articulation on a cattle metatarsal (2346, ST79) (after Evans 2000: figure 7.25).

96

Parallel cuts on a cattle mandible, made clear under magnification (708, HH75) (after Evans 2000: figure 7.26).

97

Scanning electron microscope photograph showing the well defined ‘V’ shaped cut marks on a cattle metatarsal (2346, ST79) (after Evans 2000: figure 7.27).

97

Scanning electron microscope photograph showing well defined ‘V’ shaped cuts on a cattle tibia (266, HH74) (after Evans 2000: figure 7.28).

98

State of human skeletal material at long mounds in Wessex (after Thorpe 1984: figure 4.6).

101

4.32 4.33 4.34

5.1 5.2

5.3 5.4 5.5 5.6 5.7 5.8

5.9 5.10 5.11 5.12

5.16 5.17 5.18 5.19

ix

5.20

State of human skeletal material identified as adults at long mounds in Wessex (after Thorpe 1984: figure 4.11).

101

State of human skeletal material identified as children at long mounds in Wessex (after Thorpe 1984: figure 4.12).

102

Plan and section of the central pit at Willington Plantation Quarry (Bedfordshire) (after Dawson 1996: figure 3).

103

5.23

Alington Avenue (Dorset): Phase 1, earlier prehistoric (after Davies et al. 1985: figure 2).

104

5.24

Plan and section of Chilbolton (Leckford Estate) (Hampshire), including details of central features (after Russel 1990: figure 2).

105

Child skeleton at base of ditch at Flagstones House (Dorset) (after Smith et al. 1997: figure 28).

105

A human skull fragment found on the clays at Carrigdirty Rock (Co. Limerick) (after O'Sullivan 1997a: page 15).

106

5.27

Distribution of human skeletal material at Stonehenge (after Cleal et al. 1995: figure 250).

106

5.28

Longitudinal section of Kilgreany (Co. Waterford) (after Movius 1935: figure 3).

107

5.29

Excavation of a skeleton in cist 2 at An Corran (Highland) (after Downes and Badcock 1998: figure 28).

107

5.30

Bog body found at Castle Balkeney (Gallagh) (Co. Galway) (after Ó Floinn 1995a: figure 66).

108

5.31

Schematic drawing of the excavated deposits at Carding Mill Bay (Argyll and Bute), showing sections (A-B, C-D, E-F) and a summary of the phasing of contexts (after Connock et al. 1991-2: figure 2).

108

Sketch plan and section of Quanterness (Orkney Islands), with the later roundhouse to the east (after Renfrew et al. 1976: figure 1).

109

5.33

Plan and sections of Grange (Co. Roscommon) (after Ó Ríordáin 1997: figure 2).

110

5.34

Photograph of the human skeletal material at Giants’ Hills 2 (Lincolnshire), from the west. 30 centimetre scale (after Evans and Simpson 1991: plate viia).

111

Photograph of the primary articulated teenage male skeleton with Food Vessel at Tallington 16 (Lincolnshire) (after W.G. Simpson 1976: plate 24 upper).

111

5.36

Ground plan and section of Collessie (Fife) (after Anderson 1888: figure 3 and 4).

112

5.37

General plan of Fussell’s Lodge long mound (Wiltshire) as excavated (after Ashbee 1966a: figure 2).

113

Fussell’s Lodge (Wiltshire): Distribution of human anatomical parts (after Thomas 1999a: figure 6.5).

114

Contexts of burials in early ‘Neolithic’ Wessex for children and adults (after Thorpe 1996: figure 7.6).

115

Contexts of deposition of body parts in early ‘Neolithic’ Wessex (after Thorpe 1996: figure 7.7).

115

A schematic diagram illustrating the relationship between the deposition of human skeletal material and the landscape (after Jones 1998: figure 4).

116

Phases of Beckhampton Road (Bishops Cannings 76) (Wiltshire) (after Pollard 1993: figure 2.36).

118

Reproduction of plate XIV from Nilsson’s The Primitive Inhabitants of Scandinavia showing ground plans of Swedish passage graves and a reconstruction diagram of an Eskimo house (after Tilley 1998: figure 4).

119

Comparison of ‘tombs’ (stalled cairns - A, B, C) and ‘houses’ (D) in the Orkney ‘Neolithic’. A. Kierfea Hill. B. Knowe of Craie. C. Knowe of Yarso. D. Knap of Howar (after Hodder 1984: figure 7).

120

5.45

Plan of Skara Brae (Orkney Islands) (after Shee Twohig 1981: figure 286).

121

5.46

Abstract design at the Maeshowe (Orkney Islands) chambered ‘tomb’ (left) overlain by runic inscription (right) (after Bradley et al. 2001: figure 6).

122

5.21 5.22

5.25 5.26

5.32

5.35

5.38 5.39 5.40 5.41 5.42 5.43

5.44

x

5.47

Incised motifs on the Orkney Islands at the Cuween Hill chambered ‘tomb’ (1-3), the Holm of Papa Westray South chambered ‘tomb’ (4) and the Quoyness chambered ‘tomb’ (5) (after Bradley et al. 2001: figure 8).

123

Incised motifs on the Orkney Islands at the Quoyness chambered ‘tomb’ (6 and 7) and the Wideford Hill chambered ‘tomb’ (8) (after Bradley et al. 2001: figure 9).

124

Incised motifs on the Orkney Islands at the Wideford Hill chambered ‘tomb’ (after Bradley et al. 2001: figure 10).

125

Incised motifs from the chambered ‘tomb’ at Maeshowe, the ‘house’ Barnhouse and a cist slab from Brodgar from the Orkney Islands (after Bradley et al. 2001: figure 13).

125

Skara Brae (Orkney Islands). Decorated stones from ‘Hut’ 7 and 8 (after Shee Twohig 1981: figure 287).

126

Skara Brae (Orkney Islands). Decorated stones from the Passages (after Shee Twohig 1981: figure 288).

127

5.53

Skara Brae (Orkney Islands). Stones 42-5 (after Shee Twohig 1981: figure 289).

128

5.54

Skara Brae (Orkney Islands). Decorated stones from the passage and miscellaneous locations (after Shee Twohig 1981: figure 290).

129

Windmill Hill (Wiltshire) Context 630, Trench F. Early ‘Neolithic’. The distal half of the humerus of an ox, with the shaft of the femur of a child inserted into the marrow cavity (after Grigson 1999: figure 161).

130

Predominant character of human skeletal material in the chambers of three chambered structures: 1) Burn Ground (Gloucestershire); 2) Notgrove (Gloucestershire); 3) West Kennet (Wiltshire) (after Thomas 1988b: figure 6).

132

Relative representation of human bones, animal bones and pottery vessels in different phases of the secondary deposit at West Kennet (after Thomas and Whittle 1986: figure 6).

133

Cut marks on a human lower jaw bone from Gough’s New Cave (Somerset) (after Barton 1999: figure 2.5).

134

Cnoc Coig (Argyll and Bute); (a) schematic section (note vertical exaggeration); (b) plan showing structures within midden and hearth areas (after Smith 1992: figure 8.8).

136

Horizontal plot showing the distribution of all human skeletal material in the levels at Cnoc Coig (Argyll and Bute) (after Nolan 1986: figure 63).

137

5.61

Grotte Margaux: Plan of the early ‘Mesolithic’ cave (after Cauwe 2001: figure 6).

139

5.62

Grotte Margaux: Distribution of human skeletal material in the cave (after Cauwe 2001: figure 7).

140

Abri des Autours: Plan of the ‘Mesolithic’ cave (spelling from original) (after Cauwe 2001: figure 11).

141

Skateholm II. Grave XXI with a buried dog. A deer antler lay along the back of the dog and three flint knives were found on its stomach (above). A decorated hammer of red deer antler was placed on the chest of the dog (below) (after Larsson 1989a: figure 7).

143

Section and deposits of Ogof-yr-Ychen (Cave of the Oxen) (Pembrokeshire) (after Davies 1989: figure 7.4).

144

Feature location plan of the northern area at Ferriter’s Cove (Co. Kerry) (after Woodman et al. 1999: figure 2.1).

145

5.67

Plan of 5MT10010 (Colorado, U.S.A.) (after Billman et al. 2000: figure 2).

148

5.68

Plan of pithouse Feature 3, 5MT10010 (Colorado, U.S.A.) (after Billman et al. 2000: figure 3).

149

Plan of pithouse Feature 13, 5MT10010 (Colorado, U.S.A.) (after Billman et al. 2000: figure 4).

150

Cut marks on an adolescent femur from 5MT10010 (Colorado, U.S.A.) (after Billman et al. 2000: figure 7).

151

Bar chart showing the discovery rate of ‘Neolithic’ buildings in England, Wales and the Isle of Man (after Darvill 1996: figure 6.2).

156

5.48 5.49 5.50 5.51 5.52

5.55

5.56

5.57 5.58 5.59 5.60

5.63 5.64

5.65 5.66

5.69 5.70 6.1

xi

6.2

Ground plan of the building at “Benie Hoose” (Shetland Islands) (after Barclay 1996: figure 5.2).

156

Ground plan of the ‘Neolithic’ building at Standing Stones of Yoxie (Shetland Islands) (the black marks are upright stones and the darker-toned area is a later alteration) (after Barclay 1996: figure 5.1).

157

Ground plan of the ‘Neolithic’ building at Gruting School (Shetland Islands) (black spots are post-holes, post-settings or post-supports) (after Barclay 1996: figure 5.1).

157

Ground plan of the ‘Neolithic’ building at Scord of Brouster 3 (Shetland Islands) (the crosshatched area is a hearth and the black spots are post-holes, post-settings or post-supports) (after Barclay 1996: figure 5.1).

158

Ground plan of the ‘Neolithic’ building at Stanydale ‘Temple’ (Shetland Islands) (the black marks are upright stones, except where marked ‘p’ which are post-holes) (after Barclay 1996: figure 5.1).

158

Ground plan of the ‘Neolithic’ building at Ness of Gruting (Shetland Islands) (the crosshatched area is a hearth) (after Barclay 1996: figure 5.1).

159

6.8

Plan of the structure at Claish Farm (Perth and Kinross) (after Barclay et al. 2002: figure 3).

159

6.9

Plan of the structure at Lough Gur B (Co. Limerick) (after Cooney 2000a: figure 3.1).

160

6.10

Plan of the structure at Lough Gur K2 (Co. Limerick) (after Cooney 2000a: figure 3.1).

160

6.11

Plan and section of Lough Gur C (Co. Limerick) (after Cooney 2000a: figure 3.3).

161

6.12

Plan of the Lough Gur C complex (Co. Limerick) (after Cooney 2000a: figure 3.4).

162

6.13

Plan of the structure at Slieve Breagh 1 (Co. Meath) (after Cooney 2000a: figure 3.2).

163

6.14

Plan and activity phases of Ballyharry (Co. Antrim) (after Cooney 2000a: figure 3.5).

163

6.15

Plan of the structure at Ballygalley 1 (Co. Antrim) (after Cooney 2000a: figure 3.1).

163

6.16

Plan of the ‘Neolithic’ structure at White Horse Stone (Kent) (after Oxford Archaeological Unit 2000: page 451).

164

Pre-excavation ground plans for ‘Mesolithic’/‘Neolithic’ structures at Bowman’s Farm (Hampshire) (after Green 1996: figure 7.3).

166

Mount Sandel Upper (Co. Derry): ‘Hut’ A showing hearths a, b, c and d; pits 1 and 2; postholes in solid black (after O’Kelly 1989: figure 9).

166

Plan of Broomhead Moor Site 5 (South Yorkshire), showing stake-holes (black) and hearth areas (stippled) (after Radley et al. 1974: figure 2).

167

Location, plan and profiles of stake-holes in trench 1, Bolsay Farm (Argyll and Bute) (after Mithen et al. 2001a: figure 4.9.23).

167

Plan and profile of stake-holes F40 and F31, trench II, Bolsay Farm (Argyll and Bute) (after Mithen et al. 2001b: figure 4.10.19).

168

Kindrochid (Argyll and Bute): Plans and profiles of features in areas 1 and 2 (after Marshall and Mithen 2001: figure 4.7.8).

168

Plan of the ‘Mesolithic’ structure at Broom Hill (Hampshire) (after O’Malley 1978: page 118).

169

6.24

Morton on Tay (Fife): Plan of occupation II (after Megaw and Simpson 1979: figure 2.28).

169

6.25

Deepcar (South Yorkshire): The distribution of local and foreign stones; quartzites (black), river-worn grits (stippled), local flags (plain) and three possible hearths are indicated (hatched). The section, the flint horizon (lightly stippled) and disturbed areas (shaded) are recorded (after Radley and Mellars 1964: figure 2).

170

‘Neolithic’ chambered structures rebuilt in the ‘Iron Age’: a) Quanterness (Orkney Islands); b) Howe Phase 3/4 (Orkney Islands); c) Howe Phase 5 (Orkney Islands); d) Clettraval (Western Isles); e) Unival (Western Isles) (after Hingley 1996: figure 2).

170

6.3

6.4 6.5

6.6

6.7

6.17 6.18 6.19 6.20 6.21 6.22 6.23

6.26

xii

6.27

Above: Bar chart of artefact scatters by period. PA = ‘Palaeolithic’; ME = ‘Mesolithic’; NE = ‘Neolithic’; BA = ‘Bronze Age’; UN = Undated. Below: Bar chart of multiperiod artefact scatters in which the numbers refer to the number of these periods represented in the multiperiod artefact scatter (after Schofield 2000: figure 5.2).

173

Bar chart showing some of the results of the pilot study, arranged by county. Non-disc. Unkn. = non-discrete unknown; part disc. = partially discrete: minor = minor alterations only to original scale of artefact scatter: excav. = excavated (after Schofield 2000: figure 5.3).

174

6.29

Some results from the pilot study undertaken in 1994-5 (after English Heritage 2000: page 6).

175

6.30

Radiocarbon dates for Peterborough Wares, Grooved Wares and Beakers (uncalibrated dates bp) (after Thomas 1999a: figure 5.10).

179

Spong Hill (Norfolk). A) Comparison of material from unstratified and stratified contexts, B) Hypothetical reconstruction of depositional history of artefacts (after Healy 1987: figure 2.1).

179

Maddle Farm and Vale of the White Horse Fieldwalking Survey: Results of fieldwalking the same field in consecutive years (after Tingle 1987: figure 7.2).

180

6.33

Core reduction sequence and its archaeological correlates (after Schofield 1991b: figure 4).

182

6.34

Strontium isotope ratios for the deciduous and permanent tooth enamel and dentine of the three juveniles (A, B and D) and permanent tooth enamel and dentine for the adult female (C). Error bar represents 2į errors calculated from 10 replicates of the NBS 987 standard (after Montgomery et al. 2000: figure 1).

185

Lead isotope ratios for the deciduous and permanent tooth enamel (large symbols) and dentine (small symbols) of the three juveniles (A, B and D) and permanent tooth enamel and dentine for the adult female (C). Errors on the Pb isotope ratios were determined from replicate analysis of the NBS 981 standard as 0.014% for 207Pb/206Pb and 0.024% for the 208 Pb/206Pb (n = 27, 2į). These are smaller than the symbols at this scale (after Montgomery et al. 2000: figure 2).

186

Location map showing Monkton-up-Wimbourne (Dorset) relative to the approximate extent of the Cretaceous chalk geology and the Mendips orefield to the north west. Approximate ranges for lead and strontium isotopes for each geology are indicated (after Montgomery et al. 2000: figure 4).

186

Strontium isotope ratios for the juveniles’ tooth enamel relative to its time of formation in year before death. Error bars associated with the permanent teeth represent the maximum period of enamel mineralisation. The adult female, C, is represented by dashed lines indicating the values of her enamel and dentine (after Montgomery et al. 2000: figure 3).

187

The inferred seasons of use of the Oronsay middens (Argyll and Bute), based on faunal evidence (after Richards and Mellars 1998: figure 2).

189

Distribution of main raw material sources in relation to Waun Fignen Felen (Powys) (after Barton et al. 1995: figure 6).

189

Principal sources of non-flint stone identified through petrological analysis. The hatched area represents the main outcrops of flint and the triangles mark the position of ‘Neolithic’ flint mines (after Edmonds 1995: illustration 5).

190

Distribution map of Group VI axes in Ireland identified by thin section (open circles indicate county or area only provenance) (after Cooney and Mandal 1995: figure 4).

191

6.42

Distribution of porcellanite axes in Ireland (after Cooney 2000a: figure 6.14).

192

6.43

Distribution of stone axes from Cornish (petrological group I) and Cumbrian (petrological group VI) sources (after Edmonds 1993b: figure 7.2).

193

Distribution of stone axes from sources in north (petrological group VII) and south (petrological group VIII) Wales (after Edmonds 1993b: figure 7.3).

193

Location of porcellanite axes in Britain and tuff axes in Ireland (after Cooney 2000a: figure 6.16).

194

Simplified geological map of the area around Cherhill (Wiltshire), showing clay outcrops and the approximate locations of sample points. A: ST 994 791; B: SU 012 771; C: SU 029 748; D: SU 024 744; E: SU 018 684 (after Darvill 1983: figure 27).

197

General location map of the Avebury region (Wiltshire) (after David et al. 1999: figure 7).

200

6.28

6.31 6.32

6.35

6.36

6.37

6.38 6.39 6.40

6.41

6.44 6.45 6.46

7.1

xiii

7.2

‘Neolithic’ features of the Avebury region (Wiltshire), referred to in figures 7.3, 7.4 and 7.5 (after R.W. Smith 1984: figure 1). AV - Avebury: Henge; BR - Beckhampton Road: Earthen Long Mound; G.55 - Avebury G.55: Round Mound; G.6a - West Overton G.6a: Round Mound; HA - Hackpen: Occupation Debris; HK - Hemp Knoll: Round Mound; HP - Horslip: Earthen Long Mound; SA - Sanctuary: Timber Circle; SB - Silbury: Mound; SS - South Street: Earthen Long Mound; WA - Waden Hill: Occupation Debris; WK - West Kennet: Chambered Mound; WKA - West Kennet Avenue: Occupation Debris beneath stone avenue; WH - Windmill Hill: Causewayed Enclosure.

201

Schematic reconstruction of the early ‘Neolithic’ landscape of the Avebury region (Wiltshire) (after R.W. Smith 1984: figure 8).

203

Schematic reconstruction of the middle ‘Neolithic’ landscape of the Avebury region (Wiltshire) (after R.W. Smith 1984: figure 9).

204

Schematic reconstruction of the late ‘Neolithic’ landscape of the Avebury region (Wiltshire) (after R.W. Smith 1984: figure 10).

205

7.6

Avebury region (Wiltshire): Earlier ‘Neolithic’ (after Thomas 1999a: figure 9.1).

206

7.7

Avebury region (Wiltshire): Middle/later ‘Neolithic’ (after Thomas 1999a: figure 9.4).

207

7.8

Avebury region (Wiltshire): Later ‘Neolithic’/Beaker (after Thomas 1999a: figure 9.7).

208

7.9

Plan of Avebury (after Malone 1989: page 85).

214

Plan of the excavated features and details of the inner timber and stone circles at The Sanctuary (rings B to G) (after Pollard 1992: figure 1).

215

Plan of the West Kennet Avenue, showing the excavation trenches, position of the stones, Beaker pits and the location of the ‘late Neolithic’ occupation debris (after Malone 1989: figure 67).

216

Schematic plan of Silbury Hill (Wiltshire), with locations of principal excavations (after Whittle 1997c: figure 5).

217

Plan of the West Kennet palisade enclosures (Wiltshire). Much of the layout of enclosure 2 is based on air photographic evidence (after Whittle 1997c: figure 28).

218

Longstones Field (Wiltshire): Above, location in the Avebury region; below, plan of excavations (the hatched lines are Medieval ridge and furrow) (after Gilling et al. 2000c: page 429).

219

Longstones Field (Wiltshire): Plan of the 2000 excavations. Trenches shown in stipple outline relate to the 1999 season (after Gillings et al. 2002a: figure 1).

220

Plan and summary cross-section of the Easton Down long mound (Wiltshire). The solid line on the plan is an estimate of the extent of the built mound (after Whittle et al. 1993: figure 2).

221

Hemp Knoll (Wiltshire). Plan of the occupation debris, round mound and ditches (after Robertson-Mackay 1980: figure 2).

222

7.18

General plan of West Overton G6B (Wiltshire) (after Smith and Simpson 1966: figure 2).

222

7.19

Composite plan of Manton Down long mound (Wiltshire). Chamber shown restored to its original form (after C.T. Barker 1985: figure 2).

223

Drawing of Old Chapel (Wiltshire) by William Stukeley, 1723 (after C.T. Barker 1985: figure 5).

223

7.21

Plan of Temple Bottom (Wiltshire) (after C.T. Barker 1985: figure 3).

224

7.22

General plan of West Kennet (after Piggott 1962: figure 2).

224

7.23

Plan and ditch summary of Millbarrow (after Whittle 1994a: figure 2).

225

7.24

Drawing of Shelving Stone (Wiltshire) by John Britton (no date) (after C.T. Barker 1985: figure 4).

225

Plan of human skeleton in the pre-bank pit in trench BB at Windmill Hill (after Whittle et al. 1999a: figure 76).

227

Photograph of human skeleton in the pre-bank pit in trench BB at Windmill Hill (after Whittle et al. 1999a: figure 75).

227

7.3 7.4 7.5

7.10 7.11

7.12 7.13 7.14

7.15 7.16 7.17

7.20

7.25 7.26

xiv

7.27

The distribution of struck flint in test pits around the Easton Down long mound (Wiltshire) (after Whittle et al. 1993: figure 6).

229

7.28

Plan of Roughridge Hill (Bishops Cannings 62a) (after Pollard 1993: figure 2.33).

230

7.29

Plan of Windmill Hill (after Whittle et al. 1999a figure 14).

232

7.30

Animal and human bones in discrete clusters along the base of the ditch at Windmill Hill (after Malone 1989: figure 34).

233

Excavation of child skeleton in OD IIIB at Windmill Hill. Inside of the ditch is to the left (after Whittle et al. 1999a: figure 29).

233

Finds distribution in MD X and MD XI at Windmill Hill (after Whittle et al. 1999a: figure 34).

234

7.33

Sections of OD I and OD III at Windmill Hill (after Whittle et al. 1999a: figure 26).

235

7.34

Finds distribution in OD III at Windmill Hill (after Whittle et al. 1999a: figure 25).

236

7.35

Child skeleton resting on bottom of OD III at Windmill Hill (after Smith 1965a: plate VIIIa).

237

7.36

Plan of Rybury (after Cunliffe 1993: figure 2.5).

237

7.37

Plan of Knap Hill (after Connah 1965: figure 1).

238

7.38

Sections of Knap Hill (after Connah 1965: figure 4).

238

7.39

Detailed plan of the chambers at West Kennet (after Piggott 1962: figure 8).

241

7.40

Photograph of the north-east chamber at West Kennet (after Piggott 1962: plate xva).

242

7.41

Photograph of the north-east chamber at West Kennet (after Piggott 1962: plate xvb).

242

7.42

Photograph of the south-east chamber at West Kennet (after Piggott 1962: plate xvi).

243

7.43

Photograph of the south-west chamber at West Kennet (after Piggott 1962: plate xviib).

244

7.44

Photograph of the human skeletal material in the blocking of the north forecourt area at West Kennet (after Piggott 1962: plate xixa).

244

Distribution of human skeletal material from rings B to G at The Sanctuary (after Pollard 1992: figure 6).

245

7.31 7.32

7.45 7.46

The adolescent skeleton beside stone hole C12 at The Sanctuary (after Pitts 2001: figure 3). 245

7.47

Section through the ditch of Avebury (after Malone 1989: page 83).

246

7.48

Avebury from the southern entrance (after Pollard 1993: figure 2.39).

247

7.49

Avebury from the north-west sector and finds from stone holes (after Pollard 1993: figure 2.43).

248

Pit with human skeletal material by stone hole 22b at West Kennet Avenue (after Smith 1965a: plate XXXVIa).

249

Pit disturbed by human skeletal material of stone 25b at West Kennet Avenue (after Smith 1965a: plate XXXVIb).

249

Longstones Field (Wiltshire): The enclosure ditch and primary deposits I Trench 23 (after Gillings et al. 2002a: figure 3).

251

7.53

Plan of Cherhill (after Evans and Smith 1983: figure 3).

253

7.54

Distribution of worked flint found in the field walking survey in North Field (Wiltshire) (after Whittle et al. 2000: figure 2).

254

Distribution of artefacts found in the field walking survey in North Field (Wiltshire) (after Whittle et al. 2000: figure 3).

255

Distribution of worked flint found in the test pit survey in North Field (Wiltshire) (after Whittle et al. 2000: figure 4).

256

Distribution of artefacts found in the test pit survey in North Field (Wiltshire) (after Whittle et al. 2000: figure 5).

257

7.50 7.51 7.52

7.55 7.56 7.57

xv

7.58

The distribution of struck flint around the Roughridge long mound (Wiltshire), as recovered by surface survey (after Whittle et al. 1993: figure 7).

258

West Overton (Wiltshire), cutting DN and D, south east section. WO86-1 and WO-V are mollusc/ostracod sample columns (after Evans et al. 1993: figure 22).

259

West Overton (Wiltshire), cutting DN, south east section. Lynchet in Avebury Soil (after Evans et al. 1993: figure 23).

260

Avebury (Wiltshire), profiles of the Avebury Soil showing dating and environment. Stippled = layer 6; unbroken lines = layer 7; broken lines = layer 8. Pits 8 and 9 are from the 1984 excavations, incorporated into cuttings E and J respectively (after Evans et al. 1993: figure 9).

261

7.62

West Overton (Wiltshire), cutting P, north west section (after Evans et al. 1993: figure 25).

262

7.63

Avebury (Wiltshire), cutting J, north section (upper) and plan at layer 5 (lower). In the plan: a, charcoal; b, earlier features or features in isolation; c, later features. A, D, F, G and I are features of layer 5 (after Evans et al. 1993: figure 8).

263

Distribution of all humanly struck flint recovered from the Avebury region (Wiltshire) in 1983 (after Holgate 1988a: figure 6.13).

264

C.1

Plan and section of Tooth Cave (Swansea) (after Harvey et al. 1967: plate).

269

C.2

Plan and section of Ossom's Crag Cave (Yellersley Tor Cave) (Staffordshire), showing sectors I-XIII of excavation (after Bramwell et al. 1987: figure 3).

270

A reconstruction of the features excavated at Dolní VČstonice I (Moravia, Czech Republic). Key: 1. kiln, 2. charcoal, 3. ash, 4. limestone blocks, 5. bone, 6. edge of artificial depression, 7. post-holes and 8. stake-holes (after Gamble 1999: figure 7.14).

298

El Juyo (Cantabria, Spain): Limits of dugouts, with rubble walls and platform (after Freeman et al. 1988: figure 1.16).

299

El Juyo (Cantabria, Spain): Level 4, an early stage of construction (after Freeman et al. 1988: figure 1.17).

300

El Juyo (Cantabria, Spain): Level 4, a later stage of construction (after Freeman et al. 1988: figure 1.18).

300

Articulated skeleton from Grotta Paglicci (Abruzzi, Italy). The height of the body is estimated at 1.60 m., A: thirty deer canines pierced for suspension found around the skull. B: Cypraea shell. C: block of hematite found below the right tibia. D: bone chisel. 1-11: flint tools including blades, end scraper, burins and a knife (after Gamble 1986: figure 5.18).

301

Hajduþka Vodenica Ib (Iron Gates Gorge, Yugoslavia). Later building structure in the central space: rectangular hearth with a side channel, and a threshold. Burials are around the hearth and at the rear of the structure, i.e., at the entrance to the chamber ‘tomb’. The chamber ‘tomb’ was not yet excavated when this photograph was taken (after Radovanoviü 1996: figure 3.40).

302

Lepenski Vir I (Iron Gates Gorge): In the foreground the foundation of house No. 24 (Lepenski Vir Ie), in the middle distance houses No. 19 (Lepenski Vir Id), 31 (Lepenski Vir Ic) and 23 (Lepenski Vir Ie) and in the background, near the Danube bank, houses Nos 10 (Lepenski Vir Ic), 11 (Lepenski Vir Id), 9 (Lepenski Vir Id) and 7 (Lepenski Vir Ic) (after Srejoviü 1972: plate 7).

303

Hearth in ‘house’ No. 24, Lepenski Vir Id (Iron Gates Gorge). The long sides of the hearths are bordered by a frieze of equilateral triangles made of small reddish stone slabs embedded vertically in the floor (after Srejoviü 1972: plate 21).

303

Padina (Iron Gates Gorge, Yugoslavia): Two men buried in sitting position (after Radovanoviü 1996: figure 4.1).

304

Riparo di Villabruna (Veneto, Italy): Burial of an adult male with some rocks covering the skeleton (after Mussi 2001: figure 7.33).

304

Tingby (Kalmer, Sweden) ‘house’. 1, post-hole, 2, peg-hole, 3, unspecified feature, 4, hearth, 5, stone, 6, limits of occupation layer (after Westergren 1995: figure 5).

305

Graves in the central part of the Vlasac (Iron Gates Gorge, Yugoslavia) terrace (after Radovanoviü 1996: figure 4.11).

306

7.59 7.60 7.61

7.64

A2.1

A2.2 A2.3 A2.4 A2.5

A2.6

A2.7

A2.8

A2.9 A2.10 A2.11 A2.12

xvi

A2.13

Eastern sector of the Vlasac (Iron Gates Gorge, Yugoslavia) terrace. A young man buried in sitting position with the skull of an older man (after Radovanoviü 1996: figure 4.15).

306

Plans and sections of Achnasavil (Argyll and Bute) (after Carter and Tipping 1991-2: figure 6).

445

Plans and sections of Achnasavil (Argyll and Bute) (after Carter and Tipping 1991-2: figure 3).

446

A3.3

Early ‘Neolithic’ features at Ardnadam (Argyll and Bute) (after Rennie 1984: figure 5).

447

A3.4

‘Neolithic’ features at Ardnadam (Argyll and Bute) (after Rennie 1984: figure 6).

448

A3.5

Late ‘Neolithic’ features at Ardnadam (Argyll and Bute) (after Rennie 1984: figure 8).

449

A3.6

Plan of structure at Ardnave (Argyll and Bute), hearths are represented by cross-hatching (after Ritchie and Welfare 1983: figure 4).

450

Post-excavation plan of Balbridie (Aberdeenshire) (after Fairweather and Ralston 1993: figure 2).

451

A3.8

Plan of Ballachrink (Isle of Man) (after Burrow 1997: figure A1.1).

451

A3.9

Plan of Ballavarry (Isle of Man) (after Burrow 1997: figure A1.6).

452

Bally Lough (Co. Waterford): Density of lithic scatters per in Fornaught Strand Basin (Zvelebil et al. 1992: figure 3).

453

A3.11

Ballygalley (Co. Antrim): Plan of major excavated area (after Simpson et al. 1990: figure 2).

453

A3.12

Ballygalley (Co. Antrim): Density distribution of flakes and debitage (after Simpson et al. 1990: figure 5).

454

Ballygalley (Co. Antrim): Distribution of flint artefacts and pottery (after Simpson et al. 1990: figure 6).

455

A3.14

Plan of the structure at Ballyglass 1 (Co. Mayo) (after Cooney 2000a: figure 3.1).

456

A3.15

Plan of the structure at Ballynagilly (Co. Tyrone) (after Cooney 2000a: figure 3.1).

456

A3.16

Plan of Barford structure C (Warwickshire) (after Darvill 1996: figure 6.5).

457

A3.17

Plan of Barford structure B (Warwickshire) (after Darvill 1996: figure 6.10).

457

A3.18

Plan of Barkhale (West Sussex) with excavation trenches (after P.C. Leach 1979: figure 2).

458

A3.19

Barleycroft Farm (Cambridgeshire) (after Evans et al. 1999: figure 3).

459

A3.20

Barleycroft Farm (Cambridgeshire): Buried soil fieldwalking distribution (after Evans et al. 1999: figure 2).

460

A3.21

Plan of Barnhouse excavations (Orkney Islands) (after Richards 1992b: page 446).

461

A3.22

General plan of building 2 at Barnhouse (Orkney Islands), showing access routes (after Barclay 1996: figure 5.4).

462

A3.23

Ground plan of building 8 at Barnhouse (Orkney Islands) (after Barclay 1996: figure 5.2).

462

A3.24

Features in Bay Farm II (Co. Antrim) layer 6 (ph = post-hole, black dots = potsherds) (after Mallory 1994: figure 4).

463

A3.25

Plan and section of Beacon Hill (Humerside) (after Moore 1966: figure 1).

463

A3.26

Plan of Beaghmore (Co. Tryone) (after May 1953: figure 1).

464

A3.27

Structure from Belle Tout (East Sussex). Possible post-holes are stippled and intrusive flint lying upon the surface of the natural flint rubble are represented in solid black (after Bradley 1970: figure 4).

465

A3.28

Structures from Belle Tout (East Sussex) (after Bradley 1970: figure 7).

466

A3.29

Plan of Bharpa Carinish (Western Isles) (after Crone 1993: figure 4).

467

A3.30

Occupation debris beneath a mound at Biggar Common (South Lanarkshire) (after Johnston 1997: illustration 6).

468

A3.31

Plan of Billown Circle (Isle of Man) (after Burrow 1997: figure A1.8).

468

A3.32

Plan of Billown Quarry 1995-97 (Isle of Man) (after Darvill 1999: figure 4).

469

A3.1 A3.2

A3.7

A3.10

A3.13

xvii

A3.33

Plan of structure 1 at Blairhall Burn (Dumfries And Galloway) (after Strachan et al. 1998: illustration 4).

470

Schematic plan of structures 2 and 3, pits and stake-holes at Blairhall Burn (Dumfries And Galloway) (after Strachan et al. 1998: illustration 6).

471

Plan of excavated features at Bolam Lake (Northumberland) (after Waddington 1999: figure 6.10).

472

A3.36

Brambletye Manor Farm (East Sussex) (after Tebbutt 1974: figure 2).

472

A3.37

Late ‘Neolithic’ and early ‘Bronze Age’ flint scatters and flint axes on the Brighton Downs (after Gardiner 1990: figure 8).

473

Brixworth (Northamptonshire): Detailed flint distribution of sites 26 and 27, 9-13 and 25. Each dot represents approximately 3 flints (after Martin and Hall 1980: figure 2).

474

A3.39

Pits and features from Bromfield (Shropshire) (after Stanford 1982: figure 4).

475

A3.40

Plan of Broome Heath (Norfolk) (after Wainwright 1972: figure 2).

476

A3.41

Plan of trench F1 at Bullock Down (East Sussex) (after Drewett 1978b: figure 2).

477

A3.42

Plan of trench B at Bullock Down (East Sussex) (after Drewett 1977b: figure 6).

478

A3.43

Plan of the excavation in field 1 at Burley Road (Leicestershire) (after Clay 1998: figure 8).

479

A3.44

Results from the fieldwalking survey at Burley Road (Leicestershire) showing the distribution of burnt flint. • 1-3; Ɣ 4-6;over 6 (after Clay 1998: figure 6).

480

Results from the fieldwalking survey at Burley Road (Leicestershire) showing the distribution of all lithics by standard deviation • 1-2ı; Ɣ 2-4ı;over 5ı (after Clay 1998: figure 2).

481

A3.46

Plan and sections of Cairnwell (Aberdeenshire) (after Rees 1997: illustration 5 and 6).

482

A3.47

The south west part of the cairn at Camster Long (Highland): Pre-cairn features (above); density of finds per square metre (below) (after Masters 1997: illustration 5).

483

Carlisle Airport (Cumbria): View from E, showing the eroded hollow (A-A) and the Dshaped structure (B), with one of the large stones of the alignment (C) removed from its modern pit. Scales are 2 metres (after Flynn 1998: Online).

484

A3.49

Plan and section of Carnaby Top 13 (Humberside) (after Manby 1974: figure 8).

484

A3.50

Plan and section of Carnaby Top 18 (Humberside) (after Manby 1974: figure 8).

485

A3.51

Section of pit at Carzield (Dumfries And Galloway) (after Maynard 1993: figure 2).

485

A3.52

Plan of pits at Cassington (Oxfordshire) (after Case 1982a: figure 68).

486

A3.53

Castell Bryn-Gwyn (Isle Of Anglesey): Plan (after Harding and Lee 1987: page 331)

487

A3.54

Plan and sections of features along the Caythorpe Gas Pipeline (Humerside) (after Abramson 1996: figure 3).

488

Plan and sections of pits along the Caythorpe Gas Pipeline (Humerside) (after Abramson 1996: figure 5).

489

A3.56

Plan of Cefn Cilsanws (Merthyr Tydfil) (Warwickshire) (after Darvill 1996: figure 6.10).

489

A3.57

Plan of Cefn Glas (Rhondda, Cynon, Taff) (after Clayton and Savory 1991: figure 2).

490

A3.58

Plan of Chippenham ‘Barrow’ 5 (Cambridgeshire) (after Darvill 1996: figure 6.10).

491

A3.59

Churston (Devon): (a) Distribution of finds, (b) Phases of prehistoric activity (after Parker Pearson 1981: figure 2).

492

A3.60

Plan of structure 1 at Clegyr Boia (Pembrokeshire) (after Williams 1953: figure 5).

493

A3.61

Plan of structure 2 at Clegyr Boia (Pembrokeshire) (after Williams 1953: figure 6).

493

A3.62

Corntown (The Vale of Glamorgan): Aerial photographic transcription from oblique sources. Grid at 50 metre intervals (after Burrow et al. 2001: figure 7.5).

494

Corse Law (South Lanarkshire): Lithic distribution: (a) flint; (b) chert (after Clarke 1989: illustration 4a, b).

495

Cowie Road (Stirling): Overall plan (after Rideout 1997: illustration 3).

496

A3.34 A3.35

A3.38

A3.45

A3.48

A3.55

A3.63 A3.64

xviii

A3.65

Cowie Road (Stirling): Main excavated area (after Rideout 1997: illustration 4).

496

A3.66

Excavation plan and diagrammatic section of Craike Hill (Humerside) (after Manby 1958: figure 2).

497

The distribution of later ‘Neolithic’ artefacts and structures in Cranborne Chase (Dorset) (after Barrett et al. 1991: figure 3.2).

498

Late ‘Neolithic’ and early ‘Bronze Age’ flint scatters, selected finds and constructions in Cranbourne Chase (Dorset) (after Gardiner 1990: figure 10).

499

Deer's Den, A96 Kintore and Blackburn By-Pass (Aberdeenshire): Plan of ‘Neolithic’ pits, showing number of pot sherds, chipped stone and vessel numbers (after Alexander 2000: illustration 5).

500

A3.70

Plan of Donegore (Co. Antrim) (after Sheridan 2001: figure 13.2).

501

A3.71

Artefact scatter at Downton (Wiltshire) (after Pollard 1999a: figure 5.2).

502

A3.72

Downton (Wiltshire): Plan of the excavation (1956-1957) and distribution of worked flint (after Higgs 1959: figure 2).

503

A3.73

Downton (Wiltshire): Stake-hole area (after Higgs 1959: figure 12).

504

A3.74

Downton (Wiltshire): Section of stake-holes and cooking holes (after Higgs 1959: figure 10).

505

A3.75

Dunragit (Dumfries and Galloway): Air-photograph plot with areas excavated 1999-2000 (above and 1999-2000 Trench A and AA (below) (after Thomas 2000b: figure 10.4).

506

A3.76

Durrington 68 (Wiltshire): The timber setting (after Pollard 1995b: figure 2).

507

A3.77

Density distribution map of all collected struck flint in East Berkshire (after Ford 1987a: figure 8.4).

508

A3.78

Plan of possible post-structures at Eaton Heath (Norfolk) (after Wainwright 1973: figure 4).

509

A3.79

Plan of Eilean Domhnuill, Loch Olabhat (Western Isles) (after Barclay 1996: figure 5.5).

509

A3.80

Reconstruction drawing of the final phase 1 at Eilean Domhnuill, Loch Olabhat (Western Isles) (after Armit 1991: page 287).

510

Elmton (Derbyshire): Distribution of fields yielding datable ‘Mesolithic’ (above) and Earlier ‘Neolithic’ (below) tools and debitage. Scale 1:50000 (after Knight et al. 1998: figure 3).

511

Elmton (Derbyshire): Distribution of fields yielding datable Later ‘Neolithic’/Early ‘Bronze Age’ tools and debitage. Scale 1:50000 (after Knight et al. 1998: figure 4).

512

Eye Hill Farm (Cambridgeshire): Overall worked and burnt flint densities (after Edmonds et al. 1999: figure 6).

513

A3.84

Eye Hill Farm (Cambridgeshire): General distributions (after Edmonds et al. 1999: figure 9).

514

A3.85

Plan of excavated pits and features at Fourknocks Ridge (Co. Meath) (after King 1999: figure 3).

515

A3.86

Plan of main excavated area at Glendhu (Co. Down) (after Woodman 1985c: figure 2).

516

A3.87

Plan of the late ‘Neolithic’ structural remains at Gorhambury (Hertfordshire) (after Neal et al. 1990).

517

A3.88

General plan of Grandtully (Perth and Kinross) (after Simpson and Coles 1990: illustration 3).

517

A3.89

Plan of Gwernvale (Powys) (after Darvill 1996: figure 6.4).

518

A3.90

Plan of structure at Gwithian (Cornwall). Phase 1 (after Megaw 1976: figure 4.3).

518

A3.91

Plan of structure at Gwithian (Cornwall). Phase 1 (after Megaw 1976: figure 4.4).

519

A3.92

Plan of Haldon (Devon) (after Willock 1933-6: plate 58).

520

A3.93

Plan and section of Hartendale (Humberside) (after Manby 1974: figure 29).

521

A3.94

Plan of Hazard Hill (Devon) areas A and B (after Houlder 1963: figure 3).

522

A3.95

Plan of Hembury (Devon) (after Darvill 1996: figure 6.4).

523

A3.96

Plan of High Peak (Devon) (after Pollard 1966: figure 1).

523

A3.97

High Peak (Devon): Section of Trench A, B and C (after Pollard 1966: figure 3, 4 and 5).

524

A3.67 A3.68 A3.69

A3.81 A3.82 A3.83

xix

A3.98

High Peak (Devon): Section of Cutting GA and H and Trench GB (after Pollard 1966: figure 6, 7 and 8).

525

Plan of the trial excavations at Hindwell Ash (Powys) (after Gibson 1999: figure 20).

526

Ground plan of the excavated remains beneath Hindwell Ash Barrow (Powys) (after Gibson 1996: figure 9.4).

527

A3.101

Plan of Hockwold-cum-Wilton (Norfolk) (Warwickshire) (after Darvill 1996: figure 6.10).

528

A3.102

Honey Hill (Cambridgeshire): Distributions (after Edmonds et al. 1999: figure 2).

529

A3.103

Honey Hill (Cambridgeshire): Plots (after Edmonds et al. 1999: figure 3).

530

A3.104

Horse Pastures (Derbyshire): The distribution of lithic within the two excavation trenches by artefact type (after Barnatt and Robinson 1998: figure 5).

531

Horse Pastures (Derbyshire): Sherd distribution within the two excavation trenches by period (after Barnatt and Robinson 1998: figure 6).

532

A3.106

Plan of Hurst Fen (Suffolk) (after Pollard 1999a: figure 5.4).

533

A3.107

Section of the deepest part of the valley sediments in Itford Bottom (East Sussex), Trench B (after Bell 1983: figure 10).

534

A3.108

Plan of Kemp Knowe (Humerside) (after Darvill 1996: figure 6.5).

535

A3.109

Kilncombe (East Sussex): The distribution of flint waste and retouched pieces (after Bell 1983: figure 6).

536

Kilncombe (East Sussex): The distribution of Beaker and other prehistoric pottery (after Bell 1983: figure 4).

537

A3.111

Kinbeachie (Highland): Excavated features in Trench 5 (Barclay et al. 2001: illustration 2).

538

A3.112

Kinbeachie (Highland): The rectangular structure in Trench 5 (Barclay et al. 2001: illustration 3).

538

Lithic artefact distribution at King Barrow Ridge (Wiltshire) (after Entwistle and Richards 1987: figure 3.5).

539

Plans of structures at Knap of Howar (Orkney Islands): Secondary phase of structure 2 with block entrance, primary phase of structures 2 and 1 (after Ritchie 1982: figure 3).

540

A3.115

Area of occupation at Knowth (Co. Meath) (after Eogan 1986: figure 85).

541

A3.116

Plan of structures 6 and 8 at Knowth (Co. Meath) (after Cooney 2000a: figure 3.2).

542

A3.117

Timber circle from Knowth (Co. Meath) (after Eogan and Roche 1999: illustration 11.1).

542

A3.118

The palisade trenches of the enclosure at Knowth (Co. Meath) (after Sheridan 2001: figure 13.3).

543

Plan and section of a remnant of Structure A at Lamb’s Nursery, Dalkeith (Midlothian) (after Cook 2000b: illustration 3).

544

Sectional profiles of ‘Neolithic’ pits Context 90 (top) and Context 60 (bottom) at Lamb’s Nursery, Dalkeith (Midlothian) (after Cook 2000b: illustration 4).

544

A3.121

Plan of Langford Lodge (Co. Antrim) (after Waterman 1963: figure 2).

545

A3.122

Plan of Levens Park (Lancashire) (after Darvill 1996: figure 6.10).

546

A3.123

Plan of the structure at Links of Noltland (Orkney Islands), two phases showing (after Clarke and Sharples 1985: figure 4.5).

546

A3.124

Plan of Lismore Fields (Derbyshire) (after Garton 1987: page 252).

547

A3.125

Plan of Little Paxton (Cambridgeshire) (after Darvill 1996: figure 6.10).

547

A3.126

Lyles Hill, showing the ‘Bronze Age’ and later banked enclosure, the cairn, the position of the 1936-37 trench (“x”) and the 1987-88 trench (“A,B”). Insert: The palisade trenched and other features in part of B (after Sheridan 2001: figure 13.4).

548

A3.127

Earliest features at Machrie Moor (North Ayrshire) (after Haggarty 1991: illustration 2).

549

A3.128

Timber structures at Machrie Moor (North Ayrshire) (after Haggarty 1991: illustration 5).

549

A3.99 A3.100

A3.105

A3.110

A3.113 A3.114

A3.119 A3.120

xx

A3.129

Stake lines at Machrie Moor (North Ayrshire) (after Haggarty 1991: illustration 7).

550

A3.130

Inserted burials and other features at Machrie Moor (North Ayrshire) (after Haggarty 1991: illustration 16).

550

A3.131

Section at Mad Mans Window I (after Woodman 1992: figure 2).

551

A3.132

Section through chipping floor at Mad Mans Window IIB (after Woodman 1992: figure 6).

551

A3.133

Schematic plan of Mad Mans Window III chipping floor (after Woodman 1992: figure 8).

552

A3.134

(a) Section through Mad Mans Window III chipping floor; (b) Section through pit (after Woodman 1992: figure 9).

553

Total numbers of flakes from the first season’s fieldwalking at Maddle Farm (Berkshire) (after Tingle 1987: figure 7.3).

554

Distribution of implements from the first season’s fieldwalking at Maddle Farm (Berkshire) (after Tingle 1987: figure 7.4).

555

Middle Avon Valley (Hampshire): Density variation for lithic artefacts within the survey area (after Light et al. 1994: figure 10).

556

A3.138

Plan of Mill Street, Driffield (Humerside) (after Darvill 1996: figure 6.4).

557

A3.139

Mount Pleasant (Derbyshire): Fieldwalking data collected in the autumn of 1983. One symbol equals one artefact (after Garton and Beswick 1983: figure 2).

558

Mount Pleasant (Derbyshire): Area D artefact scatter (after Garton and Beswick 1983: figure 3).

559

Mount Pleasant (Derbyshire): Area A artefact scatter, (A) shows the distribution of pottery and tools, (B) shows the distribution of burnt flakes (after Garton and Beswick 1983: figure 4).

560

The ‘Neolithic’ phase at Mount Pleasant Farm (Rhondda, Cynon, Taff) (after Lane 1986: figure 1).

561

A3.143

Distribution of chert at Nether Exe (Devon) (after Silvester et al. 1987: figure 2).

562

A3.144

Distribution of flint at Nether Exe (Devon) (after Silvester et al. 1987: figure 3).

563

A3.145

Excavated feature from Newton (Argyll and Bute) (after McCullagh 1989: figure 4).

564

A3.146

Plan of the structure at Newtown (Co. Meath) (after Cooney 2000a: figure 3.1).

565

A3.147

Plan and section of North Carnaby Temple 2 (Humberside) (after Manby 1974: figure 15).

565

A3.148

Plan and section of North Carnaby Temple 7 (Humberside) (after Manby 1974: figure 15).

566

A3.149

Plan and section of North Carnaby Temple 9 (Humberside) (after Manby 1974: figure 15).

567

A3.150

Plan and section of North Carnaby Temple 12 (Humberside) (after Manby 1974: figure 15).

567

A3.151

Plan of North Carnaby Temple 15 (Humberside) (after Manby 1974: figure 16).

568

A3.152

Sketch plan of North End, Walney Island (Cumbria) (after Barnes 1956: figure 1).

569

A3.153

Density distribution map of all collected struck flint at North Stoke (Oxfordshire) (after Ford 1987a: figure 8.3).

570

A3.154

Plan of Padholm Road, Fengate (Cambridgeshire) (after Darvill 1996: figure 6.5).

571

A3.155

Plan of Park Farm (Isle of Man) (after Burrow 1997: figure A1.19).

571

A3.156

Late ‘Neolithic’ flint scatters in the Peak District (Derbyshire) (after Bradley and Hart 1983: figure 5).

572

Flint scatters with barbed-and-tanged arrowheads in the Peak District (Derbyshire) (after Bradley and Hart 1983: figure 7).

572

Flint scatters with leaf-shaped arrowheads in the Peak District (Derbyshire) (after Bradley and Hart 1983: figure 3).

572

A3.159

Penhale Point (Cornwall): Flint debitage distribution (after Smith 1988: figure 15).

573

A3.160

Penhale Point (Cornwall): Distribution of flint retouched pieces and other stone objects (after Smith 1988: figure 17).

574

A3.135 A3.136 A3.137

A3.140 A3.141

A3.142

A3.157 A3.158

xxi

A3.161

Plan of Playden (East Sussex) (after Cheney 1935: figure 2).

575

A3.162

Poldowrian (Cornwall): ‘Neolithic’ features (after Smith and Harris 1982: figure 6).

576

A3.163

Plan and sections of the pits at Puddlehill (Bedfordshire) (after Field et al. 1964: figure 5).

577

A3.164

Rackham (West Sussex): Plan of features (after Holden and Bradley 1975: figure 2).

578

A3.165

Rackham (West Sussex): Distribution of burnt flint (after Holden and Bradley 1975: figure 3).

578

A3.166

Rackham (West Sussex): Distribution of charcoal (after Holden and Bradley 1975: figure 4).

579

A3.167

Rackham (West Sussex): Distribution of cores and hammerstones (after Holden and Bradley 1975: figure 5).

579

Rackham (West Sussex): Distribution of waste flakes (after Holden and Bradley 1975: figure 6).

580

Rackham (West Sussex): Distribution of flint implements (after Holden and Bradley 1975: figure 7).

580

Lake or marshland occupation at Rathjordan (Co. Limerick): Plan and cross-section (after O’Sullivan 1998: figure 10).

581

A3.171

Excavation plan of Rhos-y-Clegyrn (Pembrokeshire) (after Lewis 1975: figure 3).

582

A3.172

Plan of the structure at Rinyo (Orkney Islands) (after Simpson 1971: figure 26).

583

A3.173

Plan and sections of pits at Risby Warren (Lincolnshire) (after Riley 1957: figure 2).

584

A3.174

Flint distributions at Rodney Stoke (Somerset) (after Broomhead 1991: figure 2).

585

A3.175

Plan and section of Rudston Wold East Reservoir 3 (Humberside) (after Manby 1974: figure 5).

586

Salcombe Hill (Devon): Prehistoric features from a sketch map by Hutchinson, 1909 (after Pollard and Luxton 1978: figure 1).

587

A3.177

Plan of Sale’s Lot (Gloucestershire) (after Darvill 1996: figure 6.5).

588

A3.178

Slipper Low Farm (Derbyshire) (after Garton and Kennett 1996: figure 2).

589

A3.179

South Dorset Ridgeway Ridge: ‘Neolithic’ and ‘Bronze Age’ landuse and occupation (after Sharples 1991: figure 28).

590

Late ‘Neolithic’ and early ‘Bronze Age’ flint scatters, flint axes and selected finds on the south-east Sussex Down between Lewes and Eastbourne (after Gardiner 1990: figure 9).

591

Flintwork and settlement evidence in southern Anglesey. Contours in metres above OD (after Holgate and Smith 1981: figure 2).

592

Stonehenge Environs (Wiltshire): Distribution of all worked flint (after Richards 1990: figure 10).

593

Stonehenge Environs (Wiltshire): Distribution by weight of prehistoric pottery (after Richards 1990: figure 16).

594

Storey’s Bar Road, Fengate (Cambridgeshire): Plan of the late ‘Neolithic’ and ‘Bronze Age’ features. Contours in metres OD after removal of topsoil (after Pryor 1978: figure 6).

595

A3.185

Plan of Stretton-on-Fosse Site 5 (Warwickshire) (after Darvill 1996: figure 6.5).

596

A3.186

Plan of Swarkestone (Derbyshire) (after Simpson 1971: figure 25).

596

A3.187

Plan of the structure at Tankardstown 1 and 2 (Co. Limerick) (after Cooney 2000a: figure 3.1).

597

Plan of trench 1: ‘Neolithic’ features at Templecorran (Co. Antrim) (after Crothers 2000: figure 3).

597

Plan of trench 1: Later ‘Neolithic’ features at Templecorran (Co. Antrim) (after Crothers 2000: figure 4).

598

Plan of trench 2: ‘Neolithic’ features at Templecorran (Co. Antrim) (after Crothers 2000: figure 7).

598

Plan of structure at Thirlings (Northumberland) (after Waddington 1999: figure 7.8).

599

A3.168 A3.169 A3.170

A3.176

A3.180 A3.181 A3.182 A3.183 A3.184

A3.188 A3.189 A3.190 A3.191

xxii

A3.192

The distribution of worked lithics from surface collection across Thornborough (North Yorkshire) (after Harding 2000: figure 2.5).

600

General plan of the pre-mound features at Upper Ninepence (Powys) (after Gibson 1999: figure 31).

601

The distribution of ‘Neolithic’ structures and flint finds in the Walton Basin (Powys) (after Gibson 2000: figure 3.4).

602

A3.195

Plan of occupation debris 1 at Willington (Derbyshire) (after Wheeler 1979: figure 3).

603

A3.196

Plan of occupation debris 2 at Willington (Derbyshire) (after Wheeler 1979: figure 5).

604

A3.197

Windy Ridge (Co. Antrim): Excavation of trench 8, showing features recovered in Levels 1 and 2 (after Woodman et al. 1994: figure 2).

605

Windy Ridge (Co. Antrim): Excavation of trench 8, showing features recovered in Levels 3 and 4 and distribution of pottery (after Woodman et al. 1994: figure 3).

606

Windy Ridge (Co. Antrim): Excavation of trench 8, showing distribution of struck flint (after Woodman et al. 1994: figure 4).

607

Windy Ridge (Co. Antrim): Excavation of trench 8, showing distribution of quartz, burnt flint, cores and retouched tools (after Woodman et al. 1994: figure 5).

608

Plan and section of structure at Woodhead (Cumbria) (after Hodgson 1940: figure 7).

609

A3.193 A3.194

A3.198 A3.199 A3.200 A3.201

xxiii

List of tables 5.1

Percentages of weathered and root-marked human skeletal material within categories of types of deposit (after McKinley forthcoming).

88

5.2

Radiocarbon dated skeletal material associated with cut marks (after McKinley forthcoming).

92

5.3

Surface modification of the human skeletal material, expressed as a percentage of the total assemblage (after Evans 2000:figure 7.3).

98

Distribution of cut marked bone by location and phase, expressed as a percentage of the total for each individual location (after Evans 2000:figure 7.9).

98

Distribution of cut bone within age categories, shown as a percentage of the total number of cut bone (after Evans 2000: figure 7.11).

99

Deposition of human skeletal material documented by recent investigations at ‘Mesolithic’ funerary contexts in northern France and Belgium (after Scarre 2002a: table 1).

138

Summary of raw material, technological and typological characteristics of later ‘Mesolithic, earlier ‘Neolithic’ and later ‘Neolithic’ flint assemblages (after Holgate 1988a: table 5.7).

177

Chronologically diagnostic traits of late ‘Mesolithic’ to later ‘Neolithic’ flintwork in the upper Thames Valley (after Bradley and Holgate 1984: table 1).

178

Expected assemblage characteristics for ‘domestic’ and ‘industrial’ area assuming a policy of extra home-range production (after Schofield 1991a: table 10.1).

181

Sr and Pb isotope ratios and compositions for the tooth samples. En - enamel, De - dentine. Strontium isotope ratios were undertaken by TIMS, external reproducibility of standards was 0.0045% (2ı, n=10). Pb isotope ratios were undertaken by PIMMS, errors are 0.014% for 207 Pb/206Pb and 0.024% for the 208Pb/206Pb (2ı, n=27). Isotope dilution errors also account for sample heterogeneity and are therefore estimated as 14% for Pb and 10% for Sr (after Montgomery et al. 2000: table 1).

185

Strontium and lead isotope ratios and compositions for the Monkton chalk (MC) and soil (MS) leachates. Strontium isotope ratios were undertaken by TIMS, external reproducibility of standards was 0.0045% (2ı, n=10). Pb isotope ratios were undertaken by PIMMS, errors are 0.014% for 207Pb/206Pb and 0.024% for the 208Pb/206Pb (2ı, n=27) (after Montgomery et al. 2000: table 2).

187

Outline summary proposed by Whittle of phases in the ‘Neolithic’ of the Avebury region. Information from this will be referred to throughout the chapter (after Whittle 1994a: table 7).

202

Dates of long mound construction and theoretical minimum and maximum time until woodland regeneration. All dates in calibrated BP and at the one standard deviation level from Whittle et al. 1993. Relevant laboratory codes are: Easton Down (OxA-3762), Millbarrow (BM-2730 and BM-2729), South Street (BM-356 and BM-358a) and Beckhampton Road (BM-506b) (after Davies and Wolski 2001: table 1).

210

Theoretical maximum and minimum distances of woodland (in metres) from four long mounds in the Avebury region according to different molluscan species (after Davies and Wolski 2001: table 3).

210

Human stable isotope values from radiocarbon dated human bone and tooth samples from the Avebury region from the middle of the seventh to the late fifth millennium BP. The values presented here include work carried out exclusively for the study of human diet and a sample of į13C (‰) values gathered in the course of radiocarbon dating analysis. The locations are arranged in chronological order from oldest to youngest radiocarbon dates. Sources of the data: (5) Richards and Hedges 1999a, (8) Richards 2000a (data drawn from Appendix 1). Individuals who get a predominant (>90%) amount of their protein from marine sources will have į13C values close to -12‰. Those who have no marine protein in their diet will have į13C values closer to -20‰. Those who derive 50% of their protein from terrestrial, and 50% from marine sources, will have į13C values in the middle, about -16‰. Measuring the į15N provides information on the relative importance of plant and animal protein sources. There is an increase of about 3‰ in the į15N each step up the food chain. Therefore, if plants have an average į15N of about 3‰, herbivores that consume those plants have į15N values of 6‰ and carnivores consuming those herbivores will have į15N values of about 9-10‰.

212

5.4 5.5 5.6 6.1 6.2 6.3 6.4

6.5

7.1 7.2

7.3

7.4

xxiv

Acknowledgements This book is the result of doctoral research carried out at the University of Sydney, Australia and completed in 2002. During this time I have been fortunate enough to be supervised by Roland Fletcher. His help, criticism and guidance have been invaluable and I thank him for all he has done. My thanks to Aedeen Cremin for introducing me to prehistoric Europe, for all her help and the many books and journals she has given to me over the years. Thanks also to Peter Rowley-Conwy, Joshua Pollard, Tim Murray, Peter White, Garin Down, Richard Fletcher and Tigger Wise for various comments on drafts. In particular, I would like to thank Bronwyn Sneesby for casting her editorial eye over my work, even the tedious bits. I am very thankful for the help and encouragement given by Peter Rowley-Conwy (University of Durham), Robert Layton (University of Durham), Thomas Dowson (University of Durham), Andrew Chamberlain (University of Sheffield), Joshua Pollard (University of Wales College, Newport), John Barrett (University of Sheffield), Joanna Brück (University College Dublin), John Chapman (University of Durham) Michael Richards (University of Bradford), Julian Thomas (University of Manchester), Robin Torrance (Australian Museum), Bernard Knapp (University of Glasgow), Alasdair Whittle (Cardiff University), Richard Bradley (University of Reading) and Bob Chapman (University of Reading). Many people have been kind enough to supply me with their published and unpublished work to help with my research. Thanks must go to the following people: Michael Richards (University of Bradford), Christopher Ramsey (Oxford Radiocarbon Accelerator Unit), Paul Ashmore (Historic Scotland), Robert Mowat (Historic Scotland), Andrew Chamberlain (University of Sheffield), Alex Bayliss (English Heritage), Anna L. Brindley (Groningen Institute for Archaeology), Frances Healy (University of Newcastle), Gerry McCormac (Queen’s University Belfast), George Lambrick (Council for British Archaeology), Göran Burenhult (Gotland University College), Mark Hinman (Cambridge Archaeological Unit), Emma Wager (University of Sheffield), Michael Wysocki (University of Central Lancashire), Chris Scarre (The McDonald Institute for Archaeological Research), Joshua Pollard (University of Wales College, Newport), Alex Gibson (University of Bradford), Darden Hood (Beta Analytic Inc.), Gordon Cook (SURRC Radiocarbon Dating Laboratory), Jacqui Mulville (Oxford University Museum), Mark Robinson (Oxford University Museum), Rick Schulting (Queen’s University Belfast) John Evans (Cardiff University), John Schofield (English Heritage), Michelle Thompson (Queen’s University Belfast), Mike Pitts, Paul Budd (University of Durham and Archaeotrace Limited), Paula Reimer (Lawrence Livermore National Laboratory), Rachel E. Pope (University of Durham), Robert Hedges (Research Laboratory for Archaeology, Oxford), Steven J. Mithen (University of Reading), Susan Kus (Rhodes College), Mark Edmonds (University of Sheffield), Tim Murray (La Trobe University) and EmmaJayne Evans. Many institutions, databases and libraries were also consulted in the course of my research. I would like to thank ARChaeological Holdings SEARCH System (ArcHSearch), The British Library and the Institute of Archaeology Library in London, the Haddon Library and University Library at the University of Cambridge, Hartley Library at the University of Southampton, Ashmolean (Sackler) Library, the Bodleian Library and the Radcliffe Science Library at the University of Oxford, the Main Library Whiteknights at the University of Reading, the Dorset House Library at the Bournemouth University, the Main Library at the University of Sheffield, the Main Library at the University of Durham, CANMORE (Computer Application for National MOnuments Record Enquiries), National Library of Scotland and the Royal Commission on the Ancient and Historical Monuments of Scotland Library in Edinburgh, the Main Library at the University of Edinburgh, the University Library at the University of Glasgow, Royal Commission on Ancient and Historical Monuments in Wales (CARN), the Arts & Social Studies Library at the Cardiff University, National Library of Ireland in Dublin, the Main Library at Queen’s University Belfast, the James Hardiman Library at the National University of Ireland in Galway, the Boole Library at University College Cork, the Main Library at University College Dublin, the National Library of Australia in Canberra and the J.B. Chifley Library, the W.K. Hancock Library and the R.G. Menzies Library at the Australian National University in Canberra. In particular, I would like to thank Fisher Library at the University of Sydney, especially the Inter-library Loans section which supplied me with many books and journals from around Australia and overseas. The setting up of the Online Database took also time and effort by a number of people. Thanks goes to Richard Fletcher for helping me initially set up my database, the Archaeological Computing Lab (ACL) at the University of Sydney, Garin Down, Catherine Hardman from the Archaeology Data Service (ADS) and the people from BrainJuice for designing and helping construct my Online Database, especially Daniel Geleris. I am thankful to the participants at the Archaeology Seminars at the University of Sydney and those at the Stones and Bones Conference in Sligo, May 2002 for their helpful contributions. Further I would like to thank the many students who sat through one of my lectures and have offered their opinions and criticisms, all of which was much appreciated. The Carlyle Greenwell Bequest and the PhD Students’ Research Support Scheme, University of Sydney provided funding for overseas research. Research for this book was undertaken while in receipt of funding from the Australian Postgraduate Award, provided by the Commonwealth Government of Australia. xxv

Unfailing support has been provided by my mother and father, my three sisters, Bernadette, Therese and Claudine, my brother-in-laws, Garin and Rod, and my friends, in particular Michael and Dennis, all of who have learnt to live with the prehistoric dead from Britain and Ireland. I would also like to express my extreme gratitude to Monique Spiller, who gave me my dream, and Melissa Phillips, who encouraged me to pursue it, to both I am forever indebted. Finally, I would like to thank Bronwyn for her love, friendship and inspiration. Martin King March 2003

xxvi

Abstract A re-examination of the actual patterning of human skeletal material, animal bones and artefacts across the British and Irish ‘Mesolithic’ and ‘Neolithic’ landscape has led to a redefinition of the understanding of the dead of many species, including humans. Dichotomised categories which are generally assumed to be in some way universal, such as ‘tomb’/‘house’, ‘ritual’/‘domestic’ and ‘human’/‘animal’, cannot simply be considered as directly applicable to the interpretation of the operational characteristics of past human societies. Rather than seeking an explanation derived from standard social theory categories, different questions need to be developed about ‘materiality’ – the actuality of the pattern on the ground. The use of standard social theory that dominates the current interpretative archaeologies appears to be plausible and reasonable, because it identifies human social meaning that we know to be possible. However, this relies on a logically invalid principle that there is some universal correspondence between the material and the social. Maintaining the use of standard social theory drastically reduces the scope of human social behaviour which can possibly be identified. The risk of this prevalent practice is that archaeological assemblages and contexts are liable to be a pale imitation of the findings and declarations of the social sciences. The large corpus of literature, relating to occupation, mobility, clearing woodland, construction, the deposition of artefacts and the distribution and treatment of human and animal skeletal material, illustrates continuity in human social behaviour across the ‘Mesolithic’/‘Neolithic’ divide. The mobility argument stresses that a larger operational scale is also important. The dispersed nature of skeletal remains and artefacts across the landscape is viewed as the consequence of mobile occupation and random incidents of death and discard across the landscape. This overall ‘patterning’ need not be seen as the result of a specific elaborate ‘ritual’. Rather the dispersed human and non-human skeletal remains can be seen as the ‘fall-out’ from the operations of the dispersed occupation system. This study illustrates that archaeological assemblages and contexts have the potential to bear witness to human behaviour that has no direct parallel with historically documented or observed examples. Such behaviour should not be marginalised or normalised into a form that falls comfortably within our realms of understanding. Rather, archaeological assemblages and contexts should be studied as a suite of unparalleled signatures of past human behaviour. I argue that this should be the fundamental role of archaeology. We must expect the unexpected – pursuing a logic that allows unique human behaviour which has materialised in archaeological assemblages and contexts over the two to three million years of hominid cultural existence to be recognised.

xxvii

Notes Note on the Online Database The Online Database is catalogued at the Archaeology Data Service (ADS) under Special Collections and can be accessed at: http://ads.ahds.ac.uk/catalogue/specColl/lwtd/. Note on regions and administrative localities After each location or context mentioned in the text and the appendices its region or administrative locality is referred to in brackets. I have referred to the new administrative boundaries that have been defined across Britain and Ireland during the 1990s. I have maintained this consistency throughout the text, the appendices and the Online Database. Note on radiocarbon dates All radiocarbon dates have been calibrated using the Radiocarbon Calibration Program 4.2 from the University of Washington, based on the INTCAL98 curve, and are presented by the conventions outlined by Stuiver and Reimer 1993 and Stuiver et al. 1998. For purpose of simplifying radiocarbon dates in the text, the convention of rounding dates to the nearest 10 years for samples with standard deviations (± error in the age) greater than 50 years has been followed. However, this has not been done for radiocarbon dates in the appendices and the Online Database. To free up the text long strings of calibrated radiocarbon dates have be footnoted. Uncalibrated dates are expressed as bp (usually in brackets with the laboratory code) and calibrated dates are expressed as BP at two standard deviation (BP = Before Present, 0 BP = 1950). Note on bibliography The bibliography contains the references mentioned in the text, appendices and those in the Online Database.

xxviii

Preface This book is organised into three parts. Part one includes the introduction and two chapters discussing the changing perspectives of the ‘Mesolithic’ and ‘Neolithic’, in particular the changing way that the two periods have been viewed in relation to economy and subsistence. Continuity in economic and subsistence patterns between the ‘Mesolithic’ and ‘Neolithic’ of Britain and Ireland will be examined in detail. The evidence illustrates that ‘Neolithic’ populations were continuing a long and well-established management of plants and animals. Part two contains four chapters, and begins with a theoretical chapter which outlines and overviews the past and current use standard social theory, arguing that the use of such an approach drastically reduces the scope of human social behaviour which can possibly be identified. The following chapters then look at the evidence for human social behaviour relating to occupation, mobility, clearing woodland, construction, the deposition of artefacts and the distribution and treatment of human and animal skeletal material. The large corpus of literature illustrates the continuity that is present in the empirical evidence between the ‘Mesolithic’ and ‘Neolithic’. The same kind of practices occur to varying degrees in both ‘periods’, but some have been viewed as central to characterising the ‘Neolithic’ as a phenomenologically distinct entity. For example, while the ‘Neolithic’ witnessed some constructions using massive stones, this scale of ‘monumentality’ is actually the exception in the ‘Neolithic’. Exceptionally large constructions are not the dominant scale of constructions down to the late fifth millennium BP, rather, if the evidence is thoroughly examined, the majority of constructions in post-‘Neolithic’ contexts are more comparable in size to those in pre‘Neolithic’ contexts. As will be illustrated, the phenomenologically distinct ‘Neolithic’ has been produced through a selective viewing of the empirical evidence, the use of specific ethnographic and ethno-historic analogies and through an inherent bias that the foundations of ‘our’ modern view of the landscape and of death (ancestors) was constructed in the ‘Neolithic’. Part three, which contains a case study chapter and the conclusion, applies the arguments and observations made in Parts one and two to a case study of the Avebury region. The case study documents the archaeological and environmental data gathered over the last few centuries which identifies continuity in human social behaviour across the ‘Mesolithic’/‘Neolithic’ divide, from the early tenth to the late fifth millennium BP. The case study concludes that a complex and intermeshed patterning of human activity occurred across the landscape from the early tenth to the late fifth millennium BP and that the ‘Mesolithic’ and ‘Neolithic’ represent one tradition of ‘action’, whatever specific verbalised meanings may have been involved. Finally, the book concludes that the current discourse’s interpretive approaches adversely affect our ability to identify past human social behaviour which has no direct parallel in either the observed and/or the documented social life of the present or the recent past.

xxix

PART I INTRODUCTION: CONSTRUCTING THE PAST postprocessual/social and ideological approaches (see Bradley 1984a:11, 1997a; Edmonds 1997:99; Gijn and Zvelebil 1997; Thomas 1988a:59; Waddington 2000a:33). Thus, elements of continuity are lost as very different questions and ways of interpreting empirical data are pursued on either side of the designated chronological boundary. As Bradley (1984a:11) put it: in the literature as a whole, successful farmers [‘Neolithic’ populations] have social relations with one another, while hunter-gatherers [‘Mesolithic’ populations] have ecological relations with hazelnuts.

We still interrogate the deep past with essentially the same theoretical instruments which we use to make sense of the present. (Murray 1997:459) Complex and varied combinations of human skeletal material, and for that matter animal bones and artefacts, distributed across the British and Irish landscape have long been at the heart of archaeological investigation and interpretation in ‘Neolithic’1 studies. Yet in ‘Mesolithic’ contexts, less has been made of a similarly complex pattern. The continuity between the ‘Mesolithic’ and ‘Neolithic’ has, in practise, been obscured due to the different theoretical and interpretative approaches which directly affect the questions that are asked of the data from each ‘period’.

Imposing a notion of strict linear change onto archaeological assemblages and contexts has unduly set boundaries to which archaeologists feel obliged to adhere, which in practice impedes our understanding of prehistory (see Barclay et al. 1996:2). A division between the ‘Mesolithic’ and ‘Neolithic’ is created, despite there being no empirical reason to do so. In addition, it is important to note that the separation between processual and postprocessual archaeologies is an artificial one, based on associated theoretical baggage rather than particular archaeological method and practice (for example see appraisal by Tschauner 1996). In addition, while the two claim different epistemological reasoning, they both still consider that “meaningful explanation or interpretation of prehistoric human behaviour should mirror the forms found in contemporary social theory” (Murray 1997:456, also see Fletcher 1995:Chapter 1).

In ‘Neolithic’ contexts, archaeological assemblages of human skeletal material have long been viewed as ‘special’ and deeply symbolic, especially those in stone chambered constructions (‘tombs’) and earthen and timber mounds (‘barrows’). This apparent importance of human skeletal material, particularly in these constructions, has led to interpretations that have taken a selective view of the empirical evidence. This selective view has formed the basis for recent interpretations of the ‘Neolithic’. The very different theoretical baggage employed is just one part of the deep conceptual separation between the scholarly traditions of ‘Mesolithic’ and ‘Neolithic’ studies. In Britain and Ireland ‘Mesolithic’ studies are dominated by processual/ecological approaches and ‘Neolithic’ studies are now dominated by

This archaeological divide between the two discourses has been further exaggerated due to the recent theoretical and philosophical direction taken by some ‘Neolithic’ scholars. Over the last ten to fifteen years the ‘Neolithic’ has been identified as a phenomenological event, a view championed by scholars such as Thomas, Tilley, Hodder, Edmonds, Barrett and Whittle. Most recently, the contextual view of the ‘Neolithic’ in terms of phenomenological meaning has stressed the increased importance of the landscape and the distinctive location of the ancestors within that landscape in terms of phenomenological meaning. The ‘Neolithic’ is characterised as a time when the population developed, and thus materialised, a greater concern for the dead and their symbolic meaning (see Barrett 1994; Edmonds 1999a; Hodder 1990a; Thomas 1999a, 1999b; Tilley 1994; Whittle 1996a). Rather than being an attribute of the ‘Neolithic’, we need to consider that this interpretation has been developed in order to facilitate the creation of yet another supposedly ‘distinctive’ attribute to define the ‘Neolithic’, a practice that has defined the history of ‘Neolithic’ discourse since the nineteenth century. This has led to the situation we now face, where a phenomenologically distinct ‘Neolithic’ is being formulated, as yet another attempt to segregate the

1. I support recent criticisms of the use of the term ‘Neolithic’ (see Czerniak 1998; Pluciennik 1998; Thomas 1993b, 1997a; Zvelebil 1998a), so inverted commas are employed to exclude its value-laden meaning from my interpretation. Further, a similar convention will be employed in discussing the ‘Mesolithic’ period and other prehistoric chronological terms. For convenience, chronologically the ‘Mesolithic’ in Britain and Ireland approximately dates from the early tenth to the middle of the seventh millennium BP and the ‘Neolithic’ dates from the middle of the seventh to late fifth millennium BP. However, the time of the ‘Mesolithic’/‘Neolithic’ transition in Britain and Ireland can vary in the literature to anytime during the seventh millennium BP.

1

interpretative tools that have been developed to cope with human social behaviour observed in contemporary or recent contexts. I will seek instead to formulate an interpretation that does not convert the empirical evidence into verbal meanings, but rather develops different questions about ‘materiality’ – the pattern on the ground – and different views of what they ‘mean’ in terms of past human actions rather than in terms of verbalised meanings of social value. The approach is specifically sensitive to the information on human social behaviour that we actually encounter in the archaeology of Britain and Ireland from the early tenth to the late fifth millennium BP. Bailey (1983:165) summarised the situation that prehistoric archaeology finds itself in as follows: Above all, new dating methods have demonstrated that human cultural history extends over a timespan of a least two million years. This poses in a new way the issue of how we are able to make use of knowledge of the past, what questions we should ask of it, and whether by archaeological investigations of human activities over this timespan, we can learn something new about human nature not available from other sources.

‘Neolithic’ as somehow distinct from the preceding ‘Mesolithic’. I argue that the current chronological divisions and the constraints this places on archaeological research have more to do with the differences between the practices of ‘Mesolithic’ and ‘Neolithic’ archaeology than it does with the current archaeological evidence. A revised picture of the relationship between the ‘Mesolithic’ and ‘Neolithic’ has been developing particularly over the last thirty years, mainly pursued by scholars whose major interest is in ‘Mesolithic’ studies. Economic, technological and residential continuity between the ‘two periods’ is now generally accepted. Despite such arguments for continuity, two very different social views of the world are devised for the ‘Mesolithic’ and the ‘Neolithic’. Thomas has argued that the current approaches taken in ‘Neolithic’ studies are “more likely to consider human beings as purposive subjects, acting in pursuit of sociallydefined goals” (1988a:59). Further, he has suggested that if ‘Neolithic’ theoretical and interpretive tools were injected into ‘Mesolithic’ studies, a more balanced and continual picture of European prehistory would be forthcoming, rather than currently where ‘Mesolithic’ archaeology “represents a period of the human past as a bland and stable epoch” (Thomas 1991b:19). Thomas (1988a:64) is in essence arguing for a ‘Neolithicising’ of the ‘Mesolithic’, which: can best be done by injecting a little history and human intentionality into the cybernetic wasteland of the Mesolithic, rather than reducing the Neolithic population of Europe to the status of automata. A suggestion to which Steven Mithen (1991, 1999a), amongst others, has rightly taken offence. However, some recent ‘Mesolithic’ studies seem to have followed Thomas’ suggestion, with work by Cummings (2000), Pollard (2000b) and Warren (2000) taking a phenomenological approach to ‘Mesolithic’ contexts.

‘Neolithic’ archaeology has fundamental theoretical and interpretive flaws that impede enquiry about the past, so much so that unparalleled and important archaeological phenomena have remained marginalised for the sake of explanatory familiarity. The current archaeological discourse is unable to break out of accepted forms of archaeological interpretation due to the continual and habitual use of standard social theory – that is the use of the terms and expressions from the suite of social anthropology/sociology/history about observed or historically documented social life. While it is argued that “it is hard to envisage how we [archaeologists] could begin to make sense of past, different worlds without it [standard social theory categories]” (Whittle 1998:853, also see Whittle 1997b:142-64; 2002), I would argue that the current way in which analogy is employed makes the identification of past and different human social behaviour problematic. The assumption has been that the unique and the non-Western/modern can be discovered in archaeological assemblages and contexts using analogies from standard social theory (see Barrett 1994; Hodder 1982a, 1990a; Tilley 1996a; Verhoeven 2000; Whittle 1996a). Further, some argue that the use of ethnographic evidence is important if archaeology is going to identify past human social behaviour rather than just recognising patterns in archaeological assemblages and contexts (Gould and Yellen 1991:283-4).

The current theoretical and interpretative approaches of ‘Neolithic’ archaeology produce a very selective view of the empirical evidence. While the scholars concerned with a phenomenological ‘Neolithic’ argue that they take a more humanising view I will argue that what is discussed in ‘Mesolithic’ studies is by no means less humanising. I will, and have argued (King 2001), that a more productive step would be a ‘Mesolithicising’ of the ‘Neolithic’. ‘Mesolithic’ archaeology is more concerned with interpretation of human social behaviour through careful and systematic analysis of the archaeological empirical data at hand to identify patterns of behaviour. The current ‘Neolithic’ archaeological practice of social explanation involves the creation of verbalised meanings that are supposed to have been associated with past human social behaviour, a practice which has been criticised by Fletcher (1989, 1995, 1996, 2002, in press) and Murray (1993, 1997, 1999, 2001, 2002).

There is no reason to expect that the range of human social behaviour that is documented ethnographically and ethno-historically is in any way sufficiently comparable to, or representative of, the social behaviour of prehistoric populations. Ethnographic and historically documented social life from an extremely limited time span and region cannot be the sum of all human behaviour. This methodology therefore allows archaeological interpretations to perceive only a tiny fraction of the great

The history of archaeological investigation has seen a succession of efforts to adopt methodologies and 2

‘Neolithic’ discourse is systematically and consistently ‘normalising’ past human social behaviour “where the conventional concepts and categories which underwrite the interpretation of human action defuse potentially disturbing archaeological data” (Murray 1992:731). Further, this practice fundamentally infringes postprocessual archaeology’s proposition that there is “no direct, universal cross-cultural relationship between behaviour and material culture” (Hodder 1991a:14). As Hodder observed back in 1982, the pitfalls of the current practices were known by the postprocessual proponents, however they have yet to abandon the short-term interpretive approach: If we interpret the past by analogy to the [ethnographic] present, we can never find out about forms of society and culture which do not exist today…what would be the point of repeating our knowledge of contemporary [ethnographic] societies by tagging labels on to societies of the past? (Hodder 1982a:14) Thus, the use of ethnographic evidence by postprocessualists places them in an epistemological quandary. Paradoxically, the assumption is that all humans have behaved in a similar manner to that which has been historically documented or ethnographically observed. However, the majority of these resources were collected and documented by Western scholars, using particular conventions of theory and classification which are familiar to them. Thus the very reason for valuing what is historically documented or ethnographically observed as an alternative to our modern reality must be questioned (see Hammersley 1992). As Martyn Hammersley notes: If it is true that what ethnographers produce is simply one version of the world, true (at best) only in its own terms, what value can it have? And there is no reason to suppose that ethnographers produce just one version of the world.

breadth and complexity of human social behaviour that has materialised in archaeological assemblages over the last 50-100000 years of modern human existence, let alone two to three million years of hominid existence. The very sporadic, fragmented and selective way in which these observations were/are collected and documented should automatically deem them problematic for the interpretation of human social behaviour identified in prehistoric archaeology. Further arguments that analogies are used as a starting point for the interpretation of human social behaviour (for example, Tilley 1996a:16; Whittle 1998:853) are all well and good, but the analogies employed ultimately impede what we can observe. One only has to look at the final interpretations, for example by Tilley (see Tilley 1996a:334-5 and pages 53-55), to see the intrinsic and dominant role played by analogies. This current method automatically pigeonholes archaeological assemblages, such as human skeletal material (King 2001), as the archaeological manifestation of a very specific form of human social life. The situation is compounded and made even more difficult because the current ‘Neolithic’ discourse is one in which the ‘radical’ proponents have emphasised the search for ‘radically’ different pasts. These ‘radicals’ wished to escape the discourse which prevailed from the 1950s to the 1980s which purportedly ‘reduced’ the ‘Neolithic’ to economic rationalism (Hodder 1990a:31; Thomas 1991a:77, also see Bender 1992). However, their ‘radical’ approaches have not subsequently been able to break free from the interpretive framework which sees their continual employment of standard social theory to truly seek to identify a unique and different ‘Neolithic’. Their failure to engage in a fundamental questioning of the habitual theoretical and methodological tools they employ, and specifically the failure to redefine the ‘social’ into a form appropriate for analysing the archaeological assemblages and contexts, has resulted in the continual production of the Past-as-Same and Past-as-Analogue (see Ricoeur 1984) despite claims to the contrary by Thomas (also see 1990a): If we accept the point that to write at all is to tame and homogenise the past, such a writing effects some kind of reconciliation between Same and Other. If we begin from the position that what we are striving to do is to free the past from ethnocentric and presentist deformations, our writing can begin with a radical separation of past and present through the use of genealogy and deconstruction, yet must end with an act of domestication. (1999a:5-6, emphasis mine)

Bradley (1998a:Chapter 1), in a recent work, begins by discussing two Trichterrandbecher (TRB) long structures at Barkaer in Denmark (Figure I.1), which neatly illustrates the interpretative problems in the ‘Neolithic’ discourse. Peter Vilhelm Glob worked at Barkaer from 1947 to 1975, but died prior to publishing his findings. Preliminary interpretations by Glob identified two of the longest houses in prehistoric Europe, each some 90 metres long (Figure I.2). Bradley discusses Barkaer in detail as he sees it as epitomising a number of problems in the archaeology of western and northern Europe, particularly in our preconceptions about what the ‘Neolithic’ should be (see Chapter 1). Bradley (1998a:9) states: Glob’s confusions were the confusions of an entire generation, faced with a Neolithic that failed to measure up to what was expected of it...Instead of evidence for stable mixed farming, there are signs of a more mobile pattern of settlement (Whittle 1996a: ch. 7). Instead of the houses of the living, they find monuments of the dead.

The problem is that the result is an ‘over-domestication’ of the evidence, which ultimately impedes the search and documentation of unparalleled forms of human social behaviour. The anomalous data present in archaeological assemblages and contexts are still reduced to classes of human sociality as attested by the familiar. Thus familiarity is chosen in preference to the ‘intransigence’ of past prehistoric human social behaviour. The current 3

Figure I.1 – The local setting of the structures at Barkaer (Jutland) (after Bradley 1998a: figure 2).

4

Figure I.2 – The ‘Neolithic’ long ‘houses’ at Barkaer (Jutland) (after Bradley 1998a: figure 1).

5

While in agreement with Bradley’s overall assessment, I also see the re-interpretation of Barkaer as continuing to epitomise the past and current interpretative conundrum of the ‘Neolithic’ European landscape – the prevailing assumption that there must be distinct locations for the living and the dead, despite evidence to the contrary.2

Conclusions No matter how committed a researcher may be to identifying ‘difference’ in archaeological assemblages, the use of standard social theory does not allow this. Maintaining the use of standard social theory, drawing on ethnographic and ethno-historic analogies, drastically reduces the scope of human social behaviour which can possibly be identified, to the point where the archaeological past is liable to be a pale imitation of the findings, declaration and categories of the social sciences and history. We must recognise the uniqueness of archaeological assemblages and contexts and our unique position to identify unparalleled forms of human social behaviour. This position is partly due to the fact that we can deal with human social behaviour on a time-scale which is not encountered in the disciplines of other cultural studies. We must begin to develop our own methodology and interpretative tools, specifically sensitive to the discipline. We should seek to be in a position to identify and interpret the operations of the actual human social behaviour represented in archaeological assemblages and contexts, and avoid the current pitfalls of interpreting prehistoric human social behaviour and verbalised meaning and move toward identifying possible new elements of human social behaviour.

Taking Barkaer as an example, Bradley states that Glob’s preoccupation with interpreting the contexts as permanent settlements (‘houses’), characterised as an integral attribute of the ‘Neolithic package’, subsequently led him to marginalise the burial evidence. However, by Bradley’s own admission, the burial evidence did not include human skeletal material, as none was excavated from the context, though this absence is put down to loss due to acid soil. Instead he stated that a stone cist “contained two pots and an amber bead: a collection of artefacts which might well have served as grave goods” (1998a:5). Is this location an occupation context with a later burial? Is the cist even a burial? Perhaps the problem is in the assumption that all human social behaviour, irrespective of time and space, should be or can be easily reducible and predictably placed into a series of basic Western/modern functional categories, such as ‘domestic’ and ‘ceremonial’/‘ritual’ (see Fowler 2000:118-9; Pollard 1998, 2000a and Chapter 6). It is extraordinary that we still expect or consider that all humans will necessarily organise their behaviour in such a way that makes it familiar to us. As M.A. Smith (1955:4, original emphasis) commented back in the 1950s: we must not make the mistake of assuming that they [prehistoric populations] could have been aiming only at things which seem justifiable to us. The Western standards of value which condition our own thought are not absolute. The potential existence of unique past practices may be lost in favour of one which is at once familiar and inaccurately portrayed. Some elementary actions are required to control this problem. For instance, throughout the book I will avoid the value laden terms ‘tomb’ and ‘barrow’. The reason for this is twofold. First, the use of such terms suggests that these locations are the definitive mortuary location in the landscape, which is not necessarily the empirical reality. Secondly, their use imposes our modern notions of the dichotomies of social life on the past.

What follows will further document the current pitfalls of the use of such data and interpretive approaches in archaeology which far from helping to formulate the ‘otherness’ of the past, result in the loss of unparalleled forms of human social behaviour in favour of interpreting a familiar and recognisable past. While a number of ‘Neolithic’ scholars, such as Pollard, Bradley, Kirk, Whittle and Thomas, have offered new and significant viewpoints on various aspects of the ‘Neolithic’, including mobility and the patterning of activity across the landscape, which will be referred to throughout, they have not produced a coherent rethink of the ‘Neolithic’ within its prehistoric context. I will argue that we should rigorously view the past as a foreign country and past social human behaviour as offering a unique perspective, which only archaeology, with its long-term and deep time depths, has the opportunity to identify and which may have no direct parallel in either observed and/or documented human social behaviour.

2. There are similar re-interpretations for other structures. For example Francis Pryor’s (1991:51, 1993:138) discussion concerning the changing interpretation of Fengate Site 11 (Cambridgeshire) from a ‘house’ to a mortuary enclosure and Padholm Road, Fengate (Cambridgeshire) being interpreted as a ‘house’ but then later interpreted as a ritual structure.

6

CHAPTER ONE: DEFINING THE ‘NEOLITHIC’ microliths. However, microliths have been found in late ‘Palaeolithic’ and even ‘Neolithic’ contexts and have not been found in all ‘Mesolithic’ industries across Europe (Dennell 1983:130; Price 1987:278-9; Telford 2002:Table 1). Other stone tools in ‘Mesolithic’ contexts, such as blades, end-scrapers and burins industries also show variability across Europe. Conventionally, no further clarification other than ‘hunter-gatherer’ has been used to describe ‘Mesolithic’ subsistence, apparently due to the general low level of preservation of organic material across Europe (Mellars 1981:13-14). While there are some spectacular examples of organic preservation, such as Star Carr and Oronsay (Legge and RowleyConwy 1988; Mellars 1987; Mellars and Dark 1998; Morrison 1980), this cannot generally be extrapolated to other contexts in Britain, Ireland or continental Europe.

Like old soldiers, archaeological periods do not die but simply fade away (in most cases, all too slowly). Today, the archaeological periods constructed in the nineteenth century can still be discerned behind the facade of twentieth century renovations. (Dennell 1983:1) As an archaeological concept the ‘Neolithic’ has undergone significant change since its development in the middle of the nineteenth century (see Czerniak 1998; Pluciennik 1998; Thomas 1993b, 1997a; Zvelebil 1998a). This chapter will outline the history of what the term ‘Neolithic’ has been associated with in order to differentiate it from the ‘Mesolithic’. The characterisations have ranged from the technological to the economic to the conventions of social relations. Despite the fluid nature of its associated attributes, the ‘Neolithic’ has generally been seen as a unified, homogeneous entity since its identification by Sir John Lubbock as a technological phenomenon (Thomas 1993b, 1998c). Though claims are made to the contrary, such a perception continues in the current discourse as the ‘Neolithic’ concept undergoes another metamorphosis into an epoch of phenomenological meaning. This reappraisal has sought to define the ‘Neolithic’ in terms of criteria based on a definitive way of viewing the landscape and treating dead that differentiates it from the ‘Mesolithic’.

Models for ‘Mesolithic’ societies have drawn heavily from ethnographic stereotypes of hunter-gatherers, in some cases from a very simplistic understanding based on a belief in universal patterns of behaviour (Spikins 2000b). ‘Mesolithic’ populations have generally been viewed as just a prelude to later ‘Neolithic’ populations. For example, Grahame Clark (1980:5) described the ‘Mesolithic’ as having: crucial significance for understanding the course of prehistory, and not least for explaining the rise and spread of the Neolithic societies that laid the foundations of the diverse civilisation of mankind. However, the 1980s and 1990s have witnessed an ‘upgrading’ of ‘Mesolithic’ populations mainly through ethnographic analysis of complex hunter-gather populations such as the Ainu and Kwakiutl. ‘Mesolithic’ populations are no longer viewed as passive victims, prisoners of their natural environment (Gijn and Zvelebil 1997:3). Evidence of ‘Mesolithic’ cemeteries, elaborate archaeological assemblages and social differentiation has been used to rank the European ‘Mesolithic’ populations as complex hunter-gatherers (Clark and Neeley 1987; Larsson 1989a, 1989b, 1990a, 1990b; Neeley and Clark 1990; Price and Brown 1985; Rowley-Conwy 1983; Schulting 1996; Zvelebil 1986). But drawing analogy between these and complex hunter-gatherers from the present and recent past is problematic. Armit and Finlayson (1992:667) indicate, “many contemporary hunter-gatherers have economies inextricably entangled with neighbouring farming economies”, making the drawing of analogies from such societies problematic and misleading (also see Wobst 1978). However, this is not always the case, as groups such as the Northwest-Coast of America and the coastal Californian hunter-gatherer populations were not in such situations.

Changing perspectives on the ‘Mesolithic’ and ‘Neolithic’ The Irish archaeologist Hodder M. Westropp was the first to use the term ‘Mesolithic’ in an 1866 lecture to separate the ‘Palaeolithic’ and ‘Neolithic’ and later the term appears in his 1872 book, Prehistoric phases; or, introductory essays on pre-historic archæology (Gräslund 1987:38; Rowley-Conwy 1996).3 As Rowley-Conwy (1986:13) has remarked, the ‘Mesolithic’ is the ‘period’ between the ‘cave painters and crop planters’. Traditionally, the ‘Mesolithic’ has been differentiated technologically from prior and later periods. ‘Mesolithic’ stone tool assemblages are commonly characterised by 3. While first published in 1872, the term was not generally accepted until fifty years later (Zvelebil 1986:5).

The English archaeologist Sir John Lubbock originally defined the term ‘Neolithic’ in his 1865 book, 7

Pre-historic Times: as illustrated by ancient remains and the manners and customs of modern savages, describing it as: The later or polished Stone Age: a period characterised by beautiful weapons and instruments made of flint and other kinds of stone; in which, however, we find no trace of the knowledge of any metals, excepting gold, which seems to have been sometimes used for ornaments. This we may call the ‘neolithic’ era. (Lubbock 1872:2) In the same work, Lubbock also coined the term ‘Palaeolithic’. These classifications, along with the definition of the ‘Mesolithic’, were part of a broader endeavour to segment prehistory on purportedly Darwinian and geological lines (Daniel 1975:85-9; Thomas 1993b:362; Trigger 1989:114-6). The assumption was that ‘stages’ were a key attribute of Darwin’s theoretical structure, and thus should be duplicated in the Humanities. This work built on the Three-Age System proposed by the Danish scholar Christian Jürgensen Thomsen in 1819 and first published in an 1836 museum guidebook (van Riper 1993:41).4 However, these terms were not just related to archaeological chronological periods, they were also subsequently defined as stages of cultural progress, as part of a “belief in the inevitability of progress” (Czerniak 1998:30).

The ‘Neolithic’ as a technological phenomenon As discussed above, Lubbock’s initial identification of the ‘Neolithic’ was based on the presence of ground and polished stone tools, in much the same way that microliths have been seen to characterise the ‘Mesolithic’. At the end of the nineteenth and the beginning of the twentieth century, pottery was also identified as a key technological and artefactual feature of the ‘Neolithic’ (Kinnes 1988:4). As Thomas (1993b:363) indicates, Lubbock himself stated that pottery could be found in ‘Neolithic’ lake dwellings and kitchen middens. Nevertheless, he marginalised the importance of pottery as a technological attribute. Childe’s concern with diffusionism resulted in the view that ground and polished stone tools, pottery and agriculture were a constantly occurring ‘package’, which archaeologically identified and defined his ‘Neolithic Revolution’ (Childe 1936, 1957:341-3). Currently, the archaeological identification of ground and polished stone tools and pottery in non/pre-‘Neolithic’ contexts (defined as such by other attributes and absolute dates) has diminished the value of these attributes as a way of defining the ‘Neolithic’. For example, polished stone tools have been identified in Ireland and Scandinavia among ‘Mesolithic’ hunter-gatherer groups (Woodman 1978:114), while pottery-using ‘Mesolithic’ hunter-gatherers have also been identified throughout Europe, including recent finds from Lepenski Vir (Yugoslavia) (Garašanin and Radovanoviü 2001; Zvelebil 1986:Figure 4; Zvelebil and Lillie 2000). In Britain, ceramics have been identified within ‘Mesolithic’ contexts at Kinloch (Highland) and Ulva Cave (Argyll and Bute). However, at these two contexts it is difficult to distinguish between the ‘two periods’, suggesting they were continually utilised into the ‘Neolithic’ (Armit and Finlayson 1992:668-9; Richmond 1999:39). Such a conundrum required those scholars who feel the need to retain the ‘Neolithic’ as a category must construct a new way of defining the ‘Neolithic’. Their quandary in attempting to differentiate the ‘Neolithic’ will be developed throughout this book.

John Lubbock’s development and definition of the ‘Neolithic’ specified it as a technological phenomenon – the presence of ground and polished stone tools were identified as the diagnostic attribute of the period. However, as the term ‘Neolithic’ began to be habitually applied and utilised by practically all European archaeologists, its diagnostic characteristics have been changed over the following century and a half. New scientific perspectives and the accumulation of empirical evidence have resulted in a wide spectrum of archaeological attributes being applied to the term ‘Neolithic’, ranging from technological to economic and the social and now phenomenological meaning. The concept of the ‘Neolithic’ has come to be associated with a whole series of technological, socio-cultural, sociopolitical, economic, subsistence and philosophical differences, supposedly distinctive to the period.

The ‘Neolithic’ as an economic phenomenon The identification of the ‘Neolithic’ as an economic phenomenon remains the dominant way of differentiating the period. The economic change associated with the ‘Neolithic’ is a shift from the exploitation of wild resources to domesticated plants and/or animals. Interestingly, Lubbock (1872:210) did not identify the domestication of plants and animals as occurring in the ‘Neolithic’, instead viewing this as a later development associated with the introduction of bronze. It was Westropp (1872) who linked the identified technological sequence, based on polished stone tools, “with an assumed economic one: thus the ‘Mesolithic’ became equivalent to hunting, the ‘Neolithic’ to pastoralism, and the Bronze Age to agriculture” (Dennell 1985:115). This general economic view was reinforced during the 1920s and 1930s by Childe’s famous ‘Neolithic revolution’ statement which identified agriculture, not just

4. However, Vedel-Simonsen’s 1813 publication, Aperçu sur les périodes les plus anciennes et les plus remarquables de I’histoire nationale, “had already argued for three periods of Scandinavian antiquity – a Stone, a Copper or Bronze, and an Iron Age” (Lowie 1937:21).

8

Neolithic of Britain might have had their origins in processes and tensions already present in the Mesolithic” (Thomas 1991b:19). However, many reconstructions of ‘Neolithic’ populations “rely on the ethno-history of European peasantry for direct historical analogies” (Zvelebil 1998a:12-3, also see Armit and Finlayson 1992:667, for examples see G. Barker 1985; Clark 1952; Higgs 1972, 1975; Halstead 1981; Zvelebil 1981). Such practices reinforce the perception that the ancestors of modern Europeans were farmers, identifying ‘Neolithic’ society as the essential root of modern society and which can thus be reconstructed in their (modern Europeans) own image (see Gijn and Zvelebil 1997; Kristiansen 1996; Rudebeck 1996, 2000; Zvelebil 1996a) (Figure 1.1). This ‘Neolithic’ past is seen as resembling the European traditional social world of houses, farms, cemeteries and so on, an argument which will be documented throughout this book.

pastoralism, in the ‘Neolithic’ (Childe 1936, 1957:341-3). Thomas (1993b:363) states that while the initial identification was based on technological change, a change in economy is now seen as a major (if not the crucial) identifying attribute of the ‘Neolithic’. Many factors and models have been proposed to examine the origins and spread of agriculture. The ‘wave of advance’, indigenous and availability models for shifts to agriculture will be elaborated on in Chapter 2, but as Dennell (1992:92) noted: the concept of the ‘Neolithic’ as signifying the appearance of agriculture probably has done more to obscure than to illuminate the nature of the processes involved. External factors, such as climate change, resource degradation or population pressure, have been cited as factors beyond the control of human populations that may have forced subsistence change (Baillie 1992; Binford 1968; Bar-Yosef and Belfer-Cohen 1992; Childe 1928; Cohen 1977; Jarman et al. 1982; McCorriston and Hole 1991; Rowley-Conwy 1984; Wright 1977). In the case of Britain and Ireland, Baillie’s (1992, 1995) recent work on tree-ring chronologies identified difficulties in constructing a chronology for the early sixth millennium BP, which Baillie (1995:147) has suggested might be due to some sort of environmental pressure, forcing a rapid shift to domesticated plant and animal resources. In western and north Europe, the elm decline, a sharp reduction in elm pollen, has been associated with climate change that coincided with the introduction of agriculture. However, this decline in elm pollen is now understood to be a result of disease, rather than climate, and only sometimes occurs after the appearance of agriculture (Peglar and Birks 1993).

The ‘Neolithic’ as defined by megaliths and earthen ‘barrows’ Prior to the Lubbock divisions of the Stone Age, the Danish archaeologist Jens Jacob Asmussen Worsaae (1849) had suggested that megalithic ‘tombs’ and associated earthen ‘barrows’, particularly those in Scandinavia, characterised the later part of the Stone Age. Despite this, early definitions generally identified megalithic structures and earthen ‘barrows’ as secondary to the other fundamental traits of the ‘Neolithic’ (Thomas 1993b:363-4). For instance, Childe, while identifying pottery, houses, polished stone tools and agriculture as characteristics of the ‘Neolithic’, was reluctant to view megalithic ‘tombs’ and earthen ‘barrows’ as a necessary part of the ‘Neolithic package’. Childe, as a diffusionist, perceived the ‘Neolithic’ as a pan-European phenomenon. As megalithic and earthen constructions were confined only to western and northern Europe, he could not identify them as articles of the ‘Neolithic package’. Rather, Childe identified those ‘monuments’ as additional attributes introduced by “apostles of the megalithic faith” to an already ‘Neolithic’ society (Childe 1957:365, also see Thomas 1993b:366).

The ‘Neolithic’ as a social phenomenon The subsistence change associated with the ‘Neolithic’ has been seen as having a crucial impact on the social and political aspects of life in the period as well (Bender 1978, 1989; Hayden 1990, 1992, 1993:192-265). The link between subsistence and social practices has been an underlying assumption that has persisted in archaeology and anthropology for the past 250 years (Pluciennik 2001). Until recently, the social lives of ‘Neolithic’ farming populations have been regarded as more ‘advanced’ than those of ‘Mesolithic’ hunter-gatherers. Zvelebil (1995a:79) argues, however, that such perceptions of ‘Mesolithic’ hunter-gatherers were derived, in the first half of the twentieth century, from assumptions about ethnographic analogies that generally (though incorrectly) described hunter-gatherer groups as having simple social and economic organisations (Pluciennik 2001).

Linked with the view that the ‘Neolithic’ was synonymous with the domestication of plants and/or animals, it has been argued that this development led to Western concepts of territory and property, of which Renfrew used to argue that ‘tombs’ and ‘barrows’ were territorial markers (Renfrew 1973, 1976; also see Armit and Finlayson 1992:667). Recently, however, megalithic ‘tombs’ and earthen ‘barrows’ have become fundamental to defining the ‘Neolithic’ in terms of phenomenological meaning (for example, Cooney 2000a, 2000b; Edmonds 1999a; Fowler and Cummings 2003; Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Tilley 1994, 1995a, 1996b; Tilley and Bennett 2001; Thomas 1991a, 1996b, 1999a, 1999b; Watson 2001). According to Thomas, it was the construction of these ‘monuments’ that heralded the beginning of the British ‘Neolithic’ at the middle of the seventh millennium BP. He links these new ‘monumental’ constructions with a major cultural and social change – a change that Thomas

Recent work has elevated the social practice and standing of ‘Mesolithic’ populations (Bradley 1998a; Clark and Neeley 1987; Ingold et al. 1988a, 1988b; Jeunesse 1996; Larsson 1989a, 1989b, 1990a, 1990b; Neeley and Clark 1990; Price and Brown 1985; Schulting 1996). Further, the possibility has been entertained among scholars that “some of the social changes which took place in the 9

Figure 1.1 – The Homeland, by J. Mánes, 1856. The national self-perception created by Czech and Slovak intellectuals referred to ancestors as peaceful farmers, an image captured in this nationalist painting (after Zvelebil 1996a: figure 10.1). Despite such sentiments, Thomas, along with the vast majority of British and Irish archaeologists, has continued to utilise ‘Neolithic’ (and ‘Mesolithic’) as a convenient shorthand despite widespread agreement of the problem of the use of chronological labels. Despite recent reappraisal of the use of the term ‘Neolithic’, the discourse continues in a similar vein, identifying it as a self-evident cultural phenomenon which can be archaeologically identified in assemblages and contexts. While I use the terms I do not imply any economic or social behaviour. Rather I use them to refer to chronological periods that have come to be associated with them.

associates with the new relationship of the populace with their landscape and their ancestors (1988b, 1991a, 1993b:388-9, 1999a). This postprocessual approach was a reaction against the processual ‘economic’ view that identified the ‘Neolithic’ as a socio-economic phenomenon. Both will be critiqued throughout this book.

The ‘Neolithic’ Discourse The identification of elements of human social behaviour with conventional chronological schemes has resulted in archaeological attempts to assign a set of cultural traits to a specific segment of time/type of assemblages and contexts. In consequence, repeated redefinitions have led historically to the term ‘Neolithic’ being swamped by ambiguity. Its definition has not been consistently associated with a single phenomenon and has gained increasingly generalised, vague social/cultural entanglements. The term has come to mean very different things but, as indicated above, it “has always drawn upon a more or less inexplicit conception of totality” (Thomas 1993b:383). Recently there has been a reappraisal of the view of the ‘Neolithic’ as a unified and homogeneous entity (for example, Pluccienik 1998:79; Price 1987:2923; Thomas 1993b; Zvelebil 1993a:64). Thomas has been one of the most vocal opponents of the term ‘Neolithic’. Referring to his 1991a work studying ‘Neolithic’ southern Britain, he comments: The character of these changes [associated with the ‘Neolithic’] has been obscured by the use of the term ‘Neolithic’, which has conditioned the way in which the available evidence has been read. (Thomas 1993b:389)

The questioning of the term ‘Neolithic’ and its recent redefinition in phenomenological terms, is part of a wider questioning of how to ‘do’ archaeology. The general ‘rethink’ has taken two major directions. The first, which will be discussed below, examines the cultural attributes associated with the ‘Neolithic’. These cultural views have seen the ‘Neolithic’ as differentiated from the ‘Mesolithic’ due to a newly developing phenomenological awareness by the ‘Neolithic’ population. The second, discussed in Chapter 2, has investigated the economic and subsistence strategies identified with the ‘Neolithic’. While these two approaches have overlapped in a number of ways, they will be examined separately. They have followed rather different logical paths, the former being argued under the postprocessual banner while the latter lies generally under the processual banner. However, they have both attempted to identify the ‘Neolithic’ as a culturally and archaeologically distinctive entity.

10

a wholesale transformation of social relations which results from adopting an integrated cultural system. Such a system has as its purpose not merely the provision of subsistence, the biological reproduction of the community, but its social reproduction, including the maintenance of power relations, knowledge and institutions. Owning a cow, or an axe, living in a house, or burying one of one’s kin in a particular way does not make a person Neolithic. It is the recognition of the symbolic potential of these elements to express a fundamental division of the universe into the wild and the tame which creates the Neolithic world. Thomas sees the symbolic recognition of the interlocking binary opposition, of culture:nature, human:animal, life:death, etc., as fundamental to the ‘Neolithic’ social order (see Whittle 1996a:360). The megalithic ‘tombs’, earthen ‘barrows’ and all that they are assumed to encompass (ritual, communal feasting, ancestor worship, etc.) define the ‘Neolithic’ and played a major role, “imposing a certain conceptual scheme upon the world” (1991a:182). Thomas presents the ‘Neolithic’ as a phenomenological and cultural phenomenon – “a system of social reproduction” (ibid.). While his work, and the work of others with a phenomenological perspective, is presented as radical, and has subsequently been viewed as radical, Ashbee (1991:268) indicates that such concerns are far from new. In his view, such cultural stances have been part of the approach to the ‘Neolithic’ as far back as Childe in the 1920s. Notwithstanding Thomas’ aims to reject the ‘Neolithic’ as an economic phenomenon, Richards states that he is unable to totally “dispense with generally muddled and discredited criteria”, as he continues to employ the term ‘Neolithic’ (1993:807).

The Phenomenological ‘Neolithic’ Farming is not only an ‘economy’ or a ‘mode of production’, but also a cultural and symbolic construct. (Olsen 1988:431) Phenomenology was first defined by the German philosopher Edmund Husserl and relates to the manner in which people experience and understand the world. In British and Irish archaeology it has been tied up with a concern for investigated landscapes, ancestors and movement around/between monuments and has built on the view that the ‘Neolithic’ witnessed an ideological (religious) change in society (see Cauvin 1978, 1989, 1994; Hodder 1990a). The phenomenological ‘Neolithic’ has been promoted by an influential group of radical British contextualists including Thomas (1991a, 1996b, 1999a, 1999b, 2001a), Barrett (1994, 1996), Tilley (1994, 1995a, 1996b; also see Tilley and Bennett 2001), Edmonds (1999a, 1999b), Parker Pearson (2000; also see Parker Pearson and Ramilisonina 1998a, 1998b) and Whittle (1996a). In addition, it has recently been attempted by the Irish archaeologist Cooney (1994, 1998, 2000a, 2000b). Tilley (1996a:73) sees the ‘Neolithic’ as the start of a new ideology which “restructured notions of time and space, death and the body, prestige and social competition”, resulting in social inequality. Proponents of ideological and phenomenological approaches generally argue that: the Neolithic was primarily not an economic phenomenon, nor a new set of social relations, nor the manifestation of an immigrant group, but the material manifestation of a new set of ideas restructuring Late Mesolithic societies and changing their social and economic condition of existence. (ibid.:72) As outlined in the introduction, the weakness of this approach is that interpretations are based on a very selective and particular way of looking at the empirical evidence. Analogies are chosen which are often highly selective. These use general social categories which merge verbal meaning, action and material culture into one entity. However, with such an approach ‘verbal meaning’ is the decisive component, with the current discourse professing to ‘understand’ past human behaviour through familiar standard social categories

Barrett views the monuments and the associated archaeological assemblages in relation to the prehistoric individual’s emotions and understanding (1994:3). Whittle sees constructions, including long mounds, chambered structures, enclosures and long ‘houses’, as helping to frame the landscape, thereby acting as a focus for routines and rites (1996a:369). For Cooney (2000a:90-1), these monuments are the physical presence of the ancestor in the landscape, as he states: Landscape is linked to the ancestors and across and within the landscape the dead people walked…the saturation of the landscape with the spiritual presence of the dead and the ancestor. Tilley, in his influential A Phenomenology of Landscape: Places, Paths and Monuments (1994), provides insights into the ‘sacred geography’ of the British ‘Neolithic’ (see Fleming 1999). Drawing on a few ethnographic accounts, mainly from the Australian Aborigines, he argues that the landscape is the arena where “Histories, discourses and ideologies are created and re-created through reference to the special affinity people have with an area of land, its topography, waters, rocks, locales, paths and boundaries” (1994:67).

Scholars arguing for a phenomenological ‘Neolithic’ have strongly emphasised continuity of subsistence, economic and residential behaviour between the ‘Mesolithic’ and ‘Neolithic’ (see Chapter 2 and 6). Thomas (1991a, 1999a) and Edmonds (1999a) have proposed that the ‘Neolithic’ was characterised by mobile communities involved in mixed farming-hunting. Residential and economic continuity from the ‘Mesolithic’ brought about the need to redefine the ‘Neolithic’ as different in phenomenological terms in order to maintain it as a culturally distinct period. Thomas (1991a:13) defines the ‘Neolithic’ as:

Edmonds’ book, entitled Ancestral Geographies of the Neolithic: Landscape, monuments and memory (1999a), can be seen as epitomising the current ‘radical’ trend in 11

developed Rowlands’ critique of the modernist and Eurocentric fantasies of later prehistory (Rowlands 1986, 1987). Such fantasies have lead to our capitalist modes of production being projected into the past. Thus, “our modernist, historically specific values and common sense notions apply” to the later prehistory of Britain (Hill 1989:17). Hill has attempted to see the past not just in relation to present concerns.

British and Irish ‘Neolithic’ archaeology. The book is experimental, with each chapter starting with a poetic and nostalgic account of life in a ‘Neolithic’ landscape. It is recounted in terms of a familiar rural setting, and as Hodder (2000:378) states in his review, the “past is familiar and the rhetoric supports the familiarity”. The fictitious stories introducing each chapter, coupled with the complete lack of references in the text and lack of evidence to support interpretations produces a book where the line between what is fiction and what is fact is lost. Unfortunately, only experienced readers of ‘Neolithic’ studies can discern such a line between the rhetoric and that which is empirically testable. These measures help to achieve a text that is impenetrable to all but experienced readers on the subject, despite Edmonds’ attempt to make the text “less determined and more open” (Edmonds 1999a:x).

Despite the theoretical claims to be writing a different past in recent ‘Neolithic’ studies, there is predictability to individual’s relationships with place and their interaction with material culture: It is almost as if one sees in such writing not the self-critical, reflective social scientist, but the lord of the manor ‘taking a walk’ around his estate, surveying ‘his’ landscape, entirely comfortable within a familiar land and nation to which he ‘belongs’. (Hodder 2000:377) The familiarity is produced through the use of ethnographic and ethno-historical analogies which give the writer a false sense that they are producing a ‘different’ past. There is a sense of arrogance in the writing. The freedom expressed in the writing, the literal speaking for the ‘Neolithic’ populations as a commonsense experience of the world which one can easily emulate, stems from a belief that we (the ‘author’/‘writer’) and the ‘Neolithic’ population are one (ibid.:379, also see Cooney 1994, 1998, 2000a, 2000b; Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Tilley 1994, 1995a, 1996b; Tilley and Bennett 2001).6 As Whitley (2002:120, however see Pitts 2003) has recently observed: British prehistory is becoming more particular, more nationalistic and more inclined to stress the uniqueness of the British sequence. The more unique the period appears (and nothing could be more unique than the British Neolithic), the more British it becomes.

Hodder’s review of Edmonds’ book also evaluates the current phenomenological trend of the ‘Neolithic’ discourse in Britain. Edmonds’ book focuses on the ‘Neolithic’ populations of Britain and Hodder uneasily detects a familiarity in the writing. The ‘Neolithic’ bodies described moving within a henge or over a cursus or through woodland are at once familiar to the reader and the writer. Hodder describes them as “universal bodies”; allowing for their relationship with the landscape as unproblematic and thus taken for granted (also see Brück 1998, 2001). The landscape is viewed in the British, or at least English sense of nationalism and nostalgia. A timeless landscape is produced linked to the image of ‘Britishness’,5 as Hodder states, similar to that created in Turner’s paintings of Stonehenge (Figure 1.2) or the image of a Victorian picnic in a landscape of monuments (Barclay 2001; Evans 1985:84; Hodder 2000:377; Lowenthal 1994:20; Thomas 1999a:11). Evans (1985:84) argues that the current perception of the prehistoric and historic landscape is related to the present cultural landscape, in particular the visual impact of monuments. Thus, the current reality, of conspicuous monuments within a ‘tamed’ landscape, translates itself into the view that the monuments were just as conspicuous in the ‘Neolithic’ landscape. However, this view is not supported by the evidence which suggests that woodland still dominated the landscape from the middle of the seventh to the end of the fifth millennium BP (see Chapter 2).

An unproblematic link has been constructed between the writer, the land, the monuments and the artefacts, linked 6. Born and raised on the other side of the world in a former convict colony of the British Empire, and having a father who never romanticised his first twenty two years on a farm in County Galway, Ireland, I cannot claim to have feelings of nostalgia or romance for the British or Irish rural landscape. It would be inappropriate for a nonindigenous Australian archaeologist to portray the indigenous experience of living within Australian precolonial (pre-invasion) landscapes in the terms espoused for the British and Irish ‘Neolithic’ (the same can be said of American archaeologists in relation to native American populations, see Hodder 2000:379).

The work of Hill on the ‘Iron Age’ of southern England questioned the ‘unproblematic’ and ‘familiar’ view of ‘Iron Age’ (Hill 1989, 1993, 1995a, 1996). His work 5. Pittock (1999:104) states that Britishness “is another word for Englishness; it is a political word…which extends Englishness over the lives of the Welsh, the Scots and the Irish”.

12

Figure 1.2 – J.M.W. Turner, Stonehenge, c.1825-28 (after Chippindale 1994: figure 80). Zvelebil and Beneš (1997:23) have judged as flawed archaeologists’ attempts to examine “historical and prehistoric landscapes in terms of social relations, relations of power, identity, and appropriation, and as a reflection of our own, modern beliefs”. They argue that such examinations, entrenched in modern ideology, are inadequate to account for the character of prehistoric landscapes. More to the point, Johnston (1998a:314-16) has argued that current archaeological practice of employing landscape as a universal is fundamentally flawed. He points out that, while current accounts (see Tilley 1994) move away from the ‘culture:nature’ dichotomy of the structuralist perspective, they continue to impose “contemporary values when they use ‘landscape’ as a way of ‘living-the-world’ in the past” (also see Brück and Goodman 1999; Lemaire 1997). Johnston develops his argument to include Ingold’s (1995) recent paper discussing the notion of ‘dwelling’. Johnston contends that the current use of landscape in archaeology dichotomises “the built world (archaeology) and the natural world (the environment), ensuring there can only be an empirical and mechanistic relationship between the two domains of existence” (Johnston 1998a:316). Johnston’s work is part of an increasing awareness that the materiality of landscape is not as straightforward as it may seem” (van Dommelen 1999:277).

to nationalism, a sense of belonging and familiarity. In Britain this is based on the rural fantasy, which permeates much that is currently produced in ‘Neolithic’ studies. As discussed above, throughout Europe the perception is that farmers are the ancestors of modern Europeans (see Gijn and Zvelebil 1997; Kristiansen 1996; Rudebeck 1996, 2000; Zvelebil 1996a). Zvelebil (1996a) has argued that the long-standing paradigm which categorised ‘Neolithic’ farmers as our ancestors and ‘Mesolithic’ huntergatherers as savage ‘others’ was one of the key factors influencing a relative lack of interest in the ‘Mesolithic’ before the mid-1980s (also see Pluciennik 2001). So the current ‘Mesolithic’/‘Neolithic’ divide in Britain and Ireland is a direct product of a ‘romantic view’ of rural Britain and Ireland presented in art and literature, with ‘Neolithic’ populations the first in a long line of people that tamed and manicured the wildscape into the rural landscape which is still valued today.7 But as will be discussed, the landscape in the ‘Neolithic’ of Britain and Ireland was far from tamed (see Chapter 2).

7. The fight for the Green Belt in England has been a major issue for both English Heritage and the National Trust.

13

theoretical archaeology in both Europe and America over the past thirty-five years (Clarke 1968; Dunnell 1982; Hodder 1986, 1988, 1990b, 1992, 1995b; Raab and Goodyear 1984; Renfrew 1982; Shanks and Tilley 1987a, 1987b). Yet these authors have rarely questioned the kind of terminology used in (re)constructing the past. I will endeavour to ask questions specifically sensitive to the human social behaviour that we actually encounter in archaeological assemblages and contexts of Britain and Ireland from the early tenth to the late fifth millennium BP.

Conclusions …analogies, it is true, decide nothing, but they can make one feel more at home. (Freud 1974 [1933]:72) The appropriateness of nineteenth century chronological and cultural labels, such as ‘Mesolithic’ and ‘Neolithic’, has recently been called into question (Pluccienik 1998, Thomas 1993b; Zvelebil 1998). While the terms may serve as useful shorthand, their habitual use and varied associations diminish their effectiveness for the description of the British and Irish archaeological assemblages and contexts. The continual use of chronological/cultural divisions to define key issues of inquiry, and the use of archaeological descriptive terms based on verbal meaning and universal social theory, aim to create a comprehensible past (see Murray 1997). The use of chronological/cultural divisions, verbal meaning and universal social theory creates two sets of problems. The first is that a difference between periods then has to be explained, rather than demonstrated archaeologically. Secondly, the procedure creates the social conventions from which explanations will be constructed.

The development of cultural and economic historical archaeology led to a ‘Mesolithic’ equated with postglacial hunter-gatherer, mobile populations and a ‘Neolithic’ associated with farming, sedentary populations. While the new phenomenological approach has moved away from this strict subsistence and settlement distinction, a view that I share, they continue to maintain that the ‘two periods’ viewed their world differently. These arguments draw on ethnographic analogies in order to (re)construct social organisation to ‘Neolithic’ populations as a means to differentiate them from ‘Mesolithic’ populations. However, these ethnographic analogies have undercurrents that reflect the traditional view of the subsistence and settlement patterns of ‘Mesolithic’ and ‘Neolithic’ populations (Zvelebil 1993b:147). For all the rhetoric for continuity in subsistence and settlement patterns promoted by the new phenomenological approach, they are still arguing for discontinuity, based on evidence drawn from ethnographic evidence. Throughout I will continually demonstrate that the period between the early tenth and the late fifth millennium BP in Britain and Ireland does not neatly fit the conventionalised ethnographic generalisations. Thus: we should desist from forcing the archaeological record into conformity with ethnographic models, or into the straightjacket of our preconceived notions about Neolithic or Mesolithic society. (Zvelebil 1993b:159)

I argue that this practice only serves to insinuate difference while playing down continuity between the time span covered by the two terms. I will endeavour to illustrate that the labels, ‘Neolithic’ and ‘Mesolithic’, however useful and adequate in previous archaeological investigation, have now lost their usefulness in describing the detail and complexity of the British and Irish archaeological assemblages and contexts. Actively pursuing continuity between the ‘Mesolithic’ and the ‘Neolithic’ and avoiding the rhetoric of the current ‘Neolithic’ discourse, patterns of long-term human social behavioural traits present in archaeological assemblages and contexts can be identified (see King 2001). Coming to terms with chronological labels, as well as the problem of finding appropriate methods of interpreting archaeological assemblages and contexts, have plagued

14

CHAPTER TWO: ECONOMY AND SUBSISTENCE IN THE ‘MESOLITHIC’ AND ‘NEOLITHIC’ At the core of the ‘wave of advance’ model, as the name suggests, is the view that agriculture is a superior subsistence practice that would always be chosen over a hunter-gatherer existence. This, coupled with progressively later radiocarbon dates for agriculture as you move across Europe from southwest Asia, is seen as evidence for the gradual march (estimated by Ammerman and Cavalli-Sforza at about 1 kilometre a year) of domesticated plant and animals westwards across the European continent, via population growth and individual migration (Figure 2.1) (Clark 1965). However, this model has been criticised due to its simplistic view of the evidence related to dates, population growth, and migration. Furthermore, empirical evidence suggests that subsistence practice cannot be simply dichotomised either as hunter-gathering or agriculture, as subsistence practices are actually far more complex (Bateman et al. 1990; Dennell 1985; Fix 1996; Gkiasta et al. 2003; Kent 1989; Lahr et al. 2000; Meiklejohn 1985; Sims-Williams 1998; Thomas 1996c; Whittle 1996a; Zvelebil 1989, 1995b, 1996a; Zvelebil and Lillie 2000; Zvelebil and Zvelebil 1988). More specifically, it has been demonstrated that the ‘wave of advance’ model does not fit the spread and agricultural colonisation of the Bandkeramik and the Cardial cultures, with these cultures moving at rates of at least five kilometres a year (Barnett 2000; Price et al. 1995).

The important role of Mesolithic communities in the development of the record that is termed Neolithic, it would seem, has regularly been underestimated. (Richmond 1999:40, original emphasis) This chapter will outline the changing perspectives on the economic and subsistence strategies that have been associated with the ‘Neolithic’ over the last eighty years. In particular, it will focus on changes in the last thirty years, which have attempted to understand the processes associated with the domestication of plants and animals. These approaches have focused on the contribution and role of hunter-gatherer populations in the transition to agriculture. The implications of settlement patterns will also be discussed, with suggestions that continuity in clearance of woodland and subsistence from the ‘Mesolithic’ to ‘Neolithic’ may correspond with a similar mobile residential system of occupation between the ‘two periods’ (see Chapter 6).

Approaches to Economic Change in the ‘Neolithic’ ‘Wave of Advance’ Model: Agriculture, Genetics and Linguistics Economic and subsistence continuity from the ‘Mesolithic’ to the ‘Neolithic’ has been preached by two main schools, which will be referred to as the advocates of the ‘availability model’ and ‘indigenous model’. These two models both argue, to differing degrees, that economic and subsistence changes identified in the ‘Neolithic’ had their foundation in proceeding ‘periods’. Both schools differ greatly from long held traditional views about the spread of agriculture across Europe from southwest Asia expressed in the ‘wave of advance’ model (Figure 2.1). Proposed by the archaeologist Ammerman and the human geneticist Cavalli-Sforza, the ‘wave of advance’ model derived from the Childean diffusionist model conceptualised in the 1920s and 1930s (Childe 1936) and drew on a process noted earlier by Clark (1965). This model identifies agriculture as an imported phenomenon in western and northern Europe, with no significant contribution by indigenous hunter-gatherer populations. Childe (Childe 1936, 1957:341-3) envisaged agriculture, that is domesticated plants and animals, as an imported ‘Neolithic package’ from southwest Asia, brought by foreign colonists who, amongst others things, brought pottery, burial rites and new architecture (Price 2000a:3). However, as noted by a number of scholars, the ‘wave of advance’ does not necessarily require people to have advanced (see Ammerman and Cavalli-Sforza 1971, 1984; Zilhão 1993).

Advocates of the ‘wave of advance’ model have argued for both colonisation and adoption for the movement of ‘Neolithic’ traits across Europe. In addition to archaeological evidence, physical anthropology and genetic patterns have been employed to supply additional information on the relationship between ‘Mesolithic’ and ‘Neolithic’ populations and their connection with modern European populations. Zvelebil (1995b:38) argues that despite some claims (Diamond 1992:241; Hagelberg 1995:178; Renfrew 1992:463), the results have been far from conclusive. In the 1970s work edited by Schwidetsky (1973), the physical differences between ‘Mesolithic’ and ‘Neolithic’ populations were compared and it was found that there were relatively few differences (Price 2000b:303). Since the 1970s more ‘Mesolithic’ human skeletal material has been available for comparison with ‘Neolithic’ assemblages, with results confirming the earlier results of Schwidetsky (y’Edynak and Fleisch 1988; Formicola 1986; Jackes et al. 1997b; Jacobs 1994; Lillie 1996). It has been argued that the variations in skeletal morphology, which have been identified between ‘Mesolithic’ and ‘Neolithic’ individuals, “are the result of longer-term trends and do not reflect significant physical differences between the two populations” (Price

15

Figure 2.1 – Ammerman and Cavalli-Sforza’s ‘wave of advance’ model for the spread of agriculture into Europe. The radiocarbon dates are expressed in years bp (redrawn from Ammerman and Cavalli-Sforza 1971:685). The broken lines represent probable regional variations in the rates of spread (after Thomas 1996c: figure 17.1). and colleagues have examined the geographic patterning of modern human genetic traits across Europe (Ammerman and Cavalli-Sforza 1984; Bertranpetit and Cavalli-Sforza 1991; Cavalli-Sforza 1991, 1996; CavalliSforza et al. 1994). Their work claims to show a clinical distribution of frequencies that closely resembles the distribution of radiocarbon dates for the spread of ‘Neolithic’ across Europe. Cavalli-Sforza (1996:53) argues that this pattern is the result of population expansion, “the spread of agricultural settlers…into regions inhabited sparsely by hunter-gatherers, who initially had a different genetic background”. For the genetic gradient to form, demic diffusion and indigenous adoption of agriculture must have occurred across Europe.

2000b:303, emphasis mine, also see Meiklejohn and Zvelebil 1991). These include a general decrease in stature from the ‘Upper Palaeolithic’ to the ‘Neolithic’, with an increase in the later ‘Neolithic’ and a general reduction in tooth size from the ‘Mesolithic’ to the ‘Neolithic’ (Frayer 1984; Meiklejohn 1985; Meiklejohn et al. 1988). Further, the dramatic increase in dental caries amongst individuals in ‘Neolithic’ populations is generally recognised as resulting from a change of diet to a higher level of carbohydrates, rather than physical difference between the ‘two’ populations (y’Edynak 1978, 1989; Meiklejohn and Zvelebil 1991). Despite this evidence, it must be noted that there have only been a few studies that have involved the direct comparison between ‘Mesolithic’ and ‘Neolithic’ human populations within the same area. One example is the work of Jackes et al. (1997a, 1997b) in central and southern Portugal, with the evidence illustrating population continuity between the ‘Mesolithic’ and ‘Neolithic’. Finally, work by Harding (1990; Harding et al. 1990) illustrates no correlation between modern human cranial measurements and those measurements from ‘Neolithic’ populations, concluding that: There is no evidence from this analysis for the Neolithization of Europe by demic diffusion coupled to the spread of agriculture. (Harding et al. 1990:54)

Despite the generalised and uncomplicated conclusions of Cavalli-Sforza, there is much debate and disagreement on the correlation of maps of human genetic traits and the spread of agriculture. Price states, “there is in fact no evidence to link the two phenomena other than speculation” (2000b:304). Others have argued that drawing any correspondence between modern genetic patterns and prehistoric populations is questionable and complex due to time spans involved and the changing identities among ethnic groups (Pluciennik 1996). Renfrew (1996:85) argues that it cannot be shown that modern genetic frequencies, which represent recent human population movements, are the same as those in ‘Neolithic’ Europe. Further, Price (2000b:304) questions why genetic frequencies would have changed so

Human genetics is another source of information that has been utilised to identify the relationship between ‘Mesolithic’ and ‘Neolithic’ populations. Cavalli-Sforza 16

clock (Howell and Mackey 1997; Jazin et al. 1998).

dramatically across Europe. He argues that the colonists would presumably have carried the same genes across all of Europe. Cavalli-Sforza’s argument relies on the assumption that ‘Mesolithic’ populations were genetically homogenous across Europe. However, this is inconsistent with the evidence we have regarding the European ‘Mesolithic’ population. It also appears that not all genetic frequencies follow the pattern identified by Cavalli-Sforza. Work by Sokal and Barbujani with their colleagues (Barbujani et al. 1995; Sokal et al. 1991, 1992, 1996) has attempted to confirm the relationship between the gene frequencies and the initial ‘Neolithic’ radiocarbon dates, but found no conclusive evidence. Moreover, Fix (1996) has suggested that modern gene frequency clines in Europe, rather than the result of demic diffusion, are the result of natural selection. There has also been much dispute on the causes and directionality of the clines (gene frequency gradients), especially in relation to social and cultural phenomena (Brown and Pluciennik 2001; Clark 1998; Fix 1996; Renfrew 1998; Richards et al. 1996). However, recently the geographic distribution of Y-chromosome haplotypes from modern Europeans has been presented in support of the ‘Neolithic’ demic diffusion model (Chikhi et al. 2002), “suggesting that colonising farmers from southwest Asia contributed 70-90% of the genes in the population of each Neolithic settlements with an average contribution of 50% across the continent” (Bentley et al. 2003:63, also see Semino et al. 2000).

While questions of colonisation versus adoption have been asked in Britain and Ireland, the evidence, either biological or genetic, has not been forthcoming to resolve the question in this particular region (Woodman 2000, however see Evison 1999). There has been the publicised identification of the Cheddar Man skeleton excavated from Gough’s New Cave (Somerset) (see the Online Database and Appendix 6) and dated to the early ‘Mesolithic’,8 which still has descendants surviving in the area today (Barham et al. 1999)! Linked to the ‘wave of advance’ model has been the suggestion that the distinctive Indo-European languages spread with the first agriculture into Europe (de Laubenfels 1981; Krantz 1988; Renfrew 1987, 1989, 1992, 1994b, 1996, 2002). Renfrew’s initial argument was fundamentally linked to the ‘wave of advance’ model, as it argued that colonisation was the only mechanism through which the process could take place (1987). More recently this link between the two has become less definite, wit Renfrew (1994b, 1996) now suggesting that the distribution of language families and the spread of agriculture around the world are closely connected. Thus, Renfrew is packing one more trait into the ‘Neolithic package’, an addition that has been widely criticised (Gimbutas 1988; Hines 1997; Lasker and Crews 1996; Mallory 1989; Moore 1995; Otte 1995; Sherratt and Sherratt 1988; Sims-Williams 1998; Zvelebil 1995b, 1998c; Zvelebil and Zvelebil 1988). Most criticism centres on the lack of a demonstration by Renfrew of the association between the distribution of Indo-European languages and the early ‘Neolithic’, and the poorly understood relationships between genetics, artefacts and language. Renfrew, by loosening the direct association between the spread of language and the spread of agriculture, confuses the basis of his argument and effectively destroys it. Price (2000b:307) points out: If migration is not essential to the spread of language, then the spread of the Neolithic [agriculture] need have little to do with the issue. Language could have spread at any time before

However, recent work studying modern European mitochondrial DNA (mtDNA) has produced very different results than those that support demic diffusion (Richards et al. 1996; Sykes 1999; Wilkinson-Herbots et al. 1996). Analysis of modern human populations mtDNA suggests that more than 85 percent of European genes probably date to the ‘Upper Palaeolithic’, rather than the ‘Neolithic’. Recent work by Cavalli-Sforza and Minch (1997) has produced similar results. The mtDNA does imply some small-scale colonisation (10-15 percent of total mtDNA lineages) of Europe from southwest Asia during the ‘Neolithic’. Preliminary results indicate a “rapid penetration from the southeast, both westward and northwestward, followed by a much more gradual intermixing with the numerically dominant indigenous Mesolithic inhabitants” (Richards et al. 1997:253). Although mtDNA mutation rates are fairly well established and understood (Macaulay et al. 1997; Sykes 1999), questions have been raised concerning the mtDNA

8. The two calibrated radiocarbon dates from human skeletal material found in Gough’s New Cave (Somerset) are: 10550-10530 BP (.010), 10510-10110 BP (.892) and 10080-9920 BP (.099) (9100±100 bp, OxA-814); 1064010620 BP (.007) and 10580-9740 BP (.993) (9080±150 bp, BM-525).

17

Europe” (Zvelebil and Zvelebil 1988:578).9 Agriculture, according to this model, is not necessarily an introduced phenomenon; rather, agro-pastoral farming was added to existing subsistence strategies by indigenous peoples through interaction and exchange with agricultural populations (Figure 2.4). The process would have been slow and complex, all of which are allowed by the availability model, as the model is not strict and allows for a multitude of situations and outcomes, resulting from the varying lengths of the three phases in the transition to agriculture: availability, substitution and consolidation (Figure 2.4 and 2.5).

or after the introduction of farming. The ‘wave of advance’ model offers a very simple explanation for the movement and adoption of agriculture across Europe from southwest Asia. Despite the varied and detailed evidence offered it has been unable to convincingly link population movements in the ‘Neolithic’ and the spread of agriculture. It fails to acknowledge the very complex subsistence practices which populations utilise and the many paths and outcomes that different subsistence practices can take and lead to.

The availability model argues that hunter-gatherer populations were familiar with both the practice of agriculture and the material cultures associated with it prior to its adoption, witnessed through the movement of artefacts, including organic material, between ‘Mesolithic’ and ‘Neolithic’ populations (see Larsson 1988; Newell et al. 1990; Runnels and van Andel 1988). For example, from Scandinavia, “shoe-last” adzes of amphibolite from eastern Europe, along with ‘Danubian’ groundstone shaft-hole axes have been excavated from late ‘Mesolithic’ contexts. Pottery with domesticated cereal impressions, bone combs and t-shaped antler axes from ‘Neolithic’ populations have also been excavated from these contexts (Figure 2.6) (Andersen 1970, 1981; Fischer 1982; Price 2000c; Price and Gebauer 1992; Tilley 1996a:48-51; Vang Petersen 1984). In south central Europe ‘Neolithic’ artefacts have been found in late ‘Mesolithic’ contexts, including polished stone axes at Jägerhaushöhle and Falkensteinhöhle and grinding stones at Henauhof Nordwest (Jochim 1993, 2000:195; Taute 1978). In the east Baltic there was trade in seal fat, amber, axes and pottery between hunter-gatherers and farmers (Figure 2.7) (Zvelebil and Lillie 2000). In Finnmark in Arctic Norway, a ‘Mesolithic’ context, south Baltic polished flint axes have been identified (Johansen 1979). Moreover, the occurrence of domesticated plants and/or animals in later ‘Mesolithic’ contexts in central

Indigenous and Availability Models Dennell’s (1983:170-89, 1985, 1992) indigenous model argues that indigenous hunter-gatherers were responsible for the transition to agriculture and pastoralism, through contacts along the European Atlantic fringe. In relation to the ‘Mesolithic’ of western Europe, Price (1987:292-3) states that only Belgium, the southern Netherlands and the Paris Basin show any conclusive evidence of early agriculture being introduced. For Britain and Ireland, Williams (1989:518) concludes, “hunter-gatherers could have played a rôle in the introduction of cultivation and the shifts to farming”. Dennell (1992:93) has recently argued that the ‘Neolithic’ throughout Europe “is primarily about the coexistence of farming and hunter-gatherers, many of whom gradually acquired agricultural resources and developed them on their own accord”. He offers three summaries of the transition to agriculture; (1) southeastern and central Europe and the Alps, (2) the Mediterranean Perimeter, western, eastern and northern Europe and (3) Northeastern Europe, and the wetlands and highlands (Figure 2.2). The second of these regions incorporates Britain and Ireland. In this region, the initial introduction of domesticated plants and animals did not make a significant impact on subsistence, with hunter-gatherer activity still the major form of subsistence. For much of the area, this pattern lasted for upwards of 1000 years or more (Figure 2.3).

9. The case of migration by LBK farmers has recently been strengthened with strontium isotope measurements being taken from human skeletal material from Flomborn and Schwetzingen in Germany. The results indicate that substantial and varied migration occurred during the LBK of the Rhine Valley from approximately 7300-7000 BP within the surrounding regions (Price et al. 2001).

The availability model, which is in fact a variant of the indigenous model, has been developed in the work carried out by Zvelebil and Rowley-Conwy, among others (Zvelebil 1986, 1993b, 1994, 1995a, 1996b; Zvelebil and Rowley-Conwy 1984, 1986; Zvelebil and Dolukhanov 1991; Zvelebil and Zvelebil 1988). As outlined above, Zvelebil rejects the ‘wave of advance’ model claiming that there is “no archaeological evidence of immigration or colonization by new people of areas hitherto occupied by indigenous hunter-gatherers...With the possible exception of the LBK [Linearbandkeramik] culture in Central Europe, and some areas of southeast

18

Figure 2.2 – Three archaeological patterns for the transition to crop agriculture in Europe. The vertical axis represents the length of time taken from the first introduction of cereals and legumes to the establishment of a predominantly agricultural way of life. The horizontal axis represents the importance of farming relative to hunting and gathering during the transition. I, cereals appeared suddenly in the early ‘Neolithic’, along with domestic livestock and pottery. Although there may be a slight overlap with the local ‘Mesolithic’, communities entirely dependant upon foraging seem to apply to large areas of southeast and central Europe and the Alps. II, there is no clear point at which domestic crops and animals and pottery were first used in the ‘Mesolithic’. Sites used specifically for farming are very rare, and foraging seems to have provided most of the food for several hundred years. In areas such as much of the Mediterranean perimeter and western, eastern and northern Europe, agriculture was of minor importance for up to 1000 years after the first appearance of its components. III, crop agriculture is never wholly successful and was often tried intermittently during prehistory. This pattern predominated in much of north eastern Europe and in many highland and wetland areas (after Dennell 1992: figure 5.3).

19

Figure 2.3 – The agricultural frontier in north western Europe at about 6450 BP (after Bradley 1998a: figure 4).

20

Figure 2.4 – The three-stage availability model of the transition to farming (after Zvelebil 1996b: figure 18.1).

Figure 2.5 – The transition to farming along the southern rim of the Baltic in terms of the three-stage availability model. Dates calibrated BC (after Zvelebil and Lillie 2000: figure 3.9). 21

Figure 2.6 – Forager-farmer contacts expected during the earlier part of the availability phase, when co-operation prevails over competition. The example is drawn from the LBK and the Ertebølle (after Zvelebil 1996b: figure 18.5).

Figure 2.7 – Forager-farmer interaction in the east Baltic, c. 5000-4000 BC (after Zvelebil and Lillie 2000: figure 3.10).

22

Europe and possibly Scandinavia,10 illustrates that huntergatherers were familiar with domesticates prior to their significant contribution to subsistence strategies (Jochim 2000; Lüning et al. 1989; Tilley 1996a:48-51; Zvelebil and Lillie 2000).

various sources of evidence to argue that the adoption and spread of farming across the British Isles and southern Scandinavia between 6050 to 5750 BP was triggered by a significant “climate change and its interaction with soil resources, which had a critical impact on the viability and sustainability of early agriculture in these cooler, maritime regions of north-west Europe” (ibid.:11). Their work offers evidence from western Scotland which indicates that the adoption of agriculture coincided with a prolonged period of relatively dry climate conditions and an increase in the annual temperature range that began c. 6050 BP and ended c. 5150 BP, with the driest conditions centred around 5750 BP (Anderson 1998; Anderson et al. 1998).11 Bonsall et al. (2002:15) conclude that there: can now be little doubt that north-west Europe experienced significant climate change around 4000 cal BC [5950 BP] associated with drier conditions and an increase in the annual temperature range. This, they argue, fits well with the phase of stronger meridional atmospheric conditions in the North Atlantic between c. 6050 and 4950 BP inferred from glaciochemical data derived from the Greenland Ice Sheet Project 2 (GISP2) ice core (O’Brien et al. 1995).

External Factors and Subsistence Change External factors, such as climate change, resource degradation or population pressure, have been cited as factors beyond the control of human populations that may have forced subsistence change (Baillie 1992; Binford 1968; Bar-Yosef and Belfer-Cohen 1992; Bonsall et al. 2002; Childe 1928; Cohen 1977; Jarman et al. 1982; McCorriston and Hole 1991; Rowley-Conwy 1984; Wright 1977). For Rowley-Conwy (1984), reduced salinity in the Baltic Sea reduced marine resources, especially oysters, and forced Ertebølle foragers to adopt agriculture. While mussels and cockles continued to be deposited in shell middens and there is continued evidence of the large-scale use of marine resources into the early ‘Neolithic’, the mussels and cockles could not replace the oyster because they are available in different seasons (Pedersen 1997; Rowley-Conwy 1984). However, it is clear that in some regions of Europe and southwest Asia agriculture appeared in areas with stable and abundant resources (see Bar-Yosef and Meadows 1995; Hayden 1990; Price and Gebauer 1992).

The climate change which they argue for would have represented an ‘improvement’ with respect to cereal cultivation through an extension of the growing season. Thus, the change in climate would have facilitated the uptake agriculture and this “provides an underlying mechanism to account for the relatively sudden appearance of the Neolithic throughout this region between 4100 and 3800 cal BC [6050 and 5750 BP]” (Bonsall et al. 2002:17). Thus, not until this climate ‘improvement’ was it possible for the expansion of cereal cultivation and animal husbandry to become widely established in Britain, Ireland and southern Scandinavia (also see McLaren 2000:97; Thomas 2003:71). This hypothesis allows for the possibility of earlier attempts to establish cultivation and/or husbandry into these areas, attempts which would have been unsuccessful, one could argue, due to unfavourable climate conditions.

In the case of Britain and Ireland, Baillie’s (1992, 1995, also see Baillie and Brown 2002) recent work on tree-ring chronologies identified difficulties in constructing a chronology for the early sixth millennium BP, which Baillie (1995:147) has suggested might be due to some sort of environmental pressure, forcing a rapid shift to domesticated plant and animal resources. In western and north Europe, the elm decline, a sharp reduction in elm pollen, was associated with climate change that coincided with the introduction of agriculture. However, this decline in elm pollen is now understood to be a result of disease, rather than climate, and only sometimes occurs after the appearance of agriculture (Peglar and Birks 1993). Recent work by Bonsall et al. (2002) has drawn on

11. This dry phase has been identified in other regions of Scotland (Binney 1997; Tipping 1995) and in The Netherlands (Dupont 1986). Drier conditions are also suggested by lake levels at this time from Scotland (Smith 1996), southern Finland (Sarmaja-Korjonen 2001) and southern Sweden (Digerfeldt 1988).

10. The problems associated with the identification of domesticated and wild animal species and the problems with isolated cereal-type pollen will be discussed on page 47.

23

‘Mesolithic’ and ‘Neolithic’, and as Woodman (2000:219) points out “opinions are much more easily discovered than information based on the observation of actual archaeological data”. Recent debate has centred on the appropriateness and usefulness of attempting to identify patterns across the entirety of Britain and Ireland. Scholars have questioned the characterisation of Britain and Ireland as a homogenous archaeological entity (see Cooney 1997, 2001). Rather, marked regional differences have been identified between such areas as northern Scotland and southern Britain (see Barclay 2001; Woodman 2000:223). This has been part of a wider initiative to break away from viewing the ‘Neolithic’, or for that matter the ‘Mesolithic’, as a unitary phenomenon or ‘package’ (Cooney 1997; Pollard 2000a; Thomas 1993b).

Economic Continuity in Prehistoric Europe Both the ‘availability model’ and ‘indigenous model’ are products of an increased effort, particularly over the last thirty years, to understand the processes associated with the domestication of plants and animals. These approaches have focused on the transition to agriculture from the perspective of the hunter-gatherers, identifying the contribution made by ‘Mesolithic’ and earlier populations. Recently, Zvelebil (1994:36) has summarised these developments in relation to the use of plants in ‘Mesolithic’ Europe. He makes three major points. The first is that the domestication of plants must have occurred before recognisable changes in morphology became apparent in archaeological contexts. Secondly, that domestication is a long process, perhaps occurring as far back as plant and animal manipulation during the ‘Upper Palaeolithic’ (Higgs and Jarman 1972). Thirdly, that there are numerous archaeological and ethno-historical examples of hunter-gatherers in North America (for example, Fritz 1990; Shipek 1989; Watson 1989) and Australia (for example, Chase 1989; Lourandos 1985) who developed highly sophisticated systems of plant use, not necessarily resulting in domestication. Zvelebil (1995a:80) argues, in relation to northern Europe, that such management strategies may have facilitated the local domestication of plants and animals.

The study of the transition from wild to domesticated plant and/or animal resources in Britain and Ireland is inhibited by the limited range of evidence that can shed light on the subject. The acidic soils that occur over much of Ireland, northern Britain and Wales, do not allow for good preservation of organic material (Moore-Colyer 1998; Woodman 2000:219). Thus, there is a mismatch between the question archaeology wants to ask and have answered, and the archaeological data that are available. The fact is that the archaeological and palaeoecological evidence available across much of Britain and Ireland is inadequate for these questions compared to areas of continental Europe. Compounding this, as outlined earlier, is the continual use of the simple ‘chronological’ terms of ‘Mesolithic’ and ‘Neolithic’, which have served to obscure the complex process of economic and subsistence change.

These developments have called into question the generally held archaeological view of the nineteenth and twentieth centuries, that the transition to agriculture was a simple movement from one stage of subsistence organisation to another. The traditional view forces the process to be perceived as a linear phenomenon, allowing no return to foraging once subsistence is characterised by domesticates. But, evidence from southern Scandinavia indicates that the Pitted Ware communities of the fifth millennium BP shifted from subsistence reliant on domesticates to hunting, fishing and some pig husbandry (Zvelebil 1996b:327). In Scotland and Denmark, marine resources contribute significantly to subsistence in postMesolithic contexts, integrated with farming, through the ‘Neolithic’, ‘Bronze Age’ and ‘Iron Age’ (Andersen 1989; Armit and Finlayson 1992; Pollard 1990; Saville and Hallén 1994).

While there is detailed evidence of late ‘Mesolithic’ and early ‘Neolithic’ populations across Britain and Ireland, they do not occur in the same region. The late ‘Mesolithic’ of Scotland is represented by coastal midden contexts with rich faunal and bone/antler artefacts, however, the evidence for the early ‘Neolithic’ is best represented in southern England. This means that characterising the transition is difficult unless one makes generalisations across the whole of Britain and Ireland (see above, Woodman 2000:220). Woodman (2000:224) has argued that, for some, this lack of evidence is viewed as a virtue. A recent survey by Simmons (1996a), commented on the scarcity of evidence in parts of England and Wales, compared to parts of Scandinavia or the Northern European Plain, suggesting: This makes inferences at once easier since there are fewer ‘hard’ points of reference that have to be built into models, and more difficult in the sense that a very little extra information has the potential to require a great deal of revision. (ibid.:39)

Economic and subsistence patterns in Britain and Ireland There is a danger that too much can be made of hunters who gather plants or farmers who fish. Each is a valid activity in its own economy without indicating precognition of a change in lifestyle or reaffirming ancestral traditions! (Woodman 2000:224)

Williams (1989) identified, based on a limited selection of radiocarbon dates, that both hunter-gatherer and domesticated subsistence methods had significant overlap, at least 300 years in Britain and 800 years in Ireland. Her work cast doubt on some dates that have been used to argue the transition, due to issues such as

Much debate has occurred on the economic and subsistence patterns in Britain and Ireland during the 24

problem with conventional radiocarbon dating, and even more for AMS dating. The old-wood effect has been illustrated at a number of contexts. The problem has been identified at wooden structures at Tankardstown (Co. Limerick) (Gowen and Tarbett 1988) where plant remains have been dated significantly younger than the wooden structures. Further, this problem is not confined to structural timber, as indicated by work at Cleaven Dyke (Perth and Kinross) during the early 1990s. Careful examination of the charcoal submitted for radiocarbon dating from a hearth beneath the bank of the cursus showed that the oak wood was already rotten when turned to charcoal, with the possibility that the wood had been dead for up to 200 years before it was placed in the fire (Barclay et al. 1995). In addition, due to the possibility of the AMS dating technique of dating very small pieces of charcoal, there is a greater chance that charcoal chosen for AMS dating could have been affected by vertical and horizontal movement within a deposit (Lanting and van der Plicht 1995:4).

pre-treatment and the old-wood effect. Woodman (2000) has recently reviewed the use of radiocarbon dates by archaeologists. Focusing on their use to study the transition to domesticated plant and animal resources, however, the points raised concern archaeology as a whole and were at the forefront of my mind when studying human skeletal material in Britain and Ireland (see page 84). Woodman (2000:225) points out four main issues, problems and suggestions for using radiocarbon dates, a number of which will be elaborated on (also see Lanting and van der Plicht 1995): 1. The practice by archaeologists of accepting the earliest or latest age in a sequence, that is, the date that is least probable (see Baillie 1991). 2. The assumption that a single radiocarbon date is an adequate estimate for the age of all the deposits within a single layer, which does not allow for post-depositional movement within a deposit. 3. The lack of questioning of the integrity of the samples, for instance charcoal could have accumulated from several events (see Ashmore 1999). 4. The use of dated timbers for construction and other purposes means that an allowance of at least 200 years should be made for radiocarbon dates obtained from such contexts (i.e. the oldwood effect, see Warner 1990).

Related to point four is point three, where the integrity of samples is not questioned. Charcoal assemblages can represent complex events of accumulation of separate pieces of charcoal over long time spans (Ashmore 1999). For this reason, the early dates from the Ballynagilly (Co. Tyrone) timber structure,12 obtained from non-specific charcoal, have been considered dubious (Baillie 1992). The assumed early date from Ballynagilly has been fundamental in arguments about the ‘Neolithisation’ of Ireland and its rejection leads to a need for a majority reappraisal of the Irish ‘Neolithic’ (see Woodman 2000).

Point two has been demonstrated at Dalkey Island (Co. Dublin) where two small shell middens were identified as late ‘Mesolithic’, based on a single radiocarbon date on charcoal and diagnostic ‘Mesolithic’ artefacts. Excavations also revealed domesticated animal bones, prompting Woodman (1976) to propose the contemporaneity of domesticated species with late ‘Mesolithic’ contexts. However, the Quaternary Faunas Project indicates that the middens represent activity and accumulations over a 4000-year period, even allowing for the marine effect. The dates range from 7920-7550 BP (1.000) (7250±100 bp, OxA-4569) (a seal bone) to 34403430 BP (.004), 3400-3060 BP (.973), 3050-3020 BP (.017) and 3010-3000 BP (.006) (3050±70 bp, OxA4567) (a cattle bone), with another cattle bone dated to 5710-5450 BP (.872) and 5410-5330 BP (.128) (4820±75 bp, OxA-4571). Thus, the cattle bones from the midden date to after the early sixth millennium BP and well into the recognised bounds of domestication, refuting Woodman’s initial claims in 1976 (Woodman 2000:228; Woodman et al. 1997). A similar case occurred at Moynagh Lough (Co. Meath), with cattle bones apparently occurring with ‘Mesolithic’ artefacts (Bradley 1991). Within the ‘Mesolithic’ levels, dated by charcoal to 6190-5920 BP (1.000) (5270±60 bp, GrN-11443), a Bos taurus bone was also excavated. However, an Accelerator Mass Spectrometry (AMS) date of 17201410 BP (.990) and 1400-1390 BP (.010) (1660±70 bp, OxA-4268) was obtained, indicating that the cattle bone was in fact early medieval and had penetrated through 50 centimetres of later deposits to lie within the ‘Mesolithic’ layer (Hedges et al. 1997b; Woodman 2000:229).

12. Some the earliest calibrated radiocarbon dates from Ballynagilly (Co. Tyrone) on non-specific charcoal are: 6731-6391 BP (.952), 6371-6344 BP (.028) and 63406313 BP (.020) (5745±90 bp, UB-305); 6656-6651 BP (.004) and 6641-6281 BP (.996) (5640±90 bp, UB-307); 6496-6301 BP (1.000) (5625±50 bp UB-197); 6467-6461 BP (.003), 6450-6168 BP (.906), 6147-6108 BP (.044), 6098-6091 BP (.004), 6072-6057 BP (.012), 6047-6016 BP (.020) and 6014-6000 BP (.010) (5500±85 bp UB559); 6300-5985 BP (.958) and 5973-5939 BP (.042) (5370±85 bp UB-304); 6280-6216 BP (.097) and 62135911 BP (.903) (5295±90 bp, UB-18) (6190-5932 BP (1.000) (5290±50 bp UB-551); 6282-5729 BP (1.000) (5230±125 bp UB-199); 6166-6151 BP (.021), 61056100 BP (.004), 6089-6080 BP (.008), 5996-5855 BP (.743), 5848-5843 BP (.004) and 5829-5749 BP (.220) (5165±50 bp, UB-201).

Point four, the old-wood effect, has been recognised as a 25

provides dates of the events which archaeologists seek to study”.13 He further points out that the period of interest for the ‘Mesolithic’/‘Neolithic’ transition (6200-6000 BP) is when there is thought to have been considerable fluctuation in the 14C levels in the atmosphere. Dates from this time range should therefore be viewed as less reliable than usual (ibid. 240-1, also see Pearson et al. 1986). All of which makes the study of the timing of the transition all the more difficult.

Thus, dating non-specific charcoal assemblages is problematic and does not date a single event. Such assemblages should be treated with caution to avoid creating false chronologies (ibid.:227-9). There is the further difficulty of comparing radiocarbon dates due to the variable accuracy of measurements of samples dated and because of the different laboratories and methods employed. This is particularly true with the growing tendency to utilise AMS dating, rather than traditional/conventional radiocarbon dating methods, and the difficulties in comparing results from these two methods (Baxter 1990; Binder 2000:117; Woodman 2000:256). Pilcher (1993:28) points out, in relation to the reliability of conventional radiocarbon dates in palynology and other archaeological studies of rates of change, that: The routine radiocarbon dates available from most laboratories have been shown to have inaccuracies that only allow the calendrical date to be specified to the nearest half millennium. Attempts to interpret time differences or real ages closer than 500 years by conventional dates are simply not valid. Thus, one could argue that claims of a 300-year overlap in Britain between hunter-gatherer and domesticated subsistence patterns is rather redundant. Such an overlap should be expected when using conventional radiocarbon dating methods, rather than AMS dating (see Baillie 1992). Thus for Britain, and perhaps Ireland, the use of established conventional dates may be inappropriate for the questions which are currently being asked. We may need to reconsider the questions we ask of the evidence. The perceived overlap between hunter-gatherer and domesticated subsistence methods may not be an archaeological ‘reality’, rather a factor of the dating methods employed. Conversely, we may be only identifying part of the overlap of subsistence strategies.

‘Mesolithic’ Activities David Clarke’s influential 1976 paper argued that plant foods would have made a substantial contribution to the ‘Mesolithic’ diet. The lack of such assemblages, he suggested was due to lack of preservation rather than lack of use. Eighteen years on, Zvelebil’s 1994 paper on plant use in the ‘Mesolithic’ illustrated that human-plant and human-animal relationships had been well-established prior to the ‘Neolithic’. Drawing on macrobotanical remains, palynological data, artefactual evidence and the human biological record, Zvelebil contends that these relationships were both crucial and fundamental to the ultimate development of agriculture in later prehistory. Zvelebil concludes that ‘Mesolithic’ populations were involved in “intensive plant use strategies” (1994:64). There are problems with identifying the mechanism that produced cereal-type pollen, as it is produced by both wild and cultivated grasses (Andersen 1979; Dickson 1988; Edwards 1989a; O’Connell 1987). Therefore, cereal-type pollen in post-glacial context could be accounted for by genetic mutation of wild grass. This has been documented by O’Connell’s work in the Connemara National Park and at Lough Namackanbeg (Co. Galway) He identified, independent of any human manipulation, cereal-type pollen possibly due to a short lived polyploidy of native grass species. The cereal-type pollen at Connemara National Park was dated to 7824-7801 BP (.039) and 7793-7664 BP (.961) (6915±40 bp, GrN13145) and for Lough Namackanbeg, with inference from its position in the pollen profile, O’Connell suggests dates of about 7570 bp. This led to the recommendation “that single or occasional cereal-type (wheat of Triticum type) pollen, recorded in contexts where corroborative evidence

Archaeology as a discipline must accept and acknowledge the different limitations of the various dating methods that have been and are currently being employed. This is particularly true of ‘Neolithic’ studies, which tend to operate on shorter time scales than ‘Mesolithic’ studies (Gijn and Zvelebil 1997:3). While much hope is invested in AMS dating, and rightly so, the questions asked and expectations of results must be sensitive to the limitations of the technique. As Woodman (2000:229) concluded “in most cases 14C dating rarely

13. Woodman does point out that an obvious exception is the AMS dates obtained from worked organic materials.

26

of non-tree pollen may be interpreted as the creation of small clearings (Smith et al. 1989). From Star Carr (North Yorkshire) recent palaeoecological and radiocarbon dating evidence indicates that repeated deliberate burning of woodland occurred over several centuries from approximately 10700 bp (Day and Mellars 1994).

for human activity is lacking or weak, be not regarded as signifying arable farming” (1987:220). Similar identifications have been made in pre-elm decline contexts in Yorkshire (Bush and Flenley 1987) and from Oakhanger (Hampshire) (Rankine et al. 1960). Zvelebil’s argument for intensive plant use strategies by ‘Mesolithic’ populations was strongest in eastern Europe at Sarnate, Latvia. From Sarnate, the evidence consisted of a 40-centimetre thick layer of waterchestnuts over several square metres deposited around hearths and other occupation debris. The evidence in Britain and Ireland was less spectacular (Zvelebil 1994:Table 2). For example, a large number of charred hazel nuts have been excavated at Broom Hill (Hampshire) (O’Malley 1976, 1978). However, recent evidence of gatherers in the Southern Hebrides Mesolithic Project at Staosnaig (Argyll and Bute) on the Isle of Colonsay, off the west coast of Scotland, shows the exploitation of a very large number of hazel nuts and a substantial quantity of lesser celandine (Ranunculus ficaria L) (Mithen 1999b:54; Mithen and Finlay 2000; Mithen et al. 2001). The evidence from the vicinity of Staosnaig suggests intensive exploitation of hazel nuts at around 7716±26 bp (85508410 BP). This involved tens or hundreds of thousands of hazel nuts being burnt or roasted, along with substantial use of lesser celandine and crab apples. While it is difficult to estimate the number of trees involved, as many as 5000 trees could have been exploited. Mithen et al. argues that if this occurred over a short period it may have affected the local vegetation (Mithen et al. 2001:233). Earlier environmental work by Andrews et al. (1987) which recognised a sudden decrease in all tree species at around 7800 BP in the reconstruction of the vegetation history of southeast of the island may have identified this exploitation.

Clearing woodland, especially by fire, would attract grazing animals (Jacobi et al. 1976; Mellars 1976b; Simmons 1969, 1996a, 1996b). Thus, one can envisage ‘Mesolithic’ populations involved in animal husbandry, keeping animals in natural and human-created clearings (Pryor 1988:72). While this cannot be identified as true domestication, it has been argued that it illustrates that complex relationships existed between humans, plants, and animals in the ‘Mesolithic’. However, there are problems in distinguishing between natural and anthropogenic effects on the woodland. Woodland is subject to natural processes, including disease, death, wind throw, lighting strikes and animal grazing (Buckland and Edwards 1984; Edwards and Whittington 1997; Peterken 1996). However, the evidence in Britain, and to a certain extent in Ireland, for human impact on the vegetation is not seen elsewhere in northwest Europe, which remains quite puzzling (Bell and Walker 1992; Woodman 2000:246). Bonsall et al. (2002:11) research in the Oban region (Argyll and Bute) suggests that prior to 5750 BP cereal-type pollen occurrence and high charcoal concentrations reflect climate changes rather than indicators of anthropogenic modification of vegetation by hunter-gatherers and early farmers. They argue that those periods of increased charcoal frequencies across the Oban region at c. 9000 BP, c. 6400 BP, c. 5900 BP and c. 4250 BP correlate with phases of relative dry climate identified from peat humification analysis conducted in north west Scotland (Anderson 1998; Anderson et al. 1998). A similar correlation has been reported by Tipping (1996) from northern Scotland at 8950 BP, where increased charcoal counts in pollen diagrams witnessed a change to a drier climate. Thus it is argued that natural fires played an important role in the ecology of early to mid-Holocene woodlands, especially in pinewood areas (Bennett 1989, 1995; McVean and Ratcliffe 1962) and deciduous woodlands (Moore 1996).

Evidence indicates that ‘Mesolithic’ populations in Britain and Ireland were modifying their environment through deliberate manipulation and through clearing and burning. They were also taking advantage of natural clearings of woodland. In some cases lithics have been found in the pollen contexts themselves or nearby (Brown 1997; Bush 1988; Cummins 2000; Day 1993, 1996a; Day and Mellars 1994; Edwards 1989b, 2000b; Edwards and Mithen 1995; Edwards and Ralston 1984; Edwards et al. 1991; Gearey et al. 2000a; Innes and Blackford 2003; Jacobi et al. 1976; Jones 1976; Knox 1954; McCullagh 1989; Mellars and Dark 1998; Mithen 2000a, 2000b, 2000c; Mithen et al. 2001; Radley and Mellars 1964; Richmond 1999; Simmons 1969, 1975, 1996a, 1996b; Simmons and Innes 1987, 1988, 1996a, 1996b, 1996c; Smith 1970; Smith et al. 1989; Spikins 1999, 2000a; Spratt and Simmons 1976; Thomas 1989; Tipping 1994; Tipping et al. 1993; Turner and Hodgson 1983; Williams 1985; Zvelebil 1994, however see Edwards 2000a). Manipulation by thinning-out of the canopy is witnessed by charcoal concentrations in the stratigraphy of ‘Mesolithic’ contexts and by the presence of light demanding plant species within pollen diagrams, such as sorrel (Rumex), ribwort plantain (Plantago lanceolata) and heather (Calluna) (Fisher 1982:300; Richmond 1999:6). At Peacock’s Farm, an increase in the proportion

In Ireland, there is evidence in the way of AMS dates that domesticated animals occur in ‘Mesolithic’ contexts (see Woodman 2000). Within Irish archaeology it is accepted that the domesticated dog (Canis familiaris) existed in pre-‘Neolithic’ contexts, for example at Mount Sandel (Co. Derry) (Van Wijngaarden-Bakker 1985). However, since 1976 cattle, and to a lesser extent sheep, have also been thought to occur in ‘Mesolithic’ contexts (Woodman 1976). Despite this, it was not until AMS dating of the material that the animal bones could be directly associated with pre-6000 BP contexts. While evidence from Dalkey Island and Sutton (Co. Dublin) have been dismissed due to problems of contamination and misidentification, secure examples still exist (see Woodman et al. 1997). Ferriter’s Cove (Co. Kerry), has produced an AMS date of a Bos taurus’ ulna dated to 6450-6420 BP (.028), 6410-6170 BP (.959) and 613027

causewayed enclosures and long mounds were constructed within such boundaries (Rouse and Evans 1994; Davies 1999, see page 210).

6120 BP (.014) (5510±70 bp, OxA-3869) (ibid.:138). Kilgreany Cave (Co. Waterford), has produced an AMS date of a Bos taurus’ tibia dated to 6170-6130 BP (.086), 6120-5840 BP (.746) and 5830-5750 BP (.168) (5190±80 bp, OxA-4269). In Britain, subsistence and mobility have been inferred from the animal bones recovered from Star Carr (South Yorkshire) (Caulfield 1978; Clutton-Brock and Noe-Nygaard 1990; Day 1996b; Legge and RowleyConwy 1988; Mellars and Dark 1998; Pitts 1979).

Analysis has indicated that many constructions, such as causewayed enclosures at Croftton, Knap Hill, Windmill Hill (Wiltshire), Offham, Combe Hill, Whitehawk (East Sussex); Bury Hill, Barkhale, The Trundle (West Sussex) and Maiden Castle (Dorset) and the long mounds and cairns at Giants’ Hills 2 (Lincolnshire), Dyffryn Ardudwy (Gwynedd), Barclodiad y Gawres, Bryn yr Hen Bobl (Isle Of Anglesey) and Easton Down (Wiltshire) were built in woodland or small clearings, which in some cases regenerated rapidly after construction (Drewett 1994:19; Drewett et al. 1988:35-6; Evans and Simpson 1991; Hemp 1936; Lobb 1995; Lynch et al. 2000; Oswald et al. 2001; Powell 1973; Powell and Daniel 1956; Thomas 1999a:42; Thomas 1982; Whittle et al. 1993, 1999a). A study of the molluscan evidence from the dry valley sequence at Brook (Kent) produced evidence of clearings followed by regeneration (Kerney et al. 1964:165).

The empirical evidence, which has been gathered over the last 40 years, illustrates that ‘Mesolithic’ populations manipulated their environment and also took advantage of natural clearings to control plant and animal populations. ‘Mesolithic’ populations were involved in animal husbandry and horticulture. The human social behaviour documented in the ‘Mesolithic’ archaeological assemblages and contexts laid the foundations of the later relationships and dependency between humans, plants and animals. Rather than the ‘Neolithic’ representing the consolidation of this interdependency between humans, plants and animals, the ‘Neolithic’, like the ‘Mesolithic’, witnessed a process of gradual intensification that continued into the third millennium BP of Britain and Ireland.

Using pollen to reconstruct environments at the local scale is problematic as it lacks temporal resolution, due to movement by water, animals and wind (Bonny 1976; Evans and O’Connor 1999:70-1; Peck 1973). Research in the Shetland Islands has identified that pollen, given appropriate conditions, can travel tens to hundreds of kilometres (Tyldesley 1973). There are taphonomic issues that must also be addressed. The varied preservation of certain pollen types, due to rainfall and soil alkalinity, means that identifying woodland composition and density is unreliable, especially on a local scale (Scaife 1987:1267; Scaife and Burrin 1992:83). Recently, the analysis of insect remains from ‘Neolithic’ contexts has shown that in some contexts the proportion of tree and wooddependant Coleoptera decreased. However, at others the proportion of Coleoptera remained significant into the ‘Neolithic’ and even in contexts identified as later ‘Neolithic’, for example Rowlands Track (Somerset) and Buscot Lock (Oxfordshire) (Robinson 2000a).

‘Neolithic’ Activities Woodland Clearances Recent literature has argued that British and Irish ‘Neolithic’ populations adapted and developed new subsistence practices into their established practices (see Holgate 1988a; Richmond 1999; Robinson 2000a; Thomas 1999a; Whittle 1990b; Williams 1989). Despite the evidence for woodland clearance in the ‘Mesolithic’, the traditional view has been that the ‘Neolithic’ population transformed the ‘wildscape’ into an open landscape (Darvill 1987c; Mitchell 1972a; Pennington 1975). In this traditional view, the opening of the landscape was a necessary practice for the development of agriculture. But the limitations of reconstructing woodland from pollen diagrams and fossil mollusan assemblages have been recognised since the 1970s, which complicates the whole issue of identifying woodland clearances (Bonny 1976; Carter 1990; Davies 1999; Edwards 1979, 1993, 1998; Needham and Macklin 1992; Peck 1973; Richmond 1999; Rouse and Evans 1994; Scaife 1987; Scaife and Burrin 1992).

An overview of the complex picture of woodland across Britain has recently been undertaken by Richmond (1999). The conclusion reached for the sixth and fifth millennium BP is that the traditional view of the landscape being opened up for widespread agricultural activity cannot be maintained. Richmond (1999:31) concludes: From East Anglia to Wales and from Central Southern England to Scotland communities are seen to have had a limited effect upon the landscape, which was often not very different from earlier Mesolithic practices. Similar conclusions have been made elsewhere. In Ireland, Cooney (2000a:219) has stated that the land was still dominated by woodland during the ‘Neolithic’ (also see O’Sullivan 1996). Needham and Macklin’s (1992) review of over 100 published papers on Holocene river alluviation concluded that it was only towards the end of the fifth millennium BP (late ‘Neolithic’) that rates increased related to human activities. Earlier instances of run-off related to periods of increased precipitation or

Work based solely on molluscan assemblages must be questioned in light of Carter’s 1990 paper and earlier work by Edwards (1979). Their work demonstrated that care should be taken in drawing inference from molluscan assemblages, as without good temporal and ecological resolution definition, patterns can be exaggerated. Attempting to identify woodland-grassland boundaries is problematic unless a spatially-oriented multi-sampling strategy is employed, as it has been demonstrated that the molluscan response does not exactly match the vegetation structure. This work is a major consideration due to evidence which identifies that 28

The fifth millennium BP witnesses a continuing pattern of the construction of stone, timber and earthen structures within small patches of cleared woodland. Hence again clearances were created for construction. Most palynological data for the fifth millennium BP come from buried soils beneath such constructions and from the associated ditches (however see Evans 1993 and Evans et al. 1993 for off-site environmental study). Richmond (1999:61, also see Allen 1997:127; Lynch et al. 2000:45; Thomas 1999a:222-3) states that: This may have contributed to a false vegetational picture across the landscape. In other words, the identification of open habitat species at twenty monument sites does not have to indicate an open landscape; it may rather suggest spurts of clearance around the monuments in a landscape dominated by forest. The picture that prevails throughout the early tenth to the late fifth millennium BP is one of a mosaic forest of irregular clearings restricted spatially and temporally, associated with mobile occupation (see Chapter 6). Rather than creating an open agricultural landscape, a wooded landscape persisted with limited alteration due to anthropogenic processes. Where woodland reduction did occur, it was more likely due to natural ecological factors, such as thin or poor soils and high altitude (Richmond 1999:78). When human activity is a factor, it is more generally associated with the creation of grazing areas and for constructions.

rising sea levels. Clare, Clapham and Wilkinson (Clare 1995; Clapham et al. 1997) have been examining Britain’s submerged woodlands in an attempt to reconstruct their ecology. Thus far, evidence suggests a mosaic character for late ‘Mesolithic’/early ‘Neolithic’ woodland: rather “than being a homogenous mix of different tree species, the forest on dryland consisted of stands of trees in which particular species were more common than others” (Clare 1995, also see Allen and Gardiner 2000a, 2000b; Barker and Webley 1978). Further, the size of the trees and their distribution indicates that the woodland would have consisted of trees of varied diameter (Figure 2.8). Thus, the clearing of such woodland may have been highly selective, based on the ease with which specific types of trees could be removed and the specific needs of the population. Evidence from Sweet Track (Somerset) (Figure 2.9) built of oak, ash, hazel, willow, poplar, alder, dogwood, birch timbers, indicates that construction began in the early sixth millennium BP. According to tree-ring dating, the woodland timber resources used for the Sweet Track (Somerset) had been carefully managed for up to 150 years and then selected prior to felling (Hillam et al. 1990:215; Zienkiewicz 1996:2). Work by Edwards (1985) in Co. Tyrone in Ireland supports the argument that vegetation was a mosaic of different species, with limited clearings for cultivation and to facilitate animal husbandry (Cooney 2000a:41-2; Mitchell 1976, 1986) (Figure 2.10).

Woodland Regeneration? According to Clare (1995), some parts of the woodland may have been impenetrable, offering a potential explanation for the patchy and ad hoc distribution of archaeological artefacts and features in the landscape. Similarly, Brown suggests (1997:138 and 142) that settlement location in ‘Mesolithic’ and ‘Neolithic’ Britain may have been limited to natural clearings in the woodland, thus characterising the occupation patterns as opportunistic dwelling in naturally open areas in an otherwise wooded landscape (Clapham et al. 1997:269, also see Chapter 6). Clearances, both natural and anthropogenic, occurred in many places for reasons other than cultivation (see Brown 2000). In addition to occupation, clearings were used to build stone, timber and earthen construction and also would have affected access for occupation. Brown argues (1997:139) that evidence from waterlogged contexts throughout western and northern Europe indicates woodland management rather than the clearing of mature trees. Further, Brown suggests that in “Britain, it is not until the mid-late Neolithic that we see the construction of structures such as palisaded causeway enclosures which would have required substantial quantities of large timbers” (Brown 1997:139). Clare’s (1995) interpretation suggests that, if the majority of the smaller trees were cleared and the larger trees were retained, the result would have been irregular clearings, “producing a kind of ‘parkland’ appearance” (also see Fishpool 1999:131). The result is that the stone, timber and earthen constructions were not set in open agricultural landscapes, but rather were constructions set in small clearings

Towards the end of the sixth millennium BP, evidence suggests that, particularly in south eastern England, there was a re-establishment of woodland (Bradley 1978a, 1984a; Davies and Wolski 2001; Edwards 1979; Evans 1990; Thomas 1999a). In the Avebury region, Wiltshire, evidence exists that scrub or woodland was re-established at Knap Hill, Cherhill, South Street, Dean Bottom, Roughridge Hill, West Kennet, Millbarrow, Windmill Hill and Easton Down (Connah 1965; Evans and Smith 1983:109; Whittle 1994a, 1997b:140; Whittle 1993:232). In relation to the Easton Down long mound, Whittle et al. (1993:228) argue that: The combined evidence of woodland from the pre-barrow subsoil hollow and the speed and totality with which secondary woodland invaded the infilling ditches indicates for the first time and unequivocally the presence of woodland on the high chalklands in the middle Holocene. Questions have been raised concerning this evidence. While recent work in the Avebury region has relied upon a variety of evidence, this has not always been the case. Work at Horslip, Stonehenge (Wiltshire), Wayland’s Smithy II (Oxfordshire) and Brook (Kent) relied upon molluscan analysis of buried soils (Evans 1984; Whittle 1978:37); problems with this evidence have been discussed above. The identification of woodland regeneration in other areas of southern Britain is also questionable. In areas such as the East Anglian Breckland and the southern chalk uplands, evidence exists for increased clearance in the late ‘Neolithic’, rather than regeneration of woodland (Godwin and Tallantire 29

Figure 2.8 – Distribution of all stump diameter groupings at Hightown (Merseyside), 1994 (after Clapham et al. 1997: figure 3).

Figure 2.9 – Part of Sweet Track. Railway site. Somerset Levels (after Hillam et al. 1990).

30

Figure 2.10 – Provisional map of woodland types 5000 years ago for Britain and Ireland (after Bennett 1989: figure 1). time.

1951; Smith 1981:208). Finally, analyses of valley alluvial and colluvial deposits from the late ‘Neolithic’ have not produced the sedimentary record expected with woodland regeneration (Bell 1982; Needham and Macklin 1992).

(1999a:30) There is a current mismatch between the questions being asked and the level of palynological data and temporal resolution currently available. The situation is complicated further by lack of accuracy in inferring anthropogenic activities. Peglar’s (1993a, 1993b; Peglar and Birks 1993) work at Diss Mere in Norfolk clearly illustrated that woodland decline in the mid-Holocene was directly associated with a pathogenic attack connected with the so-called ‘elm decline’. Assuming that Holocene changes to the woodland component of the landscape were primarily due to direct human activities is incorrect. Palynological studies should continue to identify the cause of vegetational changes across the

For Thomas the question of the timing of the reestablishment of woodland in the late sixth millennium BP is not consistent with current interpretations of earlier periods: If we are not clear how much of the vegetational disturbance of the late fourth and early third millennia [BC, late sixth and early fifth millennia BP] can be attributed to human agency, it is difficult to argue that this activity decreased over 31

1993; Rowley-Conwy 2000:51). Such finds have been common in mainland Europe, such as Denmark, however they are rare in Britain and Ireland. Legge (1989) argues that cereals have not been recovered, in part due to the lack of large-scale flotation during excavation. However, Moffett et al.’s (1989) review showed that this was not the case. Of the 26 contexts listed where flotation (usually small-scale) has been used, only 922 cereal grains were recovered, that is an average of 35.5 cereal grains per context. While there are apparently exceptions like Balbridie, generally the dietary contribution of cereals does not appear to be significant. However, Rowley-Conwy (2000, also see Legge 1989) has recently used these exceptional contexts in Britain to propose that taphonomic factors have affected the amount of cultigens recovered from across Britain. He argues that taphonomic factors, such as storage procedures and building methods, have contributed to the low quantity of cereals being recovered in Britain. Jones (2000:80-1, also see Monk 2000) has suggested processing and consumption factors which may affect the survival of cereals and hazel nuts. She argues that since hazel nut shell is a by-product of consumption, it will be preserved in occupation debris, while cereal grains are part of the plant normally eaten and will less likely be preserved. She goes on to note that chaff and straw, the by-products of cereals, are less likely to occur in occupation debris as they are used for fodder (also see Hillman 1981, 1985), for bedding for domestic animals and if burnt they turn to unrecognisable ash (see Boardman and Jones 1990).

landscape, be they anthropogenic or environmental. The picture emerging is one of limited impact of humans and other factors for the majority of the early tenth to the late fifth millennium BP, punctuated by localised clearances, be they anthropogenic, environmental or a combination of both; distinguishing between these can be difficult (see Turner 1964). Cereal Domestication For several decades, cereal pollen has been identified and has been presented as evidence for the emergence of agriculture. Contexts in Ireland such as Ballynagilly (Co. Tryone), Newferry (Co. Antrim) and Cashelkeelty (Co. Kerry) have produced very early dates from the seventh millennium BP, leading to suggestions that domesticated cereals were adopted by ‘Mesolithic’ populations (Bush 1988; Edwards and Hirons 1984; Groenman-van Waateringe 1983). Similar evidence has been identified at over 20 contexts throughout Britain and Ireland (see Edwards 1989a). For over 50 years this issue has been dogged by problems in differentiating cultivated pollen from native, wild species (Edwards 1989a, 1998) and also by preservation problems of macroscopic grains (Jones 2000; Rowley-Conwy 2000). There are no stringent criteria distinguishing between wild and domesticated species of plants, leading Williams (1989) to argue that it is premature to utilise much of the published palynological evidence to argue for the domestication of cereals. Employing stringent criteria, Williams (1989:515-6, also see Rowley-Conwy 1995a) identifies only three contexts with satisfactory evidence for domesticated cereals: Cashelkeelty (Co. Kerry) at 68906820 BP (.062), 6810-6440 BP (.923) and 6430-6410 BP (.015) (5845±100 bp, UB-2413), Soyland Moor (West Yorkshire) at 6850-6840 BP (.013), 6830-6820 BP (.006) and 6800-6410 BP (.981) (5820±95 bp, Q-2394) and North Gill (North Yorkshire). However, recent radiocarbon dating of North Gill has produced uncertain results (Simmons and Innes 1996d).

The lack of cereals in archaeological assemblages has been attributed to the recovery methods employed. It has been argued that the manual recovery of plant remains, or dry-sieving with a large mesh, is likely to under-represent small cereal grains, which Jones (2000:79) suggests “may be responsible for the overwhelming predominance of hazelnuts at earlier excavations where these were the only methods of recovery”. She cites recent analysis of the plant remains from Hambledon Hill Complex (Dorset), where both manual recovery and flotation/wet sieving occurred alongside each other. Here “manual recovery methods frequently retrieved hazelnut shells but never cereals remains, in marked contrast with the results from flotation and wet-sieving” (ibid.). While the contribution of cereals cannot be disputed, I argue that it is just one component, of varying significance, in a diverse subsistence strategy.

Charred cereal seeds have been recovered from ‘Neolithic’ contexts including Carn Brea (Cornwall), Dorset Cursus, Hambledon Hill Complex (Dorset), Durrington Walls, Windmill Hill (Wiltshire), Mount Farm, Barrow Hills, Barton Court Farm (Oxfordshire), Briar Hill (Northamptonshire), Great Wilbraham, Hembury (Devon), Offham Hill, Belle Tout (Sussex), Beckton Farm (Dumfries and Galloway), West Row Fen, Pakenham (Suffolk), Springfield, Stumble (Essex), Spong Hill, (Norfolk), Deeping St. Nicholas (Lincolnshire), Skara Brae (Orkney Island), Scord of Brouster (Shetland Islands) and Fengate (Cambridgeshire). In Balbridie (Aberdeenshire), Tankardstown (Co. Limerick), Ness of Gruting (Shetland Islands) and Lismore Fields (Derbyshire), large assemblages of charred cereals have been recovered (Fairbairn 1999; Monk 1988; T. Pollard 1997; Richmond 1999:32-3).

Grain impressions have been found on ‘Neolithic’ pottery, with about half of all examples coming from Windmill Hill (Wiltshire), offering indirect evidence of cultigens (Dennell 1976; Richmond 1999:33). ‘Field systems’ have been identified from East Anglia in England and at Co. Mayo and Co. Clare in Ireland (Figures 2.11 and 2.12) (Byrne 1986; Caulfield 1983; Caulfield et al. 1998; Cooney 1997:28, 2000a; MooreColyer 1998:22; Pryor 1978). ‘Ard’ marks have been identified beneath the South Street (Wiltshire) earthen mound (Rowley-Conwy 1987). Also, a possible ‘Beaker period’ field system has been identified at Cherhill (Evans and Smith 1983:109; Evans et al. 1993:188-9). However, these attributes do not constitute reliable

The burned down rectangular structure at Balbridie (Aberdeenshire) (Figure A3.7) resulted in the preservation of some 20000 cereal grains, including emmer, barley and bread wheat (Fairweather and Ralston 32

differentiating on morphological grounds between wild and domestic species and preservation problems (see Caseldine 1980; Lynch et al. 2000:45; Maltby 1985; Moore-Colyer 1998:15). While the morphological differences between wild and domestic pig has recently been shown to be less complex than previously supposed (Albarella and Serjeantson 2002; Payne and Bull 1988; Rowley-Conwy 1995b), the same can not be said for wild and domestic cattle (however see Rowley-Conwy 1995b). There is significant overlap in the size of wild and domestic animals, especially between domestic bulls and wild cows (Grigson 1999:213). By analysing insect remains from a number of excavations, the increased role of domestic mammals at certain contexts in the ‘Neolithic’ has been identified through a great rise in the number of scarabaeoid dung beetles. Such beetles would have thrived with the increased number of domestic mammals (Robinson 2000a).

evidence for cultivation. They represent secondary evidence (Moore-Colyer 1998:22; Richmond 1999:34) and can be interpreted in several other ways. The pottery with cereal impressions could have come from other areas, even mainland Europe (see Chapter 6 for the movement and circulation of artefacts). The constructions that have been identified as field systems may have served some other purpose. The ‘ard’ marks could be identified as activity linked to construction of the earthen mounds. Evidence from the second half of the seventh to the late fifth millennium BP points to a general lack of cereal pollen and “food consumption rather than crop processing” (Richmond 1999:34). There is the general assumption that cereal pollen indicated cereal production. However, this is generally assumed rather than demonstrated (Richmond 1999:79). Further, there is the possibility that identified cereal pollens are merely genetic mutations of wild grass (O’Connell 1987). The existence of cereal pollen has been presented as evidence of cereal use, but the current evidence suggests that it was ‘an optional food’ rather than an item of subsistence as witnessed in later periods (Richmond 1999:34). The current evidence for much of Britain and Ireland from the middle of the seventh to the late fifth millennium BP can be summarised as very patchy use of cereals, with the odd spectacular find.

Discussions related to domestic livestock have all too often degenerated into arguments of their ‘special significance’, ‘symbolic’ or their importance as ‘ranked foods’, rather than elements of subsistence strategies (see Ashbee 1970; Bradley 1984a; Cooney 1997:25, 2000a:98-99; Entwistle and Grant 1989:207; Thomas 1997a:59, 1999a:29; Whittle et al. 1999a), an explanatory practice which has also occurred in relation to cereals (Pryor 1988; Thomas 1993b; Fairbairn 1999, 2000; Thomas 1996c; Whittle 1997b:137; Whittle et al. 2000). This is due to their consistent association with socalled ‘ritual’ and ‘funerary’ contexts (Richmond 1999:75-6) and their spatial separation from wild species at certain contexts. For example, at Durrington Walls, Horslip, Marden and Woodhenge (Wiltshire) wild species are confined to the outer ditches and domestic species were placed in more central locations (Ashbee et al. 1979; Pollard 1995a; Richards and Thomas 1984). Further, there has been much discussion on the significance of so-called ‘head and hooves’ assemblages at Hemp Knoll (Wiltshire) and at Irthlingborough (Northamptonshire) which contained at least 184 domestic cattle skulls and a few auroch skulls (Ashbee 1970; Davis and Payne 1993; Grant 1991; RobertsonMackay 1980). One may ask why the ‘special significance’ cannot be bestowed upon the wild species. Despite what symbolic value we give the assemblages today, eating them would still contribute to the subsistence of the population.

Domestic Livestock As outlined earlier, woodland clearances were restricted spatially and temporally, an interpretation that is further reinforced by faunal assemblages that are dominated by species suited to a woodland habitat (see Barrett et al. 1991:20). In relation to domestic livestock, cattle and pig are dominant, with sheep generally poorly represented. The nature of the woodland clearances in the ‘Neolithic’, as in the ‘Mesolithic’, would have been suitable for attracting grazing animals, be they wild species like roe deer, red deer or auroch, or domestic livestock like cattle or pig (Richmond 1999:34; Whittle 1997b:124). Cattle are identified as particularly important from the middle of the seventh to late fifth millennium BP, with dairy herding seen as a primary activity for ‘Neolithic’ populations (Barker 1985:200, 206; Harrison 1985:87-8; Legge 1989). During the fifth millennium BP, evidence indicates that pigs out-represented cattle in archaeological assemblages (Figure 2.13) (Grigson 1981; Whittle 1978). The introduction of domesticated animals, such as cattle, sheep and goats, in Ireland has been seen as particularly important to the population as there are virtually no large mammals in Ireland prior to domestication. The Quaternary Faunas Project identified no red deer bones (Cervus elaphus) radiocarbon dated in any ‘Mesolithic’ contexts, leaving wild pigs (Sus scrofa ferus) as the only large mammal available for subsistence and domestication (Van Wijngaarden-Bakker 1985; Woodman 1976; Woodman and O’Brien 1993; Woodman et al. 1997).

Domestic livestock, particularly cattle and pig, are found in many contexts from the middle of the seventh to the late fifth millennium BP. With mobility (see Chapter 6) and woodland cover continuing to be features of the ‘Neolithic’, livestock management or animal husbandry would have been an important element of a diverse subsistence strategy. Wild fauna were also utilised, and while not considered throughout the literature as holding the same ‘special significance’ as newly domestic species, contributed to the overall subsistence strategy of ‘Neolithic’ populations – a point that must be acknowledged.

As with the identification of cereal, studies of domestic livestock have been influenced by problems of 33

Figure 2.11 – The field systems at Rathlackan (Co. Mayo) and Roughaun Hill (Co. Clare) (after Cooney 2000a: figure 2.7).

Figure 2.12 – The layout of Céide Fields (Co. Mayo) (after Cooney 2000a: figure 2.2).

34

Figure 2.13 – Faunal assemblages from southern Britain: Ratio of pigs:cattle:sheep (after Thomas 1999a: figure 2.5). the middle of the seventh millennium BP continued to contribute to the subsistence of the population at different times of the year. Excavated evidence suggests that ‘Neolithic’ populations relied heavily on wild fauna and wild flora from woodland, rivers and coastlines (Richmond 1999:76). Grigson (1981:223) states that the view that domestic species were dominant in ‘Neolithic’ contexts has produced a false impression of domestic livestock predominance. Certain bones recovered that have been identified as domestic may in fact be wild species due to the problems of identifying wild or domestic variants that was outlined previously. Interestingly, recent DNA evidence suggests that British cattle were introduced from outside as fully domestic animals, indicating that wild cattle as not domesticated in Britain (Troy et al. 2001).

Continued Utilisation of Wild Resources Despite new domestic flora and fauna becoming integrated into ‘Neolithic’ subsistence strategies of Britain, virtually all the vertebrate fauna utilised through the ‘Mesolithic’ period survive into the ‘Bronze Age’. Down to at least the fourth millennium BP, wild fauna species are recovered in small quantities from archaeological contexts, including auroch, red deer, roe deer and boar (Ashbee 1970:158ff; Grigson 1966, 1981; Richmond 1999). Bird and fish bones have been found in the ‘Neolithic’, but generally in small quantities. It has been argued that this situation is related to recovery and post-depositional destruction rather than the nature of the subsistence practices (Louwe-Kooijmans 1993:81). Marine resources, such as shellfish continued to be of great importance in Scotland well into the ‘Iron Age’ and also in Ireland in the ‘Neolithic’ (Armit and Finlayson 1992; Liversage 1968; Österholm and Österholm 1984; Pollard 1990; Saville and Hallén 1994). Cooney (2000a:219) sees fishing, gathering and hunting as continuing to be utilised on a seasonal basis in Ireland during the ‘Neolithic’.

Human diet: Bone stable isotope analysis Stable isotope analysis of human bone collagen and teeth is a well-established technique, widely applied in archaeology since the 1970s in South Africa and North America (Richards 2000a:123, 2001; Vogel and van der Merwe 1977). The technique can provide information on the protein sources in the diets of the last 5-15 years of a person’s life, depending on bone element samples but regardless of an individuals age or sex (Ambrose 1993; Lovell et al. 1986; Richards 2000a:123; Schwarcz and Schoeninger 1991, however see Schoeller 1999). All foods which humans consume have specific stable isotope ratios, or ‘signatures’, that are maintained (with

With the continued dominance of woodland across the landscape, some wild flora would have been plentiful at different times of the year (Zvelebil 1994). The important role of wild flora within ‘Neolithic’ subsistence strategy has long been argued (Entwhiste and Grant 1989; Greig 1991; Moffett 1991; Moffett et al. 1989; Thomas 1991a, 1999a). Preservation of such remains is the major problem, with barks, roots, stems, leaves, flowers, fruits and fibres only surviving under certain exceptional conditions. Those subsistence strategies utilised prior to 35

analysis. This is particularly true of radiocarbon dates carried out using the ion-exchange procedure that seem to alter the isotopic results slightly (Rick Schulting personal communication; Schulting and Richards 2002b:163). However, these samples do produce data which can be discussed more broadly (see Appendix 1).

modification) through the food chain. There are two stable isotopes of carbon (12C and 13C) and the ratio of the two is the carbon stable isotope ratio, denoted by the symbol į13C. For nitrogen, the two isotopes of interest are 15 N and 14N, and their ratio is called the į15N value. The 13 į C and į15N values of various foods are fairly well known.

Palaeodietary Evidence in Britain and Ireland

Measuring the į13C provides information on the levels of marine (fish, shellfish) compared with terrestrial (milk, blood or meat) sources of protein in the diet. Individuals who get a predominant (>90%) amount of their protein from marine sources will have į13C values close to -12‰. Those who have no marine protein in their diet will have į13C values closer to -20‰. Those who derive 50% of their protein from terrestrial, and 50% from marine sources, will have į13C values in the middle, about -16‰. Measuring the į15N provides information on the relative importance of plant and animal protein sources. There is an increase of about 3‰ in the į15N each step up the food chain. Therefore, if plants have an average į15N of about 3‰, herbivores that consume those plants have į15N values of 6‰ and carnivores consuming those herbivores will have į15N values of about 9-10‰. These hypothetical values (based on the fact that the į15N of air is 0‰) are close to the published values for European fauna (Bocherens et al. 1991, 1994, 1995; Bonsall et al. 1997; Murray and Schoeninger 1988). This allows an estimate of the relative contribution of plants and animal proteins in a human diet. Also į15N values tend to be higher in marine and aquatic systems than in terrestrial systems (Katzenberg 1989). This is because these food chains are longer, with higher-order carnivores. Humans who consume significant amounts of marine fish and/or mammals can have į15N values of 15‰ or higher (Richards 2000a, forthcoming; Schulting and Richards 2000).

From the Scottish west coast, three ‘Mesolithic’ human samples from the Cnoc Coig shell midden on Oronsay (Argyll and Bute), radiocarbon dated to the seventh millennium BP,14 indicated a human diet that was largely marine based, with similar values being obtained on a sea otter from Cnoc Coig (Argyll and Bute). A kilometre south at the ‘Mesolithic’ shell midden of Caisteal nan Gillean II (Argyll and Bute), a human bone radiocarbon dated to the second half of the seventh millennium BP15 indicates a mixed marine/terrestrial human diet (Richards and Mellars 1998; Schulting 1998a:208; Schulting and Richards 2000). At Aveline’s Hole (Somerset) three human bones, radiocarbon dated to the late eleventh and the first half of the tenth millennium BP,16 have also been analysed. The results indicate that some marine input may have occurred in the ‘Mesolithic’ human diet, but the majority of protein was from terrestrial sources. At Worm’s Head Cave (Swansea) a į13C value of -18.8‰, obtained from a human bone radiocarbon dated to the late eleventh and the first half of the tenth millennium BP,17 is indicative of some (10% or less) marine foods in the diet, but most of the dietary protein came from terrestrial 14. The three calibrated radiocarbon dates are: 6720-6700 BP (.022), 6670-6400 BP (.974) and 6370-6360 BP (.004) (5740±65 bp, OxA-8004); 6490-6460 BP (.069) and 6450-6300 BP (.931) (5615±45 bp, OxA-8019); 6410-6180 BP (1.000) (5495±55 bp, OxA-8014). 15. The calibrated radiocarbon date is: 6400-6370 BP (.102), 6360-6170 BP (.883); 6130-6120 BP (.015) (5480±55 bp, OxA-8005). 16. The three calibrated radiocarbon dates are: 1055010530 BP (.010), 10510-10110 BP (.892) and 100809920 BP (.099) (9100±100 bp, OxA-799); 10210-9660 BP (.979), 9650-9630 BP (.015) and 9610-9610 BP (.006) (8860±100 bp, OxA-800); 10150-10050 BP (.098), 10040-9990 BP (.045) and 9960-9540 BP (.857) (8740±100 bp, OxA-1070). 17. The calibrated radiocarbon date is: 10150-9600 BP (.991) and 9570-9570 BP (.009) (8800±80 bp, OxA4024).

It is important to remember also that while freshwater organisms, excluding salmon, typically have ‘terrestrial’ carbon isotope signatures, the ‘hard water’ effect of early Holocene lakes, such as those in the Vale of Pickering, could produce į13C values that mimic a marine signature and also have elevated values (compared to a purely landbased terrestrial diet). However these will generally not be as high as those found in marine systems due to the longer food chains in marine environments. It is important to note that other components of the diet, such as carbohydrates and lipids, are not reflected in the measurements on bone collagen, only the protein intake (Schulting and Richards 2000). Richards and Schulting, amongst others, have collected į13C and į15N values, along with radiocarbon dates of human skeletal material from a number of contexts across Britain and Ireland. Samples from both ‘Mesolithic’ and ‘Neolithic’ contexts have been chosen to identify any dietary or subsistence changes. This evidence will be presented first as it offers a detailed picture of dietary patterns. Following this, radiocarbon dated samples that have measured į13C levels, as part of the dating process will be discussed. These measurements are not as accurate as those studies specifically for palaeodietary 36

millennium BP,21 and Rockmarshall (Co. Louth), with a human bone sample radiocarbon dated to the seventh millennium BP,22 show a significant reliance on marine protein in the human diet. From the inland ‘Mesolithic’ context of Killuragh (Co. Limerick) there is no marine element in the human diet indicated from two human bone samples radiocarbon dated to the seventh millennium BP23 (Woodman et al. 1999:143-4). In Wales at Paviland Cave (Goat’s Hole and Fox Hole) (Swansea) a number of samples that span the ‘Mesolithic’ and ‘Neolithic’ have been dated and analysed. A ‘Mesolithic’ human bone sample has been radiocarbon dated to the first half of the eighth millennium BP.24 The three remaining ‘Neolithic’ samples have been radiocarbon dated to the sixth millennium BP.25 The į13C value indicates that all individuals lacked marine protein in their diet despite the close proximity to the sea. The į15N value of the three ‘Neolithic’ samples indicates high levels of animal protein in the diet. The į15N value for the ‘Mesolithic’ is either related to high levels of animal proteins in the diet or childhood breast-feeding (Richards 2000b; Schulting and Richards 2002c).

sources (Richards and Schulting 2000; Schulting and Richards 2002c). A surprising result came from a sample radiocarbon dated to the tenth millennium BP18 from Oreston Third Bone Cave (Breakwater Quarry) (Devon), which despite its close proximity to the coast, produced a į13C value of 20.2‰ indicating that all dietary protein came from terrestrial sources. At the cave Ogof-yr-Ychen, 5 human samples, one from the ‘Mesolithic’ radiocarbon dated to 8020-7660 BP (1.000) (7020±100 bp, OxA-2574), indicated considerable use of marine protein, but terrestrial foods were a significant part of the diet (Richards and Schulting 2000; Schulting 1998a:210; Schulting and Richards 2002c). From two other caves with ‘Mesolithic’ human skeletal material, Potter’s Cave (Pembrokeshire), with two human bone samples radiocarbon dated to the tenth and the first half of the ninth millennium BP,19 and Small Ord Point (Daylight Rock Cave) (Pembrokeshire), with a human bone sample radiocarbon dated to the first half of the tenth millennium BP,20 two groups are clearly present. The sample from Small Ord Point (Daylight Rock Cave) (Pembrokeshire) exhibits an entirely terrestrial diet, while the two samples from Potter’s Cave (Pembrokeshire) indicates a use of approximately one-third or more marine-derived protein (Schulting 1998a:210-1; Schulting and Richards 2002c). Recent stable isotope analysis of a human femur, dated to 7681-7574 BP (1.000) (6790±40 bp, Beta-144016), recovered from a palaeochannel at Staythorpe (Nottinghamshire), provided a į13C value of -20.4‰ indicating a diet where all of the protein came from terrestrial sources. The į15N value of 9.3 indicates a diet very high in animal protein (Myers 2001).

The largest number of samples has been taken from ‘Neolithic’ human skeletal material. Four human bone samples radiocarbon dated to the sixth and the first half

21. The two calibrated radiocarbon dates are: 6490-6280 BP (1.000), (5590±60 bp, OxA-5770); 6470-6460 BP (.007), 6450-6270 BP (.902) and 6250-6200 BP (.091) (5545±65 bp, OxA-4918). 22. The calibrated radiocarbon date is: 6710-6710 BP (.001), 6660-6380 BP (.911) and 6370-6310 BP (.088) (5705±75 bp, OxA-4604). 23. The two calibrated radiocarbon dates are: 6635-6581 BP (.164), 6571-6438 BP (.798) and 6425-6411 BP (.039) (5725±35 bp, OxA-6752); 6396-6369 BP (.028), 6347-6330 BP (.014), 6317-6170 BP (.905), 6143-6111 BP (.045), 6095-6093 BP (.001), 6067-6061 BP (.005) and 6008-6003 BP (.003) (5455±50 bp, OxA-6749). 24. The radiocarbon date is: 7717-7710 BP (.007), 76957695 BP (.001), 7694-7567 BP (.969) and 7528-7512 BP (.023) (6785±50 bp, OxA-8316). 25. The three calibrated radiocarbon dates are: 5838-5832 BP (.010) and 5746-5593 BP (.990) (4940±45 bp, OxA8315); 5656-5569 BP (.582) and 5554-5471 BP (.418) (4840±45 bp, OxA-8318); 5564-5563 BP (.001), 54695288 BP (.978), 5156-5143 BP (.010) and 5099-5088 BP (.011) (4625±40 bp, OxA-8317).

From Ireland, the ‘Mesolithic’ shell middens at Ferriter’s Cove (Co. Kerry), with two human bone samples radiocarbon dated to the second half of the seventh 18. The calibrated radiocarbon date is: 9890-9880 BP (.006), 9870-9850 BP (.014), 9820-9810 BP (.013) and 9780-9470 BP (.967) (8615±75 bp, OxA-4777). 19. The two calibrated radiocarbon dates are: 9700-9470 BP (.998) and 9450-9440 BP (.002) (8580±60 bp, OxA7688); 8980-8910 BP (.126), 8900-8880 BP (.045), 88708830 BP (.087), 8810-8590 BP (.702) and 8580-8540 BP (.041) (7880±55 bp, OxA-7687). 20. The calibrated radiocarbon date is: 9890-9880 BP (.007), 9870-9850 BP (.018), 9820-9810 BP (.016) and 9780-9530 BP (.959) (8655±60 bp, OxA-7686).

37

of the fifth millennium BP26 at the Carding Mill Bay (Argyll and Bute) shell midden showed dietary lack of marine protein and a heavy reliance on animal protein in the human diet. A similar pattern has been identified from analysis of a sample radiocarbon dated to the sixth millennium BP27 at the chambered structure at Crarae (Argyll and Bute), a sample radiocarbon dated to the fifth millennium BP28 from Newport Dock NP-1 (Alexander Docks) (Newport) and two samples from Nanna’s Cave (Pembrokeshire) radiocarbon dated to the second half of the sixth and the first half of the fifth millennium BP29 (Bell et al. 2000; Schulting 1998a:209; Schulting and Richards 2000; 2002c). A human bone sample from Ogof-yr-Benglog (Pembrokeshire) dated to 5575-5536 BP (.071) and 5474-5303 BP (.929) (4660±45 bp, OxA7743) indicates a diet with no use of marine resources, with protein coming from both plant and animal resources (Schulting and Richards 2000, 2002b, 2002c). In addition, stable sulphur measurements (į34S) were obtained from two human bone samples (not radiocarbon dated) from Carding Mill Bay (Argyll and Bute). As stated above, the į13C and į15N values of these samples show no marine component to the diet, yet the į34S values of 19.7 and 19.8‰ indicate a strong marine influence. Schulting and Richards (2002b:155-7) conclude that “these individuals were consuming terrestrial foods – whether plant and/or animal – originating in coastal sea-spray zones, strongly implying that they did live on, or very near, the coast during their lifetimes but chose to avoid marine foods”.

From Hambledon Hill Complex (Dorset), stable isotope analysis has been carried out on 56 samples of human remains, 13 of which have been calibrated from 18 radiocarbon dates to the sixth and the first half of the fifth millennium BP30 (Richards and Hedges 1999a. The į13C values indicate no marine protein in the diets, as would be expected on this inland context. The į15N values are more varied, indicating that some individuals obtained almost all dietary protein from animal protein, while some others had a mixed plant and animal protein diet (Richards 2000a, forthcoming). At the chambered structure at Parc le Breos Cwm (Swansea), ten samples of human bone have been radiocarbon dated to the sixth, fifth and into

30. The eighteen calibrated radiocarbon dates are: 57085687 BP (.036), 5682-5672 BP (.016) and 5661-5587 BP (.948) (4890±35 bp, OxA-7102); 5705-5694 BP (.010), 5675-5674 BP (.001), 5661-5571 BP (.726) and 55485472 BP (.263) (4855±45 bp, OxA-7835); 5609-5567 BP (.329) and 5558-5471 BP (.671) (4815±35 bp, OxA7101); 5645-5628 BP (.021), 5614-5465 BP (.951) and 5357-5332 BP (.029) (4810±45 bp, OxA-7768); 56355634 BP (.001), 5611-5451 BP (.907) and 5379-5329 BP (.093) (4795±50 bp, OxA-7769, replicate of OxA-7768); 5591-5464 BP (.980) and 5351-5337 BP (.020) (4770±30 bp, OxA-7100); 5594-5449 BP (.841) and 5382-5327 BP (.159) (4765±45 bp, OxA-7773); 5584-5501 BP (.544), 5491-5451 BP (.226) and 5379-5329 BP (.230) (4738±28 bp, UB-4242); 5584-5501 BP (.394), 5491-5446 BP (.218) and 5413-5324 BP (.388) (4725±40 bp, OxA-7774, replicate of OxA-7773); 5583-5504 BP (.322), 5489-5442 BP (.218) and 5418-5322 BP (.460) (4715±40 bp, OxA7818, replicate of UB-4311); 5576-5532 BP (.224), 54835447 BP (.218) and 5410-5325 BP (.558) (4710±23 bp, UB-4311); 5578-5529 BP (.174), 5515-5513 BP (.002) and 5486-5318 BP (.824) (4695±40 bp, OxA-7836); 5570-5552 BP (.047) and 5471-5317 BP (.953) (4680±30 bp, OxA-7814); 5567-5557 BP (.027), 5470-5426 BP (.237) and 5424-5318 BP (.737) (4679±27 bp, UB-4243); 5580-5510 BP (.101), 5490-5280 BP (.850), 5160-5130 BP (.024) and 5100-5070 BP (.025) (4645±60 bp, OxA7045, replicate of OxA-7044); 5460-5370 BP (.131), 5330-5040 BP (.856) and 5010-4990 BP (.013) (4565±60 bp, OxA-7040, replicate of OxA-7039); 5450-5380 BP (.092), 5330-5040 BP (.902) and 5000-5000 BP (.006) (4560±55 bp, OxA-7044); 5450-5400 BP (.060), 53305030 BP (.908) and 5010-4980 BP (.032) (4550±60 bp, OxA-7039).

26. The four calibrated radiocarbon dates are: 5653-5566 BP (.482) and 5559-5469 BP (.518) (4830±45 bp, OxA7664); 5639-5631 BP (.008), 5612-5452 BP (.913) and 5377-5330 BP (.079) (4800±50 bp, OxA-7663); 55775531 BP (.149) and 5485-5317 BP (.851) (4690±40 bp, OxA-7665); 5260-5250 BP (.003), 5230-5220 BP (.001), 5210-5190 BP (.015), 5120-5110 BP (.002), 5050-4810 BP (.955), 4760-4730 BP (.015), 4720-4710 BP (.004) and 4670-4660 BP (.004) (4330±60 bp, OxA-7890). 27. The calibrated radiocarbon date is: 5586-5497 BP (.467), 5493-5448 BP (.220) and 5407-5326 BP (.313) (4735±40 bp, OxA-7662). 28. The calibrated radiocarbon date is: 4605-4603 BP (.002), 4570-4556 BP (.027), 4551-4350 BP (.941) and 4327-4299 BP (.029) (3995±45 bp, OxA-7656). 29. The two calibrated radiocarbon dates are: 5445-5414 BP (.058), 5324-5205 BP (.386) and 5195-5047 BP (.556) (4560±45 bp, OxA-7739); 5312-5038 BP (.984) and 5007-4992 BP (.016) (4520±45 bp, OxA-7740).

38

the early fourth millennium BP,31 indicating a predominantly terrestrial source for dietary protein, drawn mainly from animal protein (Richards 1998, 2000a:132; Schulting and Richards 2002c). Only į13C values were taken from an additional two human bone samples radiocarbon dated to the sixth millennium BP,32 and they exhibited a terrestrial diet (Richards and Hedges 1999a).

millennium BP,34 the į13C values indicated that the individual’s diet contained no marine protein and į15N values suggested that most individuals had a mixed plant and animal protein diet (Richards 2000a). Only į13C values were taken from an additional three human bone samples from Hazleton North, radiocarbon dated to the sixth millennium BP,35 indicating a similar terrestrial diet (Richards and Hedges 1999a).

At other chambered structures at Hazleton North (Gloucestershire), with five human bone samples radiocarbon dated to the late seventh and sixth millennium BP,33 and West Kennet (Wiltshire), with three human bone samples radiocarbon dated to the sixth

The isotope analysis results from the chambered structures at West Kennet (Wiltshire), Parc le Breos Cwm (Swansea) and Hazelton North (Gloucestershire) indicates to Richards that the people in the chambered structures represent a single group as they had similar isotope values and thus similar lifetime diets. This is different to the diets represented in the causewayed enclosure at Hambledon Hill Complex (Dorset), where a much wider range of į15N values indicates a more varied diet among the people at this location (Richards 2000a:133).

31. The ten calibrated radiocarbon dates are: 5730-5460 BP (.974) and 5360-5330 BP (.026) (4850±65 bp, OxA6496); 5650-5630 BP (.032), 5620-5450 BP (.885) and 5380-5330 BP (.084) (4805±55 bp, OxA-6492); 56105450 BP (.801) and 5410-5330 BP (.199) (4780±60 bp, OxA-6488); 5580-5500 BP (.309) and 5490-5330 BP (.691) (4710±60 bp, OxA-6491); 5590-5500 BP (.236) and 5490-5300 BP (.764) (4685±65 bp, OxA-6487); 5580-5500 BP (.147), 5490-5290 BP (.842), 5160-5150 BP (.006) and 5100-5090 BP (.006) (4660±60 bp, OxA6490); 5580-5510 BP (.105), 5490-5280 BP (.843), 51605130 BP (.024) and 5110-5070 BP (.028) (4645±60 bp, OxA-6494); 5290-5160 BP (.348), 5150-5100 BP (.105) and 5090-4870 BP (.548) (4445±60 bp, OxA-6489); 4340-4340 BP (.002), 4290-4270 BP (.012), 4260-3960 BP (.963) and 3950-3930 BP (.024) (3750±55 bp, OxA6497); 4230-4200 BP (.037), 4180-4170 BP (.013), 41603890 BP (.950) and 3870-3870 BP (.001) (3705±55 bp, OxA-6595). 32. The two calibrated radiocarbon dates are: 5580-5510 BP (.216) and 5490-5310 BP (.783) (4690±55 bp, OxA6641); 5730-5570 BP (.847) and 5540-5470 BP (.153) (4875±55 bp, OxA-6493). 33. The five calibrated radiocarbon dates are: 6280-5660 BP (1.000) (5200±150 bp, OxA-912); 5900-5610 BP (1.000) (5000±70 bp, OxA-910); 5890-5780 BP (.188) and 5770-5590 BP (.812) (4950±70 bp, OxA-905); 57205450 BP (.882) and 5410-5330 BP (.118) (4830±80 bp, OxA-911); 5590-5500 BP (.132), 5490-5260 BP (.698), 5250-5230 BP (.012), 5220-5210 BP (.004), 5190-5120 BP (.081) and 5110-5050 BP (.074) (4640±80 bp, OxA1177).

Richards and Hedges (1999) collected 78 human skeletal material samples that have been radiocarbon dated and had į13C values measured from 27 coastal and inland contexts in England and Wales (see Richards and Hedges 1999a:Table 1). The results showed that no individual dating after 5400 BP (middle of the sixth millennium BP) had any marine protein in their diets, with all į13C values close to -20‰ indicating a terrestrial food diet. This consistent pattern was found at inland and coastal contexts. While admitting to gaps in their record, Richards and Hedges (ibid.:893) conclude: that, at least in coastal areas, there was a rapid change in diet in southern Britain associated with the introduction of Neolithic material culture and the first appearance of domesticated plants and animals. 34. The three calibrated radiocarbon dates are: 5710-5700 BP (.004) and 5660-5310 BP (.996) (4780±90 bp, OxA563); 5710-5700 BP (.004) and 5660-5310 BP (.996) (4780±90 bp, OxA-451); 5600-5290 BP (.989), 51605140 BP (.006) and 5100-5090 BP (.005) (4700±80 bp, OxA-450). 35. The three calibrated radiocarbon dates are: 5860-5830 BP (.019), 5750-5470 BP (.977) and 5340-5330 BP (.004) (4880±70 bp, OxA-906); 5740-5460 BP (.974), 5370-5370 BP (.001) and 5360-5330 BP (.026) (4860±70 bp, OxA-904); 5710-5670 BP (.042), 5660-5460 BP (.928) and 5360-5330 BP (.029) (4840±60 bp, OxA-903).

39

Figure 2.14 – The ‘Mesolithic’/‘Neolithic’ transition (after Thomas 1999a: figure 2.1). (1999:40) states: “crops and cattle can be viewed as components of the traditional resource base of advanced foraging strategies” (also see Simmons and Innes 1987:399). This involved the continuing use of wild resources with livestock management, and to a lesser extent cereals, contributing to a long-established subsistence strategy. ‘Neolithic’ populations were managing a diverse subsistence strategy made up of traditional resources, with new resources developed out of the old relationships with plants and animals (Figure 2.14). Elsewhere in archaeological contexts, diverse subsistence strategy has been used to argue for mobility strategies (Bonzani 1997, see Chapter 6). There is no a priori reason to suggest that the presence of domestic plants and animals in contexts dated for the middle of the seventh to the late fifth millennium BP required a sedentary mode of occupation. Rather, the continual utilisation of wild resources and a similar limited clearing of woodland for much of Britain and Ireland, which is identified as evidence for mobility in the ‘Mesolithic’ and earlier, would suggest that a similar mobile mode of occupation, which occurred in earlier millennia, continued into the second half of the seventh to the fifth millennium BP (see Thomas 1999a:29 and Chapter 6).

A similar pattern can be identified at two ‘Mesolithic’ contexts; from Oreston Third Bone Cave (Breakwater Quarry) (Devon) and Fox Hole Cave (Swansea). As discussed earlier, the samples from these contexts are interesting, because despite their coastal location all samples show no protein from marine resources, with all dietary protein coming from terrestrial sources. Richards and Hedges’ work (also see Schulting and Richards 2002b) has identified a dietary switch from marine to terrestrial protein. Their conclusion states that the terrestrial component was wholly derived from domesticated plants and animals. However, the terrestrial component could have equally been derived from wild plants and animals, or any combination of wild and domestic terrestrial plants (also see Thomas 2003:69). While the above evidence has shown that domesticated animals, and to some degree domesticated plants, were present in the ‘Neolithic’ to varying degrees, it has also illustrated that they were just incorporated into a subsistence strategy that included the continuing utilisation of wild plants and animals.

Conclusions

Although evidence from stable isotopes indicated a switch from marine to terrestrial diets at the start of the ‘Neolithic’, there is no evidence to what extent this included only domesticated plants and animals. This pattern could have been created through a switch to wild plants and animals that were part of the subsistence strategy in the ‘Neolithic’. The tendency is for archaeologists to accept that clearances were the result of human action, cereal pollen and animal bones of a

The subsistence activities of the ‘Neolithic’ included new cereal and livestock domestication but also existing methods of subsistence dating back from earlier millennia. When the evidence is drawn together, ‘Neolithic’ populations were continuing a long and wellestablished management of plant and animals known from ‘Upper Palaeolithic’ assemblages. As Richmond 40

‘Neolithic’ cultural and phenomenological baggage. An open landscape dominated by fields, penned livestock, and dotted with ‘ritual’ and ‘funerary’ monuments, is the ‘eternal’ image of the ‘Neolithic’ lifestyle. However, we now see that the landscape was anything but open. Instead it consisted of small clearings used to attract wild fauna, to herd livestock and to construct stone, timber and earthen structures. An agricultural economy has been considered crucial to explaining the monumental stone, timber and earthen components of the landscape. It has long been argued that an agricultural surplus was required to allow these construction projects (Legge 1989:220), as if hunter-gatherer populations could not create substantial constructions (however see page 55). Following this reappraisal of the woodland cover and the subsistence strategy of the ‘Neolithic’, a reappraisal of the many stone, timber and earthen constructions found in the ‘Neolithic’ will be made in Chapter 4, arguing from continuity from the preceding millennia.

domesticated size as domesticates, which occur within an acceptable chronological framework (that is after about 6000 BP) but not before (see Woodman 2000:235). The belief is that ‘Neolithic’ populations manipulated the landscape for agricultural reasons and that any evidence that supports this long held view is to be accepted. The development and adoption of agriculture attributed to the ‘Neolithic’ since the time of Childe and before has systematically marginalised the contribution of ‘Mesolithic’ subsistence practices. The traditional view of an agricultural landscape being formed by new ‘Neolithic’ populations can no longer be substantiated on the empirical evidence. There is no evidence that the ‘Mesolithic’/‘Neolithic’ transition involved new populations in Britain and Ireland (Lahr et al. 2000), which has been at the heart of the ‘wave of advance’ model. Despite this, ‘Mesolithic’ populations are identified as being converted or assimilated by new

41

PART II CHAPTER THREE: IDENTIFYING UNPARALLELED BEHAVIOUR plausible and familiar as they represent human social behaviour which we can accept and understand, but have “little to do with their connection to evidence and much more to do with our expectations of what an intelligible human past should be like” (Murray 2001:29).

If…archaeological data are a special kind of data about human actions, a reassessment of the discipline which recognizes their uniqueness should also stress their importance as the spur for a concomitant reassessment of social and cultural theory – the very terms within which we seek knowledge of human nature. (Murray 1987:9)

The history of archaeology and standard social theory

This chapter will outline the theoretical and interpretive frameworks used in the ‘Neolithic’ discourse and the wider discipline of archaeology. These frameworks limit our ability to identify unparalleled forms of human social behaviour in the past. This will serve as the theoretical basis for my reappraisal of various aspects of ‘Mesolithic’ and ‘Neolithic’ human social behaviour in Britain and Ireland which began in the last chapter.

For it is we as archaeologists who reconstruct (or some would say construct) the past. In this we cannot avoid doing so from the standpoint of the present, and we have the responsibility of interpreting that past for our contemporaries in the present. (Renfrew 1994a:153)

That archaeological interpretations should be constrained by the evidence may seem self-evident, the evidence being the material left behind by past humans. However, since the development of archaeology as a discipline in the nineteenth century the generalised perception has been that if only the material remains were used for interpretations, knowledge produced by the discipline would be reduced to the trivial and insignificant (Murray 2001:29). In 1979 De Boer and Lathrap commented: Either they [archaeologists] must become practitioners of an overextended uniformitarianism in which past cultural behaviour is ‘read’ from our knowledge of present cultural behaviour or they must eschew their commitment to understand behaviour altogether and engage in a kind of ‘artefact physics’ in which the form and distribution of behavioural by-products are measured in a behavioural vacuum. (1979:103)

Archaeology is unique among social and historical inquiries, because it seeks to interpret human ‘active’ meaning without any direct contact with it (Trigger 1998:1). The use of standard social theory from social anthropology/sociology has a long tradition in archaeology. This relates to a deep and enduring relationship between archaeology and ethnography/sociocultural anthropology, which “is one of the most important foundations of the disciplinary culture of archaeology” (Murray 1992:730). This relationship is intimately linked to the initial development of archaeology. Those studying past societies envisaged that the worldwide findings of ethnographers could help them understand pre-industrial populations. This relationship between archaeology and social theory continues today, but is far from a ‘natural’ one – rather it is a product of the history of scholarship and the social world of nineteenth and twentieth century European culture (Lucas 2001: 188; Murray 2001).

It would seem we can either identify significant results based on questionable methods or trivial results based on sound methods. The interpretations produced are plausible and familiar as they derive from specific social meaning known from the ethnographic and anthropological literature. The assumption remains that the past should be immediately understandable in relation to the present. As has already been outlined, the former is generally the preferred choice amongst the current scholars of the ‘Neolithic’ in Britain and Ireland. At present, the ‘Neolithic’ discourse aims to produce expressions about active human social behaviour and past human interaction with artefacts, landscape and each other which are constructed in terms of conventional social behavioural expressions of social analysis, drawing substantive analogies from observed and documented human social behaviour. The interpretations produced are

Archaeological assemblages have long been seen as ‘losing’ information about social meaning and behaviour once they depart the systemic context and enter the archaeological context – that is, the verbal and active component of sociality is absent (see Schiffer 1972, 1976). The most common response to recover this ‘lost’ component since the nineteenth century has been through “structured comparison between the archaeological record and ethnographic observation – the structure being provided by what passes for social theory at any point in time” (Murray 1997:456, see Lubbock 1869, 1882; Tylor 1865, 1870, 1893). Murray (1997:456) states that this practice: ensures that the prehistoric past is not beyond the bounds of our understanding, a possibility which would severely damage the creditability of the structures through which we make sense of 42

a history to the contemporary ‘savages’ and ‘barbarians’ (see Golson 1977; Gosden 1992; Janik 1999; Lucas 2001: 187-92; Murray 2001). Their unparalleled forms of human social behaviour, which were their history, were collapsed into explaining the prehistoric past through the use of the contemporary structures of knowing about human social behaviour. To keep these structures as universals, the ‘savage’ and ‘barbarian’ cultures were presented as the modern remnants of prehistoric human socio-cultural types. Moreover, this practice made the vast time-scales, which archaeology was investigating, disappear “for the very reason that it was undimensionable by any of the ethnological or anthropological theories then available” (Murray 2001:36; also see Lucas 2001: 187-92; Orme 1974), a state of affairs which continues today.

contemporary experience. Murray states that between the years of 1858 and 1870 archaeology was clearly linked to palaeontology and geology, which allowed archaeological practitioners to claim a scientific reliability for their constructions of human social behaviour in the prehistoric past. With the discovery of high human antiquity in the nineteenth century, a crucial identification for the foundation of archaeology, the threat existed that human social behaviour in deep prehistory might be unintelligible in terms of the social perspectives of the time (Murray 1992, 1997, 2001). By using the established frameworks of ethnology and history, the potential unintelligible record of behaviour was ‘normalised’ and ‘humanised’. This allowed both archaeology and the other social sciences to maintain their legitimacy as producers of meaningful knowledge about human beings using an a priori assumption that archaeological representations of human social behaviour should parallel those of contemporary social interpretation. As Murray (2001:43) has noted: Any history of archaeology should also be a history of the retention of concepts which owe their plausibility to the cultural traditions of the discipline rather than to their usefulness in more directly engaging the archaeological past. It is the appropriateness of these tools that should be explored, which I will seek to do. This has been an alltoo-rare undertaking in the ‘Neolithic’ discourse and the wider archaeological discipline. The result has been that the observations and declarations drawn from anthropological literature have become accepted facts, which underpin the whole discipline. I will illustrate the problems inherent in asking particular questions of archaeological assemblages – questions that cannot be answered without the use of secondary referents drawn from anthropological sources. I will argue that these questions should not be asked. Rather we should develop questions that are particularly sensitive to the nature of the archaeological assemblages and contexts we encounter, leading to different or alternative ways of seeing and referring to human sociality that recognises operational behaviour as something in its own right.

However, even in the nineteenth century, advocates of this ‘commonsense’ view identified two worrying potential consequences of this practice. First – that archaeological assemblages and contexts might lose their power to vigorously critique contemporary representations of ‘humanness’. Second – that the lack of significance attributed to archaeological assemblages and contexts may diminish the reliability of assertions of archaeological knowledge. These fears were expressed in Hawkes’ influential ladder of inference. Since Hawkes’ 1954 paper, thousands of pages have been devoted to archaeological theory and interpretive frameworks. Theoretical trends have included processual, postprocessual and poststructural approaches, all identifying different epistemological grounds for interpreting archaeological assemblages and contexts. However, for all their rhetoric of distinctive approaches they “generally still adhere to the notion that meaningful explanation or interpretation of prehistoric human behaviour should mirror the forms found in contemporary social theory” (Murray 1997:456, also see Fletcher 1989). Despite much debate in archaeological theory over the last 150 years, the fundamental obstacles have only recently been confronted – the standard assumptions about the proper tools to be employed for archaeological interpretation (see Fletcher 1995:11-3; King 2001; Murray 1997). The discourse still operates on the notion that the present and the documented past hold the key to understanding human social behaviour in the prehistoric past. But the assumption remains that the past should be immediately understandable in relation to the present. Murray argues that the unintelligibility of the past has rarely been raised since the 1860s. Forms of interpretation have received attention and issues of ambiguity and polyvocality have been raised but, the idea that the prehistoric past poses a threat to the universality of the structures through which we interpret contemporary experience (social theory as well as the very identity of ‘human behaviour’) has received very little support (but see e.g. Dunnell 1982). (Murray 1993:177)

Archaeology, using the established frameworks of ethnology and history, allowed ethnology and later, anthropology, to expand its coverage back to the emergence of human social behaviour (Murray 1987:Chapter 4, 2001; van Riper 1993). Thus, new archaeological data were set within a framework of sociocultural evolution developed in the eighteenth century that assumed progress was inherent to human existence. Socio-cultural changes were explained through ethnographic analogy, with peculiarities or anomalous forms of human social behaviour being rationalised as the “product of a complex process which included elements of stasis, regression and progress” (Murray 1997:458). There was an a priori assumption that archaeological representations of human social behaviour should parallel those of contemporary social theory. This practice not only ‘normalised’ unparalleled forms of human social behaviour in the prehistoric past, making it effectively synchronous with the present, but also effectively denied

The prospect that human social behaviour existed in the past without modern reference exposed the possibility of 43

affected both through the use of explicit ethnographic examples and through generalised universal arguments. Yet the practitioners of postprocessual archaeology have long argued for the inherent contextual uniqueness of archaeological assemblages and contexts. Their aim to identify and document the contextual uniqueness of the past is impaired by their failure to address the epistemological history of archaeology and by their continual employment of standard social theory – those who ignore history are doomed to repeat it.

an unintelligible past. Rather than formulating a different kind of social theory to explain this behaviour, the discourse in the nineteenth century used conventional social behavioural terms of social analysis, drawing substantive analogies from observed and documented human social behaviour. The discovery of the time depths involved in human prehistory was a threat to science, “a threat which was defused by effectively turning deep past (oblivion) into ethnographic present (living representations of stages of European race memory). The prehistoric past became human, and properly the subject of human understanding” (ibid.:183). However, the price has been very high. The retention of conventional meanings and the need to humanise the past has produced an ontological singularity of human social behaviour from prehistory to the present. While much has been written on the wrongs of colonial, racist, etc. abuses of the past, the need for intelligibility as past human social behaviour overcomes all (Murray 1992, 1993).

Ironically, this echoes a return to the eighteenth and nineteenth century where European scholars believed that they were seeing their own past ‘frozen in time’ in the behaviour and lifestyle of the newly discovered populations of Australia, Africa and the Americas (see Lubbock 1869; Pownall 1795; Wilson 1851, see Lucas 2001: 187-92; Murray 2001; Spikins 1999:Chapter 4 for commentary). Such a view is colonialist, and ultimately racist, and entrenches notions of progress from ‘primitive’ to ‘civilised’ society. In essence, the work of Lubbock and others provided the scientific credentials and justification for colonialism (Trigger 1995:321, 1996). Furthermore, there is no reason to think that these populations should be a reflection of those of prehistoric Britain and Ireland. Humans are not so one dimensional today, let alone populations separated by thousands of years. As Hodder (1982a:39) has pointed out: there is a possibility that, after having overrun and pillaged Australian, African and American peoples militarily, economically and socially, we follow this up by an ‘intellectual colonialism’. The failing of the current ‘radical’ discourse in prehistoric Britain and Ireland is that it has been unprepared to question the particular interpretive tools employed.

The current use of standard social theory in archaeology But the past is not just a foreign country, it speaks a foreign language which we mistake for our own. (Thomas 1993b:389, also see Whittle 1997b:143). The current archaeological practice of social explanation involves the creation of a version of the verbalised meanings that are supposed to have been associated with past human social behaviour (Fletcher 1989). While the methods have been modified, standard social theory and ethnographic analogy are still seen as useful interpretive tools for identifying human social behaviour through prehistoric archaeological assemblages and contexts. This usually comes from ethnographic and anthropological literature (for example, Barley 1981; Bloch 1971, 1982; Bloch and Parry 1982; Douglas 1966; Hertz 1960; Humphreys 1981; Huntingdon and Metcalf 1979; Metcalf 1981; Munn 1986; Turner 1969; van Gennep 1960) and is used throughout the discourse (for example, Cooney 2000a; Thomas 1991a, 1999a; Tilley 1994, 1996a). Tilley for example uses ethnographic evidence to “normalize” his interpretation (Tilley 1996a:6). He claims to use ethnographic evidence as a medium through which to view the archaeological evidence, rather than a model to which the data are to be fitted or against which the data are to be tested (ibid.:2).

The use of social analysis makes the identification of past and different human social behaviour problematic. There is no reason to expect that the range of human social behaviour that is documented ethnographically and ethno-historically is in any way sufficiently comparable to, or representative of, the behaviour of prehistoric populations. The very sporadic, fragmented and selective way in which these observations are collected and documented should automatically deem them inappropriate for the interpretation of human social behaviour identified in prehistoric archaeology. We cannot presume that past human social behaviour will be the same as behaviour which has been observed and documented. This is because other human social behaviour which existed in the past need not be represented in the present. If this is the case, archaeology is in the position to identify this human social behaviour and must seek to do so because it is the only discipline that can do so.

While this work, and much of the phenomenological studies on the ‘Neolithic’, have been viewed as ‘radical’ and as a serious attempt to move way from explaining the past as a reflection of our present social and economic systems, their ‘escape’ is a superficial one. While our present social and economic systems have purportedly been replaced with those of the historically documented and observed ethnography of ‘primitive’ societies in actuality the past is still defined and interpreted by direct analogy with present and recent behaviour. This has been

It is also important to note that ethnographic evidence is not without its own problems. Scholars have their own biases, assumptions and preoccupations that cloud the documentation of an independent social reality (Hammersley 1992; Rohatynskyj and Jaarsma 2000a, 2000b). As strong relativists argue there is no empirical 44

more difficult to interpret than data in the present. However, we need to look at different data to answer different questions. Of course the past is more difficult to interpret if we feel obliged to seek answers to questions defined in the present which require the data that are available in the present. Therefore we need to consider the questions which we can ask of the data we have at our disposal.

reason to identify one interpretation as right and another wrong – what is constructed is simply a version of reality. Thus there are as many ethnographic and archaeological interpretations as persons, to paraphrase J.K. Smith’s: “there are as many realities as there are persons” (1984:386). The biases, assumptions and preoccupations of ethnographic research will impact and reduce the potential identification ranges of behaviour (Hammersley 1992; Ingold 1996). As Fabian (1999:85) comments: “Ethnographers report what they understand”.

First, it is crucial to identify that there is the general metaphysical proposition exists – what Gould calls methodological uniformitarianism – and that there are two types of analytic uniformitarianism: operational and substantive (Fletcher 1992; Gould 1965, 1987:117-26). Operational uniformitarian propositions are logically valid and define the parameters within which particular phenomena can persist or operate, but do not predefine their particular form (for example see genetics and Natural Selection). The operations of the system cannot vary (these are fixed), but specific, unexpected or unparalleled outcomes may derive from those operations (for example the products of genetic reproduction). Operational uniformitarian propositions can be used to refute existing common assumptions about the nature of human social behaviour (Fletcher 1991). For example, most agriculturalists now may well be sedentary. But by identifying that some farming populations in the present actually have residential patterns which range from sedentary to mobile and many variants in between (see Graham 1994 for mobile farmers), it follows that such variation can be expected in the past. Therefore, the prevalent assumption that farming practices in the past required or caused a sedentary residential pattern can be shown to be logically invalid.

Using analogies from social anthropology/sociology we are not therefore getting closer to possible realities, we are merely replacing one set of biases, assumptions and preoccupations for another. Arguments that such social analyses are used as a starting point for the interpretation of human sociality (for example, Tilley 1996a:1-6; Whittle 1998:853) are also questionable, as they may impede what we can observe. One only has to look at the final interpretation to see the intrinsic part the analogies have played. Tilley (1996a:6) claims “the discussion of specific ethnographic examples in the text is not to produce an interpretation – they were added after an interpretation of the archaeological data – but to ‘normalize’ what might otherwise appear ‘unwarranted’ arguments and interpretation”. However, his archaeological interpretations are inherently locked in with the ethnographic evidence and could not be made solely on the basis of archaeological evidence. As Wylie (1985:81) has commented, “analogy seems to be both indispensable to interpretation and always potentially misleading”. Two theoretical problems will now be discussed; namely the general issue of the confused and contracted way in which the general theory of uniformitarianism is discussed and employed in archaeology. This will then be followed by the specific issue of the vague and widespread identification of ‘ritual’ social behaviour in prehistoric contexts. Ritual will be discussed both as a general issue and more specifically for this research. Ritual will be identified as one of the standard categories employed throughout archaeology which is familiar to use and is assumed to be a fundamental category of human social life. Following this, a theoretical and interpretive framework will be outlined which will be used throughout which permits unparalleled forms of human social behaviour to be archaeologically identified. The issue of Uniformitarianism

Operational uniformitarianism allows us to connect present empirical evidence to equivalents in the past through general processes with unparalleled products. Importantly, operational uniformitarianism specifies how far or in what way any processes develop, and not just geological processes (cf. Hodder 1999:107). Operational uniformitarian propositions define the parameters within which particular phenomena can persist. Simpson (1970:59) calls these principles the inherent properties of the universe. An example of an operational uniformitarian principle used frequently in archaeology is the half-life of 14C (5730 years), the basis for all radiocarbon dating techniques (Fletcher 1992:44; Simpson 1970:80).

Uniformitarianism is more complex than the simple assertion that ‘the present is the key to the past’ (Gould 1965). There is also a widespread confusion on what uniformitarianism means for archaeology. Bailey (1983:174) states that uniformitarianism has come to be identified with the belief that we know more about the present than the past, thus the past should necessarily be explained in terms of processes visible in the present, a perspective which is logically incorrect. As Bailey (1983:176) argues: data from the prehistoric past, whether archaeological or otherwise, are in principle no

By contrast, statements of substantive uniformitarianism are logically invalid and refer to extrapolating some specific aspect of the present into the past (Fletcher 1995:230; Gould 1965:227-8). It is important to note that while substantive propositions are testable, that testability is very particular, and relates only to the testing of whether of not constancy of rates of processes occurs, and not to the use of specific analogies. The latter are not testable and thus cannot be shown to be logically valid. This is the fundamental flaw of social reconstruction which is based on substantive analogy. The substantive proposition assumes, “that the associations between 45

collapses the two distinct forms of uniformitarianism and obscures a myriad of complex issues. Due to the scope and variation in human social behaviour that has been historically documented and observed, particular substantive analogies can be chosen which fit almost any argument a scholar wishes to make.36 Certain interpretations and their underlying analogies become fashionable, such as the ancestor interpretation in ‘Neolithic’ Britain and Ireland (see Chapters 4 and 5), leading to a situation whereby archaeological assemblages are automatically pigeon-holed as the archaeological manifestation of a very specific form of human sociality. As Whitey (2002:119) has neatly summarised: Ancestors were to the 1990s what chiefdoms (Yoffee 1993) were to the 1970s – the explanation of choice for a whole range of archaeological phenomena…Like many recent developments in British prehistory, the universal ancestor has gone from being a suggestion to becoming an orthodoxy without ever having had to suffer the indignity of being treated as a mere hypothesis.

material features and active phenomena (such as social relationships) observable in the present are constant and can be retrodicted into the past” (Fletcher 1992:44, also see Gijn and Zvelebil 1997:4). The core problem is that a particular social practise, for example ancestor worship, need have no specific material correlate now or in the past. What is required is an explanation of why a phenomenon is constant over time and space, not just a declaration that it is. Taking radiocarbon dating, the development of radiocarbon dating in the 1950s and early 1960s assumed that the rate of formation of 14C was constant and therefore that the atmospheric concentration of 14C remained constant through time. Although this assumption was useful in the initial development of radiocarbon dating, it is now known to be false (Fletcher 1992:44; Salmon 1982:80). There has been a general and widespread collapsing of operational and substantive uniformitarianism in archaeology, as if there were only one proposition about analogy between the present and the past. Archaeological writings have generally failed to differentiate between the two forms of uniformitarianism, leading to a situation of either accepting all uniformity principles or none (Hodder 1999:45-9 and 107; Kelley and Hanen 1988:264-5; Trigger 1989:363-4). However, it is logically invalid to dismiss all uniformitarian statements as operational statements are both necessary and logically valid for the study of the past. The social reconstruction practised by many ‘Neolithic’ scholars, and other archaeologists, utilises substantive uniformitarianism, presuming that specific associations between material features and active phenomena observed or documented in the present or recent past can be generalised and projected into the past (Fletcher 1992:44, also see Gijn and Zvelebil 1997:4). For example, it cannot be shown to be logically secure to argue that the association between stone structures and ancestors in Madagascar means that a similar associate between ‘megaliths’ and ancestors occurred in the British and Irish ‘Neolithic’ (see Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Thomas 1988b; Tilley 1996a; Scarre 1994). The procedure relies on a logically invalid principle. Just because these terms and procedures are appropriate in the present in association with societies for which we have ‘verbal’ information, this does not make them valid in the past. What archaeology must do is develop its own theoretical positions and interpretations particularly sensitive to the forms of human social behaviour we encounter in archaeological assemblages and contexts.

Observed and historically documented phenomena retrodicted from the present and the recent past and described in ‘verbalised’ terms cannot be shown to be logically secure. This practice allows the past to be at once reasonable and plausible, as it identifies human behaviour that we know to be possible. However, it relies on a logically invalid principle that: the assumption has to be made about the universal uniformity of relationships between the material and the social, when that is precisely the issue which stands in need of appraisal. (Fletcher 1992:44) The key issue is that there is no reason to expect that the particulars of past human social behaviour will resemble those of the present. Implementing such logic produces the possibility that inappropriate associations and observations will be imposed on the past. Thus, the prospect exists that unparalleled forms with no observed or historically documented substantive analogies will remain unidentified because they will be ‘re-read’ as a known phenomenon. Archaeological interpretations tend to uniformitarianism without acknowledging it.

apply It is

36. For example see Fleming’s (1999) critique of Tilley’s 1994 book A Phenomenology of Landscape: Places, Paths and Monuments.

Whittle has recently argued that analogy is an important tool for scholars in interpreting prehistoric human social behaviour. He states that archaeologists “have been in too much of a hurry to make Neolithic people resemble ourselves” and that the “time has come to interpret their society in new ways” (Whittle 1996c:7). Whittle seems to favour formal analogies, one-to-one comparisons, rather than relational analogies, which he argues “seems to appeal to some sort of sense of universal meanings and principles” (1997b:143). He states that we need to start somewhere if we are going to interpret prehistoric human social behaviour. However, as pointed out earlier, this 46

‘ritual’, as well as many other terms taken for granted in prehistoric archaeological discourse such as ‘tomb’ and ‘house’, are concepts defined, formed and differentiated in the relatively recent period of Western/modern social life.

important that the different uniformitarian principles employed be identified, rather than be seen as one principle. The current trend in ‘Neolithic’ studies, and other archaeological studies, is to seek explanation of the small-scale in order to explain the larger scale. However, there is a need for a hierarchical arrangement for explanation in archaeology, as processes over different scales of space and time are present (Fletcher 1995:43-7 and 233). Generally, studies in ‘Neolithic’ Britain and Ireland have formulated and constructed arguments and interpretations on a small-scale time frame, with ‘Mesolithic’ studies tending to operate on longer time scales (Gijn and Zvelebil 1997:3). For example, the deposition of human skeletal material in a ‘tomb’ has been related in the ‘Neolithic’ to a concern for the ancestor, which has been used to explain the construction of large stone structures and the deposition of human and animal skeletal material, pottery, lithic material and other material within the ‘tomb’. By contrast I will seek to identify the larger scale operations of both the ‘Mesolithic’ and the ‘Neolithic’, avoiding the need to devise a further logically invalid explanation in terms of ‘ritual’/verbalised meaning. I will be pursuing an operational explanation rather than the current, contextual substantively based view which dominates ‘Neolithic’ studies in Britain and Ireland. I will only use analogy when it is to identify the known operational limits of a particular human social behaviour. I will seek to identify the actual relationships and how the elements of human social behaviour from the early tenth to the late fifth millennium BP are operationally linked.

The word ritual is derived from the Latin adverb rite, translated as ‘with due observance’ (Brück 1999a, 1999b; Chapman 2000b; Nordström 1997; Wilkins 1996). Attempts to define ‘ritual’ in archaeological assemblages and contexts have been rather broad and problematic. They tend to encompass activities that cannot apparently be associated with a utilitarian or domestic explanation. However, to categorise human social behaviour as secular or ritual is far from universal and is more a function of our current Western/modern social reality (Barret 1991, 1994:77-84; Brück 1999a, 1999b; Lane 1986; Leach 1966:403). Richards and Thomas (1984) and Whittle (1988a) state that ritual is a ‘structured’ or deliberate patterning of deposited objects. Consequently, the definition suggests “that domestic activity is unpatterned or recognisably less structured” (Luff 1996:1). Obviously this is not the case, since ritual can characterise the secular and domestic realms (Douglas 1966) and tacit patterning of daily life is a normal condition of community life (Fletcher 1995:Chapter 2). Further, ‘mundane’ objects can be identified in contexts that are classified as ritualistic. Thus, the situation exists where a pottery sherd can, on the one hand, be interpreted as a result of domestic activity and also due to ritual activity. Context is everything, but the problem exists of how one is to ‘see’ the context. There is no reason to expect that human social behaviour will be, or should be, so transparent in archaeological assemblages and contexts, and that it should necessarily conform to such rigid dichotomy that prevails in our current Western/modern interpretations of social life.

Identifying ritual human behaviour: vague yet familiar Ritual: All-purpose explanation used where nothing else comes to mind. (Bahn 1989b:62)

A comprehensive workable archaeological definition for ‘ritual’ has yet to be constructed, and any contextual definitions or identifications have tended to be idiosyncratic. The issue may not be in identifying what constitutes an archaeological assemblage or context as ritualistic. Brück (1999a, 1999b) has recently critiqued the issue, identifying that ‘ritual’ is frequently identified and described as non-functional and irrational, while ‘secular’ activities are identified by archaeologists as “governed by a universally-applicable functionalist logic” (1999b:311). She goes on to argue that attempts to separate ritual and secular activity are flawed due to the similarity of non-ritual and ritual deposits. Thus one cannot identify archaeologically an assemblage as either ‘ritual’ or ‘rubbish’, as both concepts are purely a Western construct based on the sacred:profane dichotomy (Bell 1992; Chapman 2000b:62; Pollard 2002:23). There is no reason to expect that this ‘Western’ dichotomy will necessarily be identifiable in the prehistoric archaeological assemblages and contexts in Britain and Ireland or anywhere else for that matter. Brück (1999b:313) states that, as far as the anthropological evidence is concerned, “ritual is generally considered practical and effective action by its practitioners. This is because different conceptions of instrumentality and

Much has been written on the identification of ‘ritual’ activity in archaeological assemblages and contexts (for example, Andersen and Boyle 1996; Bahn 1989b; Barrett 1991, 1994, 1996; Bradley 1991, 1995c; Bradley and Gordon 1988; Brück 1995, 1999a, 1999b; Chapman 2000a, 2000b; Cooney 1994, 1997; Garwood et al. 1991; Gosden 1999; Harding 1981; Hawkes 1954; Hill 1992, 1995, 1996; Hodder 1992; Kaul 1991-2, 1993; Kirk 1998; Lane 1986; Lohof 1994; Luff 1996; Malmer 1986; Miket 1985; Mount 1994; Needham and Spence 1997; Nordström 1997; Oosterbeek 1997; Renfrew 1985a; Richards and Thomas 1984; Shanks and Tilley 1982; Thorpe 1984; Topping 1996; West 1996; Whitehouse 1996; Whittle 1988a; Wilkins 1992; Wilson 1988). While an in-depth study of ‘ritual’ is beyond the purpose of my research, I would like to discuss efforts to distinguish between ‘ritual’ and ‘domestic’ activity in archaeological assemblages and contexts. Morris (1992:8) recently stated that; “Ritual is one of those words where we all know what it means but no one can define it”. Interestingly, the word is not even mentioned in the 1992 Collins Archaeological Dictionary (Bahn 1992). It is important to remember that terms like 47

identified as related to mortuary ritual in keeping with the assumed mortuary function of the context. A similar, inverted paradoxical practice will be outlined in relation to human remains within occupation contexts in Chapter 5. In that case, the ‘mortuary’ role of the ‘occupation’ context is marginalised in keeping with the assumed occupation function. Therefore, certain key elements of the empirical evidence are promoted or marginalised in different contexts to fit the prevailing interpretation. Further, even if there is a total absence of a certain assemblage, such as human skeletal material, the interpretation of a particular context as serving a distinct function, such as a ‘tomb’, is invulnerable to the empirical evidence. Empirical evidence is selectively manipulated according to a set of predefined dichotomies.

causation inform such activities”. The situation is not resolved or improved by replacing ‘ritual’ with other laden and Western/modern terms, such as “structural deposition” (Barrett 1991, 1996; Brown 1991; Chapman 2000a, 2000b; Hill 1995a; Richards and Thomas 1984), “non-domestic”, “unusual” (Whittle 1985b:220 and 233), “symbolic or placed” (Pryor 1988:114), “intentional…symbolic” (Evans 1988a:89) or “purposeful” (Richards and Thomas 1984:215). All are potential dichotomised terms and are still based on the sacred:profane dichotomy. Further, they still aim to differentiate human intent and behaviour, regardless of the empirical evidence. Assemblages and contexts are identified as ‘ritual’ if they do not fit into our understanding of things that are observed or documented historically, that is they are unparalleled forms of human social behaviour. Postprocessualism argues for and recognises, the unique internal cultural structure of societies, but also that archaeology has to be self-reflective on the Western/modern cultural system we inhabit. This approach therefore cannot logically assume that the dichotomies that seem to prevail in our modern lives constitute some universal practice. The current belief is that there is some sort of Western/modern universal commonsense of how human social behaviour should be patterned across the landscape.

Human social behaviour that could be identified as ‘ritual’ (by our modern standards) was surely occurring in Britain and Ireland from the early tenth to the late fifth millennium BP. However, the issue is the appropriate logical scale at which it can/should be recognised. There needs to be a more rigorous and concerned approach to identify such activity, such as the work of Renfrew (1985a) at Phylakopi on the island of Melos in Greece, where he systematically identified those attributes that may signify ritual or cult activity. Currently, ritual is ascribed to so many contexts that ritual as a significant and meaningful form of human social behaviour is lost. The current discipline actively seeks to identify particular discrete locations of human activity across the landscape: be they ‘domestic’, ‘ritual’, ‘burial’, etc. There is no a priori reason that these human practices should be separated into discrete locations across the landscape. Moreover, I will argue that identifying these as always having their own archaeologically distinct signatures is based on a logically invalid assumption linked fundamentally to our own ‘commonsense’ and Western/modern assumption about the patterning of daily life. I do not doubt that specific instances of ritual existed. However, I argue that archaeologists through the study of archaeological assemblages and contexts are unable to identify this scale of human social behaviour, without analogy with historically documented or observed human behaviour, risking the fallacies of substantivism.

In the case of Britain and Ireland, the situation exists where assemblages in ‘tombs’ and other ‘ritual’ contexts, such as henges, would be identified as occupation debris in other contexts. However, the assumption about the dichotomy between the ‘domestic’ and the ‘ritual’ activities is maintained, with the two “treated as distinct and bounded categories, fundamentally opposed to one another” (Brück 1999a:62, also see Brück 1999b:316 and pages 3-6 for the re-interpretation of Barkaer, Denmark). The current employment of ‘ritual’ in the archaeological discourse takes unparalleled elements of the archaeological assemblages and contexts, which do not correspond with observed or documented modern social constructs, and resolves the ambiguity caused by identifying them as ‘ritual’. Thus, if we are to find unparalleled human social behaviour in the past, the vague definitions and the continual use of the ritual to describe archaeological contexts is not the solution, rather it is part of the problem. This is all the more curious as the current discourse argues for the uniqueness of the ‘Neolithic’, but fails to implement the policy at an elementary level. The current discourse appears to use the social construct of ritual to specify that the different was really the same.

Selective examination of archaeological assemblages and contexts in the British and Irish ‘Neolithic’ has insinuated a separation between the living and the dead – creating a dichotomised landscape. The examination of the empirical evidence I am advocating in terms of operational processes of human social behaviour can therefore expose the dichotomy as a product of archaeological discourse. The discourse draws on our current Western/modern social reality where the living and the dead are perceived as segregated across the landscape in discrete locations. In such a system of logic, familiar social criteria and expressions of verbal meaning, such as ‘ritual’, are the inherently vague lubricant used to adjust the evidence to these underlying assumptions.

This situation illustrates the problem of using standard social theory to ‘organise’ archaeological evidence. The dichotomised identification of structures as ‘tombs’ or ‘houses’ creates a situation whereby data that are anomalous, that do not fit the dichotomy, have to be ‘explained away’ to maintain the dichotomy. Explicitly, in presumed ‘burial’ contexts, debris that would elsewhere be called ‘occupational’ is marginalised, being 48

been historically documented or observed. It may even lead us to see the ethnographic and historical evidence in a new way.

Identifying the Unparalleled in Prehistoric Britain and Ireland Archaeological phenomena are ambiguous, opaque and manifestly not the subject of commonsense understandings. (Murray 2002:57)

A way forward… The current archaeological practice of social explanation involves the creation of a version of the verbalised meanings which are supposed to have been associated with past human social behaviour. My interpretation of Britain and Ireland from the early tenth to the late fifth millennium BP will not attempt to convert that evidence into verbal meaning. Instead, I will examine human social behaviour by asking different kinds of ‘social’ questions about ‘materiality’ – the pattern on the ground – questions which are specifically sensitive to the information on human social behaviour that we actually encounter in archaeological assemblages and contexts, not derived from other disciplines concerned with other ways of reading human social behaviour. Due to noncorrespondence between the social and the material, there should be a move towards a greater emphasis on the actual, material characteristics and outcomes of behaviour in archaeological assemblages and contexts (see Part III).

Ethnographic studies have illustrated that there is no oneto-one correlation between the material of social life and the verbal meanings and actions of individuals, a fact recognised by the postprocessual discourse. Further, there are no “simple correlations between an economic phenomenon like agriculture and a social phenomenon like sedentism, or between irrigation and urbanism” (Fletcher in press). Fletcher argues that the material should be viewed as an “actor without intent” (1995:228), and archaeological theory should “seek to comprehend the material as a fundamental and autonomous player in community life” (in press, also see Fletcher 1996). Identifying material and social relationships in the ethnographic record and projecting these onto the past is illogical due to the non-correspondence between the material and the social. However, it goes deeper, with the current interpretive framework assuming a universal status of ‘commonsense’ human social behaviour. There is little thought of material-social dissonance. The paradox is that “if this was the case…then the story of the past would gain its specificity and unique quality from the consequent failures” (Fletcher in press).

While agreeing with Thomas that “no set of concepts or ideas developed in the present will ever grasp the whole essence of the past”, I do not agree that the logical conclusion of this is to “simply write a story” (1999a:5). This is fine if such works are presented and marketed as fiction. It is in the writing of these stories in the current discourse that the unparalleled behaviour of the past is lost. Attempting to identify the social significance and social meaning related to archaeological assemblages and contexts, results in the past being ‘analogised’ with the present and ultimately becomes just a quasi-present. If all archaeology is doing is writing stories derived from our social and cultural preoccupations of historically documented and observed human social behaviour, we need not bother with the expensive and time consuming practice of collecting archaeological data!

Notions of non-correspondence, and the related issue of non-commonsense, lead to a situation whereby there is no a priori reason to assume or expect that past societies would have organised material, and thus human social behaviour, in a way substantively analogous to historically documented and observed human social behaviour. Therefore, dichotomised categories which are generally assumed to be in some way universal, such as ‘tomb’/‘house’, ‘ritual’/‘domestic’ and ‘human’/‘animal’, cannot simply be understood to be applicable to the interpretations of the characteristics of past human societies. Rather than starting from the position that such categories exist and are there to be identified, they must be demonstrated. I will not assume that these dichotomies exist, rather I will document the actual patterning of activity and deposition across the landscape. Archaeological assemblages and contexts identified from the early tenth to the late fifth millennium BP in Britain and Ireland illustrate human social behaviour which has no specific parallel with historically documented or observed human social behaviour. There is no reason to expect past prehistoric human social behaviour to fit neatly into our predefined functional categories. If they did, the past would be predictable and its study dull. The value of archaeology is not that we are able to identify human social behaviour as similar to contemporary behaviour. Rather, it is the fact that new relationships between the material and the ‘social’ can be identified in the past, producing new insights into forms of human social behaviour. The past is interesting for the very reason that it offers an alternative to behaviour which has

Archaeological investigations should not attempt to ‘normalise’ or ‘familiarise’ the prehistoric past, “the process whereby the ontological and epistemological distinctiveness” of archaeological assemblages and human social behaviour “is lost in favour of intelligibility in terms of the conventions of short-term social theory” (Murray 1997:451, 2001). The use of standard social theory is seen as crucial if human social behaviour is to be identified in the prehistoric past. The material itself is seen as inadequate to reconstruct the activities of the population and standard social theory is viewed as the tool that can provide this information. However, we must seek to see whether there is an ontological distinctiveness to the past. This is not to say that prehistoric human social behaviour was ontologically different from that represented in historic periods. Rather, the potential exists and our interpretive tools should allow this to be documented. Currently, “in the bulk of cases such prehistoric behaviours cannot be comprehended in 49

uniformitarianism, drastically reduces the scope of human social behaviour which can be possibly identified, to the point that the archaeological past is liable to be a pale imitation of the findings and declarations of social anthropology and the other social sciences. The history of archaeological investigation has been to adopt methodologies and interpretative tools that have been developed to cope with human social behaviour historically documented and observed in contemporary or recent contexts. We must recognise the uniqueness of archaeological assemblages and contexts which can deal with human social behaviour on a time-scale that is not encountered in the disciplines of other cultural studies. We must begin to develop our own methodology and interpretative tools specifically sensitive to the discipline (see Fletcher 1995; King 2001; Murray 1997). We should want to be in a position to identify actual human social behaviour represented in archaeological assemblages and contexts, and avoid the current pitfalls of interpreting prehistoric evidence of what humans thought.

anything other than circular terms if we apply social theories derived from the ontology of the contemporary human sciences” (Murray 1997:454). But how are we to envisage the inherently problematic? The fact is that observations in archaeology over the last 200 years have been able to identify extraordinary human social behaviour which had no simple parallels in recent or observed behaviour. A case in point is the initial discovery of ‘Upper Palaeolithic’ cave art in the nineteenth century. Interestingly, due to the curiousness of this discovery, its antiquity was rigorously questioned for decades after its discovery. There have been attempts to interpret this art through ethnographic analogy, in essence ‘normalising’ its meaning. Recently, however, alternative approaches have been sought which acknowledge the strangeness of the material. Bahn and Vertut (1988:193) discussing ‘Upper Palaeolithic’ art have noted that: In the past, we have tried to ‘run’ before we can ‘walk’, trying to extract more information than was available, simply because we dislike unintelligible things. This raises the question of how to document what is foreign to us and the language we use. Of course I have used the English language to describe the past, for me that is unavoidable. I did not begin by wanting to conceive the unparalleled and unfamiliar, rather I have attempted to document the past without substantive analogy to socio-cultural phenomena in the present or the recent past. Part of this has been the avoidance of the inherent categories of ‘site’ function which plagues the current archaeological discourse. Identifying a context as a ‘tomb’ or ‘house’ does not in itself allow us to understand the prehistoric past better. Rather it creates a false sense of familiarity and understandability, which leads to the imposition of social and cultural associations apparent in the historically documented and observed contexts. Labelling a structure as a ‘tomb’ or a ‘house’ leads to the imposition of many social and cultural associations, which may or may not have been present in the past.

Archaeology needs to pursue a logic and methodology that allows the potentially unfamiliar to be identified in the past at an elementary level. Then perhaps archaeology can begin to systemically identify diverse and unparalleled human social behaviour. There needs to be a move away from attempting to find specific, substantive analogies of ‘meaning’ in the present to explain the past. Currently the potential unparalleled nature of archaeological assemblages and contexts is marginalised in favour of the preservation of standard social viewpoints. Anomalous data in archaeological assemblages and contexts are reduced to classes of human social behaviour as attested by the familiar – the ‘intransigence’ of archaeological assemblages and contexts is lost in favour of understandability. Archaeology is in the position to identify “unimagined and unimaginable fields of knowledge” and human social behaviour (Murray and Walker 1988:284). The history of archaeological theory has been one of normalising past human social behaviour through circular arguments drawn from other social sciences (Wobst 1978). Instead, we should attempt to challenge assumptions about the scope of human social behaviour that has been historically documented and observed by ethnographers and others. This can only be a tiny fraction of the great breadth and complexity of human social behaviour that has materialised in archaeological assemblages and contexts.

Conclusions Any review of archaeological theory, or its form of discourse known as theoretical archaeology, would raise serious doubts about whether archaeological data are consequential, beyond providing a locale for the colonization of the past by whatever system for comprehending the present practitioners are currently attracted to (see, e.g., Murray 2001). (Murray 2002:56-7)

Prehistoric archaeology must recognise its unique position and opportunity to identify new and challenging forms of human social behaviour, which may help us “learn something new about human nature not available from other sources” (Bailey 1983:165). This is a long way from the discourse, which has dominated archaeology since the nineteenth century, where our role has been seen as identifying prehistoric human social behaviour by social analogy with the declarations of ethnology and anthropology. Archaeology can take a much more active role in human social behaviour studies by further developing the history of humanity and

No matter how committed the researcher is in identifying ‘difference’ in archaeological assemblages, the use of standard social theory does not allow this (for example, Thomas 1999a:Chapter 1). Maintaining the use of standard social theory, drawn on ethnographic and ethnohistoric analogies and implemented using substantive 50

face of the dramatically disconfirming data, retain frameworks of interpretation or explanation because they cannot think their way through new frameworks, or because the impact of such changes on the cognitive map held by everyone else would be too great.

documenting the vast scope and variety of human social behaviour. Such a step will not be easy and will require much examination of archaeology as a discipline. As Murray (1993:183) notes: it is entirely possible that practitioners can, consciously or unconsciously, and even in the

51

CHAPTER FOUR: MAKING MEGALITHS FROM STONES discussed the implications of the terminology we employ. Highlighting the movement and deposition of human skeletal material across the landscape, he states that it is inappropriate to continue to describe these stone, timber and earthen constructions as ‘tombs’ or ‘barrows’. Bradley recommends that a more neutral term should be employed to describe these constructions, suggesting ‘mortuary monuments’. While such a critique is welcome and necessary, the new term fails to address the predicament that has and continues to dominate the archaeological discourse. Identifying these structures as ‘mortuary’ still leaves the primacy of mortuary practices located in stone chambered structures and earthen and timber mounds in the landscape.

My basic assertion is that the Neolithic of the Atlantic façade will not be understood until we realise more of the preceding and coeval Mesolithic. (Kinnes 1984:367) Chapter 2 illustrated the lack of evidence pertaining to agriculture in the ‘Neolithic’. The belief that agriculture commenced in the ‘Neolithic’ has been used to explain the existence and construction of stone, timber and earthen constructions (‘megaliths’ and ‘barrows’) across the landscape, as I will discuss below. In turn, the construction of stone chambered structures and timber and earthen mounds and their identification as ‘tombs’ and ‘barrows’, has helped develop claims for the existence of ancestors and ancestor worship in the ‘Neolithic’ (see Chapter 5). This chapter will be used to argue that, as with subsistence, ‘megaliths’ can be seen as part of a continuum in the use of stone from around 30000 BP in ‘Upper Palaeolithic’ assemblages to the beginning of the ‘Neolithic’. Further, I will argue that the identification of all stone constructions as ‘megalithic’ and of earthen, timber and stone constructions as ‘tombs’ or ‘barrows’ is not always appropriate.

Before proceeding, I must state that I do not wish to imply that these constructions were never used as depositories of the dead. My argument only follows up the implications of the observation noted by Bradley and many others in ‘Neolithic’ studies, that human remains occur in a range of locations across the landscape. Instead of ‘privileging’ the stone, timber and earthen constructions identified by archaeologists as ‘tombs’ or ‘barrows’ – by claiming that they possessed the primary or definitive ‘mortuary’ function – I wish to argue for a distributed system which did not, in practice, segregate life and death.

Identifying ‘Tombs’ and ‘Barrows’ in ‘Neolithic’ western and northern Europe

Ethnography and the megalithic structures of western and northern Europe

Since the eighteenth century antiquarians, William Borlase and William Stukeley, chambered structures and earthen and timber mounds have been described and interpreted as tombs and barrows (Fleming 1973). The funerary identification of stone chambered ‘tombs’ and timber and earthen ‘barrows’ has been based on two types of evidence: ethnographic analogy and archaeological data. The two forms of evidence have been intricately linked since the beginning of the antiquarian and archaeological investigation of these constructions in the eighteenth and nineteenth century. Sir John Lubbock (1869:Chapter 5, especially 122-25 and 164-65) identified ethnographic examples, including examples from India and Madagascar, of large stone structures that, in some cases were used as burial chambers. It is interesting to note that the Madagascar case has recently been re-examined to interpret the stone chambered structures of Britain, an issue that will be discussed below. This, coupled with evidence from antiquarian excavations, which often found human bone in the western and northern European stone chambered structures, seemed to convince Lubbock that these constructions were ‘tombs’ (Rault 1993-4:96). Despite a few exceptions (see Ashbee 1976; Case 1969; Kinnes 1984; O’Kelly 1981; Rault 1993-4) this has been the prevailing view since. Recently, Bradley (1998a:54)

The varied reasoning behind the employment of ethnographic evidence in relation to prehistoric Europe makes for some interesting reading (for example Burenhult 1987; Clark 1953; Lewis-Williams 1998; Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Scarre 1994; Tilley 1996a; Verhart and Wansleeben 1997). Parker Pearson and Ramilisonina (1998a:309) state that they employ ethnography and cross-cultural generalisations (standard social theory) “as a means of assessing the likelihood of certain aspects of social organization being shared between different cultural contexts” (emphasis mine). Scarre (1994:79) states that ethnographic analogies “can make the archaeological data less mysterious and alien”. As outlined in Chapter 3, Tilley (1996a:6) uses ethnographic evidence to “normalize” his interpretation. He claims to use ethnographic evidence as a medium through which to view the archaeological evidence, rather than a model to which the data are to be fitted or against which the data are to be tested (ibid.:2). In the case of Tilley, his reasoning is quite puzzling, as in essence his ‘archaeological’ interpretations are inherently locked in with the ethnographic evidence. I will show that his 52

claim seems unlikely, since the discourse concerning these constructions in western and northern Europe has a solid and long tradition of combining ethnographic analogy and standard social theory (for example, see Lubbock 1869, 1872 and Chapter 1). To illustrate my point, I would like to outline a number of quotes in which Tilley details his view of the ‘Neolithic’. These should be seen not only to characterise his view of southern Scandinavia, but also other regions of western and northern Europe dominated by megalith ‘tombs’ and earthen and timber ‘barrows’. I am focusing on Tilley as his view can be taken not only as representative of the new ‘radical’ works, but also, ironically, that which has prevailed since Childe. It is important to state that, by Tilley’s own admission, his interpretation is based solely on archaeological material (1996a:6). Tilley (1996a:334) states: The tombs enclosed or digested the bones of the deceased who became converted into ancestors in the process. The containment of the bones within the tomb served to incorporate an ancestral presence within its own enduring continuity, serving to legitimate the successorship of the living through their honouring of the ancestral dead. The tomb not only contained the bones but nurtured them. Artefact destruction released power and heat that cooked together the ancestral bones which were periodically fortified with fresh bones forming a supply of materials that could be harvested, like grain, to nurture the living. Further: The long mounds and dolmens became a central focus in the landscape...The social practices and relationships involved in monument building and use were obviously intimately connected with higher spiritual and ancestral powers...The ceremonies taking place outside were staged events involving song, dance, feasting and artefact sacrifice to ancestral powers...That the ancestor required incessantly food, valuables, amber, axes and pots was a constant fact of life and guide for social action during the Neolithic. As all-seeing and all-knowing forces they could make things happen for better or worse, Ancestral forces enveloped the group in a mantle of protection but if votive and tomb offerings ceased to be made their potency would go cold leaving the population at the mercy of malevolent spirits hovering at the margins of the community. (1996a:112 and 334-5, emphasis mine)

interpretations could not be made solely on the basis of archaeological evidence. Tilley uses a number of ethnographic examples in his aptly named, An ethnography of the Neolithic (1996a). In particular, he links ‘Neolithic’ mortuary practices with two “documented cases in small-scale non-Western societies” (1996a:235-41). The first case involves the practices of the Trobriand Islands communities, through which he attempts to link the ritual cycles of feasting and exchange of food, axes, pots and shells with the distribution of artefacts found in and around megalithic ‘tombs’ (see Fewster 2001). Tilley’s second ethnographic example is taken from the Merina of central Madagascar, the most commonly employed ethnographic parallel in relation to the European megalithic ‘tombs’ in recent times (for example Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Thomas 1988b; Scarre 1994). The Merina construct and use masonry tombs today (Figures 4.1 and 4.2), which are derived from earlier megalithic structures (Figure 4.3). Amongst the Merina the constructions relate to ancestors, with the tombs serving as a medium through which the ancestors, the kin group and the land merge (see Kus and Raharijaona 1998). However, the Merina situation is unusual, with the association between large stone constructions and ancestor worship being very rare in the ethnographic literature (Newell 1976; Scheinfeld 1960, also see Whitley 2002). As it is ‘anomalous’ today and in the recent past, it is surely questionable to use it as some universal model to suggest the association of large stone constructions and ancestor worship in the past with a wholly different cultural and ecological context in which even the economic foundations of society were different. Tilley fails to stress that the Merina society is literate and iron-using, and thus not exactly comparable to those societies of ‘Neolithic’ Europe. Furthermore, and very seriously, the period and region under discussion in Britain and Ireland were also characterised by a millennium-long interaction between stone-using ‘farmers’ and stone-using ‘hunter-gatherers’, a situation that has few, if any, ethnographic parallels (Whitley 2002:120-2; Zvelebil 1996b:341). Tilley (1996a:59) states, in relation to the ‘Mesolithic’ communities of southern Scandinavia: We are dealing here with a unique set of communities for which there is no direct parallel among hunter-gatherers known to us through anthropological research and which we have to try and understand in its own terms. He then, in defence of the use of ethnographic evidence in archaeology, claims that: the discussion of specific ethnographic examples in the text is not to produce an interpretation – they were added after an interpretation of the archaeological data – but to ‘normalize’ what might otherwise appear ‘unwarranted’ arguments and interpretation. (1996a:6, emphasis mine) It is hard to see how such a claim could be justified – it is as if Tilley makes the argument that his reasoning is pristine and free of the ethnographic evidence. Tilley’s

As stated above, Tilley suggests that his interpretations are based solely on archaeological evidence, with ethnographic examples used to ‘normalise’ the interpretations. However, such interpretations are not derived from archaeological observations per se, rather they are constructs using substantive analogies from the ethnography. This is done despite his claim that there is infinite capacity for diversity in human cultures which, he has argued, makes cross-cultural generalisations

53

Figure 4.1 – A Merina tomb, Madagascar (after Bloch 1971: figure 7).

Figure 4.2 – Photograph of an “above average” Merina tomb, Madagascar (after Bloch 1971: plate 3b).

54

Figure 4.3 – A Tandroy tomb from Madagascar (after Parker Pearson and Ramilisonina 1998a: figure 3). Sjögren 1986). Such a purportedly close relationship between the construction of megaliths and the arrival of agriculture on the European Atlantic fringes is linked both with the expectation that a surplus supply of food is required for unproductive labour investment and with Western/modern concepts of territory and property (for example, Case 1969; Renfrew 1973, 1976; Sjögren 1986). However, the constructions in Britain and Ireland were not accompanied by the wholesale adoption of agriculture (see Chapter 2), nor is it the case that huntergatherers do not or cannot build large constructions (P.C. Edwards 1989; Scarre 2002b).

redundant for any other purpose than “the mundane or the trivial” (ibid.:1). Yet Tilley has concluded, through the examination of the rare Merina case, that the ‘Neolithic’ European megalithic structures were ‘tombs’ used for ancestor worship, under the pretext that ethnographic evidence seems to suggest that ancestor worship is common in small-scale societies. Tilley has exercised a generalised ‘sleight of hand’, making a fallacious link between the rare practice of constructing large stone monuments concerned with ancestor worship in smallscale societies in the ethnographic record, and the societies of prehistoric Europe who built large stone constructions.

Hunter-gatherers in Australia, the Arctic Circle and the Northwest-Coast of America produced large constructions of earth, stone and timber. In Australia, for example, there are examples of eel and fish traps which artificially linked swamps and ponds, covering over five hectares (P.C. Edwards 1989; Lourandos 1976:182, 1997:64-5). In Alaska, Nunamiut hunters can erect large stone structures. In the Tulugak Lake area, lines of stones over a kilometre in length are built as aids to hunting drives. There were also over 60 stone structures and cairns which were used to preserve stored meat, sometimes built up to form corbelled domes (Binford 1978:206; P.C. Edwards 1989:18). The Nuer, a semimobile pastoralist population living on the Nile River in Sudan, constructed an earthen mound (‘pyramid’) nearly 20 metres high with large elephant tusks encircling the base and on top of the mound (Evans-Pritchard 1963:186). These examples illustrate the operational constraints of the construction of large structures in relation to subsistence. They refute the assumption that

‘Megaliths’ in Archaeological Discourse The term ‘megalith’ derives from the Greek megas, ‘big’, and lithos, ‘stone’, however, it is also applied to earthen and even timber constructions. According to Daniel (1958:14) the term was first used in the description of the stone monuments in England and other parts of the world between 1840 and 1860. Thus, the term ‘megalith’ became current at the same time the chronological/cultural terms ‘Mesolithic’ and ‘Neolithic’ were first defined (Tilley 1998:142, see Chapter 1). Early theories in relation to the ‘origin’ of megaliths have been intimately linked to the assumed ‘spread of farming’ west across Europe (see Daniel 1958). While the notion of missionaries from the East was dismissed in the 1970s due to radiocarbon dating (Renfrew 1973), megaliths and farming are still linked in many interpretations (see 55

been viewed as the foundation of a new social order in the ‘Neolithic’ (for example, Barrett 1994; Hodder 1990a; Tilley 1996a; Thomas 1991a, 1999a; Whittle 1996a). From the cultural remains, there has been an attempt to reconstruct a ‘Neolithic’ ideology. As we have seen, such an archaeological discourse produces a false cultural dichotomy between the ‘Mesolithic’ and ‘Neolithic’ periods, which masks the continuity between the actual cultural assemblages. Bradley (1998a:21) states that such analysis, while prevalent in ‘Neolithic’ studies, has not been extended to the earlier period where similar material is available for study. The issue is not that scholars have missed the similarity of the data, since some have attempted such studies in ‘Mesolithic’ contexts (see Clark and Neeley 1987; Ingold et al. 1988a, 1988b; Jeunesse 1996; Larsson 1989a, 1989b, 1990a, 1990b; Mithen 1991; Neeley and Clark 1990; Schulting 1996; Zvelebil and Jordan 1999). The issue is the different theoretical and interpretative approaches which directly affect the questions that are asked of the data. The ‘Neolithic’ discourse has been more concerned with the symbolic and phenomenological significances of assemblages and contexts, while the ‘Mesolithic’ discourse has focused upon ecological and subsistence issues.

the construction of large structures requires agriculture, or for that matter sedentism. While there is no doubt that the increase in the appearance of stone, timber and earthen constructions in the British and Irish ‘Neolithic’ landscape occurred virtually at the same time as the appearance of domesticated plants and animals, the dual and determinative linking between an agricultural economy and the ‘megaliths’ is questionable. Chapter 2 pointed out that the initial appearance of these domesticates had little immediate impact on subsistence. Therefore, their impact on the resource base of the builders of these constructions cannot be presumed to have been significant. The association between domesticates and these structures has been based on the assumption that economic change is accompanied by cultural change, or vice versa. However, it is clear from the evidence that hunter-gatherer subsistence continued to contribute significantly along the European Atlantic fringes down to at least the late fifth millennium BP. For the purpose of my research, I will only employ the term ‘megalith’ or ‘megalithic’ when discussing literature that uses such terms, otherwise I will refer to such constructions as stone structures (or earthen and timber constructions if appropriate), with a reference to the shape of the feature and the size of the stones when it is available. I hope that this will limit problems with what size of stone qualifies to be a ‘megalith’ – a matter that has troubled the discipline. According to Rault (19934:85) it is a stone which is too large for an individual to carry (c. 100 kg). Childe (1957:213) himself struggled with the use and implications of the term ‘megalith’. The use of large and small stones, and the existence of rockcut structures, made Childe question the adequacy of the term ‘megalith’ (see Tilley 1998:157). It is important to note however, that the constructions which are clearly to be labelled ‘megalithic’ and ‘monumental’, such as Stonehenge, West Kennet, Avebury (Wiltshire), Newgrange, Knowth (Co. Meath), Maes Howe (Orkney Islands), Gavrinis (Morbihan, France) etc., are at the very large and rather rare end of the scale and are relatively late in the sequence dates of construction. I argue that focusing on those examples leaves out a wide range of uses of stones and obscures continuity in the use of stones of varying sizes in constructions dating back into the ‘Upper Palaeolithic’.

Bradley (1997a, 1998a:Chapter 2) has recently attempted to address the imbalance in discussing two prehistoric graves from Dragsholm in Zealand, Denmark (Figure 4.4) – one grave assigned to the late ‘Mesolithic’ and the other to the early ‘Neolithic’. The chronological distinction was based both on radiocarbon dates and artefacts associated with the graves. Bradley (1998a:23) states that much has been made of the symbolic elements of ‘Neolithic’ burial practices, while little has been made of the ‘Mesolithic’ example. Rather, burial practices in the ‘Mesolithic’ are related to changing residential patterns, control over critical resources or economic intensification (for example Chapman 1981; Clark and Neeley 1987). Thomas and Tilley (1993) have also discussed differentiation in the study of the ‘Mesolithic’ and ‘Neolithic’ archaeological assemblages and contexts. They suggest, in relation to the shell middens at Téviec and Hoëdic (Morbihan, France), that there has been a marked dichotomy between the ‘Neolithic’ as a ‘ritual’/cultural phenomenon and the ‘Mesolithic’ as an economic phenomenon. They suggest that viewing ‘Mesolithic’ shell middens in purely economic terms fails to recognise their ‘ritual’ component (Thomas and Tilley 1993:228, also see Pollard 2000b; Warren 2001). Despite this, it is important to note that the assumed dichotomy continues to be a major component of both Tilley’s and Thomas’ work (for example, Thomas 1991a, 1993a, 1996b, 1996c, 1999a; Tilley 1993, 1994, 1996a). Most examinations of the ‘Neolithic’ are inherently locked into the assumption that the visible ‘constructions’ of western and northern Europe are an essentially ‘Neolithic’ phenomenon. This chapter has already mentioned cultural attributes (communal feasting, ancestor worship, etc.) which are said to accompany the use and creation of these stone structures. Notions of conceived perceptions of place (Bradley 1993:17), ancestors (Barrett 1994, 1996;

First, the number of theories on the development of ‘megalithic’ structures in western and northern Europe will be outlined. I will seek to demonstrate that these theories generally fail to account for the pre-‘Neolithic’ uses of stone for construction. Second, I will then argue that the earlier uses can be seen as precursors for the stone structures that are so conspicuous in the western and northern European landscape during the ‘Neolithic’.

Constructing a ‘Neolithic’ Ideology The construction of stone constructions in Europe has 56

Figure 4.4 – The ‘Mesolithic’ (left) and ‘Neolithic’ (right) burials at Dragsholm (Zealand) (after Bradley 1998a: figure 7) Bradley 1984a; Clarke et al. 1985; Cooney 2000a; Edmonds 1999a; Lucas 1997; Malone 1994; Mizoguchi 1995; Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Richards 1988; Shanks and Tilley 1982; Thomas 1991a, 1999a, 2000a; Thomas and Whittle 1986; Thorpe 1984, 1996; Tilley 1996a; Whittle 1996a), continuity of territory (Renfrew 1976) and community (Chapman 1981) have been identified as ‘Neolithic’ characteristics, expressed visibly through the material defining of space. In view of the fact that the cultural/economic, chronological dichotomy affects the way that archaeologists view the empirical evidence, as Thomas discussed (1993b:389), this chapter will examine the chronology of the development of stone, timber and earthen constructions of western and northern Europe and their associated subsistence evidence. This will lead into an examination of the use of stone in Europe prior to the ‘Neolithic’, the current standard context for the first ‘megalithic’ structures (Scarre et al. 1993).

‘Megaliths’, Subsistence and Symbols A recent school of thought, headed by Hodder (1984, 1990a, 1994) and Sherratt (1990, 1995), has focused on the ideology of ‘Neolithic’ ‘megalithic’ constructions. Somewhat curiously, the role of central European LBK farming communities in influencing the western and northern European cultural attributes is emphasised – curious in that the LBK built large trapezoidal timber ‘houses’. According to this argument, the LBK long ‘house’ was ‘transformed’ into the long burial mounds which dominated the landscape of western and northern Europe (Figures 4.5 and 4.6). Scarre et al. (1993) comments that radiocarbon dates for a number of French chambered structures within earthen long mounds make them at least as early as the LBK long ‘houses’. Early dates at Barnenez and Ile Guennoc (Finistère, France) obtained in the 1960s and 1970s, and the new dates from Bougon (Deux-Sévres, France), have restarted the debate 57

Atlantic fringes of Europe as a response to the advance of farming from the east (also see Sherratt 1990). Tilley (1996a:109) also states that the long, linear accumulation of shells and associated timber structures, which characterise Ertebølle shell fringes of Europe, contain “All the various middens and other shell middens along the Atlantic ingredients of the long barrow with wooden mortuary houses...The barrow recalls the form of the Ertebølle middens”. Boujot and Cassen (1993:479) indicate that the use of stone at Téviec and Höedic represents “the first expressions of Megalithism” (Figure 4.7). The tradition of collective burial has been identified as a ‘Mesolithic’ development, seen at Téviec and Hoëdic off the Morbihan coast in France (Case 1976; Cullen 1995; Scarre 1992, 1998, 2002a). Collective burial practice is absent from the LBK tradition, but is present in the early Breton ‘Neolithic’, suggesting that the collective aggregation of bits of dead human bodies which dominate the western and northern ‘Neolithic’ may be derived from the social life of the communities who constructed the shell middens (Scarre 1998:174-5).

that the trapezoidal mound forms in western and northern Europe from the first half of the seventh millennium BP “were an early independent development of indigenous Atlantic communities” (Scarre et al. 1993:857, also see Scarre 1992, 1998). Radiocarbon dates from thirty years ago suggested that the constructions within the trapezoidal stone cairns at Barnenez and Ile Guennoc dated back to around the seventh millennium BP.37 These dates were questioned on the basis that they were too early. However, recent radiocarbon dates from chamber F0 within a circular stone cairn at Bougon, suggests that these dates may be correct. Human skeletal material within the structure has yielded the two radiocarbon dates which calibrate to approximately 6650 BP,38 making it the oldest, securely dated chambered ‘tomb’ in western and northern Europe (Scarre et al. 1993:858). However, this assumes that the human bones provide a terminus ante quem for the construction of the structure (however see page 84). However, while these early dates are apparently now secure, the earliest trapezoidal structures in Europe are actually in the Iron Gates Gorge of the Danube (Yugoslavia) from the eighth to the sixth millennium BP (Chapman 1993:91; Kinnes 1984:368; Radovanoviü 1996; Srejoviü 1972; Whittle 1996a:25) in what are generally recognised as hunter-fisher-gatherer communities.

Bradley recently discussed an error in the long held view that the earthen mound constructions were built by farming populations. He identifies that these constructions occupied the precise “areas where agriculture was adopted late and perhaps by local hunter gatherers” (Bradley 1998a:52). Renfrew (1984:187) suggests that Childe (1925:133) was the first to state that there may have been a link between these constructions and ‘Mesolithic’ populations, identifying that the “great centres of megalithic architecture in Europe are precisely those regions where the palaeolithic survivals are the most numerous and best attested” (ibid.). Ashbee (1982), following on from Hencken (1932) and Clark (1977), states that hunter-gatherer groups may have constructed a significant number of ‘passage-tombs’ on the Isles of Scilly. Saville (1989a) and Ashbee (1982) both see ‘Neolithic’ timber and earthen mounds as having their antecedents in the ‘Mesolithic’. In Ireland, O’Kelly (1981:118) suggested that the indigenous population developed ‘Neolithic’ court-tombs. Similarly, R.W. Smith’s 1984 paper argued strongly for ‘Mesolithic’ involvement in the construction of timber and earthen mounds at Horslip (Figure 4.8), South Street (Figure 4.9)

It is important to note that the case for early western ‘megaliths’ was advocated in the 1970s, with its implications discussed then. Because of the early dates obtained from Breton constructions, much was written about the Atlantic context of early ‘megalithic’ structures. Renfrew (1976) even suggested that ‘Mesolithic’ (huntergatherer) communities constructed the ‘megaliths’ on the 37. The dates from Barnenez and Ile Guennoc include 7270-5990 BP (.995) and 5970-5950 BP (.005) (5800±300 bp, Gif-165); 6890-6280 BP (.993) and 62306210 BP (.007) (5750±150 bp, Gif-1309); 6660-6650 BP (.003), 6640-6170 BP (.872) and 6150-5990 BP (.125) (5550±140 bp, Gif-1556); 6610-6610 BP (.000) and 6550-5910 BP (1.000) (5450±140 bp, Gif-1310); 61905590 BP (1.000) (5100±140 bp, Gif-1116). 38. This is based on two radiocarbon dated human skeletal material samples: 6850-6840 BP (.013), 68306820 BP (.002), 6800-6770 BP (.073) and 6760-6500 BP (.912) (5865±65, Q-3234); 7140-7130 BP (.003) and 7010-6290 BP (.997) (5830±160 bp, Ly-1700).

58

Figure 4.5 – The plan of a Linear Pottery Culture long house at Olszanica (Poland) (left), compared with that of Kilham long barrow (East Riding of Yorkshire) (right) (after Bradley 1998a: figure 13).

Figure 4.6 – The distribution of long houses (vertical shading), long mounds and megalithic tombs (horizontal shading) in north western Europe (after Bradley 1998a: figure 12).

59

Figure 4.7 – Synthesis of the development of funerary structures (from simple grave to gallery graves, through mounds with cists and passage graves) and succession of the regional pottery styles (after Boujot and Cassen 1993: figure 3).

60

Figure 4.8 – General plan of Horslip long mound (Wiltshire) (after Ashbee et al. 1979: figure 2).

61

Figure 4.9 – General plan of South Street long mound (Wiltshire). Contours are at 20 centimetre intervals. Roman numerals are ditch cuttings (after Ashbee et al. 1979: figure 23).

62

Figure 4.10 – Plan of Beckhampton Road (Bishops Cannings 76) (after Barrett et al. 1991: figure 2.10). Direct evidence of a continual reliance on hunter-gatherer subsistence has been documented at a number of locations in relation to these constructions (see also Chapter 2). Clark (1977) suggested that an economy based on migratory fish was an important feature of the cultures making ‘passage-tombs’ in western and northern Europe. He identified a number of ‘megalithic’ structures that cluster around rich fishing grounds, such as the Isles of Scilly. However, the recent human bone stable isotope analysis, outlined in Chapter 2, indicates that after 5400 BP no humans had any marine protein in their diets (Richards and Hedges 1999a). Recently, Thorpe (1996:146) indicates that in Scandinavia, those areas with the “most monumental shell middens in the Mesolithic” are where the earthen mounds appear in large numbers (also see Tilley 1996a:109). However, it must be noted that this development does not represent a total shift from the construction of shell middens to the building of these other constructions. Shell middens continue to be utilised beyond the middle of the seventh millennium BP. For example, in Scotland, they were used into the third millennium BP (Pollard 1990; Saville and Hallén 1994) and in Scandinavia into late seventh millennium BP (Andersen 1989).

and Beckhampton Road (Wiltshire) (Figure 4.10). He bases this on pre-mound evidence relating to ‘Mesolithic’ clearings and microliths at Horslip and South Street and a ‘totem-like’ structure at Beckhampton Road. It is also relevant here to mention the Stonehenge car park pit alignment excavated in 1966, with evidence that they held large posts of about 0.75 metres diameter which had rotted in situ. The 3 pits (A, B, C) also contained flint, charcoal, decayed wood and a burnt bone. Charcoal from pit A has been radiocarbon dated to the second half of the twelfth to the first half of the tenth millennium BP.39 Charcoal from pit B has been radiocarbon dated to the second half of the tenth to the first half of the ninth millennium BP40 (Allen 1995, 1997; Allen and Bayliss 1995). 39. The calibrated radiocarbon date is: 11036-11029 BP (.002), 10974-10974 BP (.000) and 10746-9700 BP (.998) (9130±180 bp, HAR-455). 40. The calibrated radiocarbon date is: 9418-9412 BP (.002), 9401-9339 BP (.042) and 9327-8600 BP (.956) (8090±140 bp, HAR-456).

At a number of contexts, explicit examples of the continual hunter-gatherer marine resource basis have been identified, however the contribution these made to subsistence needs to be questioned in the light of recent human bone stable isotope analysis (see Chapter 2) Thick layers of limpets were excavated from the ‘passagetombs’ of La Varde and Le Dehus (Guernsey, Channel Islands) and also from the ‘gallery-tomb’ of Ville-esNovaux (Jersey, Channel Islands) (Clark 1977:45). Shell debris was also found in the passage-tombs of Gavrinis (Morbihan, France) and Vignes Jaunes (Marne, France) (Shee Twohig 1981:172, 197). Ashbee (1982:8) 63

of events and the old-wood effect has not been factored in (see page 25). Despite such criticisms, dates from France and the shell recovered from other contexts, require a reappraisal of the notion that ‘Neolithic’ farming communities constructed these stone structures. As Burenhult (1980:5) concludes: The traditional stereotype: farming communitymegalithic monument can no longer be upheld, and a development within a pre-existing Mesolithic population has been put forth as a preliminary model of the socio-economical background to the chambered tombs at Carrowmore. In my opinion, his observation is only strengthened by the recent reappraisal of the ‘Neolithic’ as an economic phenomenon of continuity with the ‘Mesolithic’, documented in Chapter 2.

documented other instances of shells being associated with ‘megalithic’ structures. Limpets were found in the mound of La Houge Bie (Jersey, Channel Islands), with oysters also being excavated from the chamber; at Les Pourciaux (Alderney, Channel Islands) and from Grantex (Jersey, Channel Islands). From Wales, oysters were excavated at Barclodiad y Gawres (Isle Of Anglesey) and cockles, mussels and limpets were encountered at Bryn Celli Ddu (Isle of Anglesey), Bryn yr Hen Bobl (Isle Of Anglesey) and Lligwy (Isle of Anglesey) (Hemp 1935:297-81; Powell and Daniel 1956). Limpets occur in the forecourt of Crarae (Argyll and Bute) and Cairnholy I (Dumfries and Galloway). Marine deposits have been identified from ‘tombs’; Embo, Tullach an t’sionnaich (Highland) and Quaterness (Orkney Islands). However, these Scottish examples have not been studied in detail and it is not known to what period the deposits relate (Pollard 1990:70).

Below, I will present a number of European predecessors for the constructions that fill the landscape from the beginning of the ‘Neolithic’. I will endeavour to illustrate that the use of stone is not isolated to ‘Neolithic’ contexts, but rather that it can be identified from around 30000 BP to the middle of the seventh millennium BP (pre-‘Neolithic’ contexts).

From Ireland, shells of cockle, periwinkle, pecten, limpets and mussel have been found in the chambered ‘tomb’ of Loughcrew Cairn H (Co. Meath), with indications that they were brought “forty miles inland from Drogheda” (Herity 1974:172). Also, the mound of a ‘passage-tomb’ at Rush (Co. Dublin) contained a thick layer of periwinkle (ibid.:173). McCormick’s (1985-6) survey of faunal remains in prehistoric Irish burials has produced a number of ‘tombs’ with shell associated with them. Shell has been found at the ‘court-tombs’ at Clontygora (Co. Armagh), Ballinran (Co. Down) and at the ‘wedge-tombs’ at Drum and Screedagh (Co. Sligo).

Pre-‘Neolithic’ Use of Stone for Construction across Europe Many examples exist across Europe for stone construction in pre-‘Neolithic’ contexts (see Appendix 2), some of which will be discussed here in detail. The shell middens of Téviec and Hoëdic off the Morbihan coast in France are important cases (Figure 4.11). Both contained inhumation burials, 10 at Téviec (Figures 4.12 and 4.13) and 9 at Hoëdic (Figures 4.14 and 4.15), with the remains of 23 and 14 individuals respectively. The graves were of a stone-built cist with a hearth on top covered with more stones, or an antler structure. The shell middens had been used continually for the disposal of the dead, with new skeletons being added, disturbing the previously buried skeletons. Based on tool typology and burial practices, late ninth to the middle of the seventh millennium BP41

Carrowmore (Co. Sligo) has been the context of the best documented claim for a ‘Mesolithic’ ‘megalithic’ monument. Large quantities of unopened mussels and oysters were excavated from Grave 4. The grave also produced radiocarbon dates, similar to those obtained at Bougon (Deux-Sévres, France) (see above, Burenhult 2001; Scarre et al. 1993), from charcoal in the small central chamber (Burenhult 1980:72, 1984:64). However, the ‘Mesolithic’ interpretation by Göran Burenhult has been strongly criticised by Caulfield (1983) and Bergh (1995), on the grounds that it is based on one radiocarbon date, the charcoal dated could have been from a number the two contexts have been regarded as contemporaneous. Charcoal from Hoëdic has been radiocarbon dated to the

41. The calibrated radiocarbon date is: 8110-8090 BP (.004) and 8040-6670 BP (.996) (6575±350 bp, Gif-227).

64

Figure 4.11 – The Bay of Quiberon region (Morbihan, France), showing locations of selected ‘Mesolithic’ sites (after Schulting 1996: figure 1).

Figure 4.12 – Plan of Téviec (Morbihan) (note the cenotaph?) (after Schulting 1996: figure 2). 65

Figure 4.13 – Grave K at Téviec (Morbihan) (after Schulting 1996: figure 4).

Figure 4.14 – Plan of Hoëdic (Morbihan, France) (after Schulting 1996: figure 3).

66

Figure 4.15 – Grave K at Hoëdic (Morbihan) with red deer antler structure (after Schulting 1996: figure 11). second half of the eighth to the seventh millennium BP.43

and six samples of human skeletal material radiocarbon dated to the eighth to the sixth millennium BP.42 Eight radiocarbon dates from human skeletal material from Téviec have been radiocarbon dated from between the

43. The eight calibrated radiocarbon dates are: 7490-7270 BP (1.000) (6740±60 bp, OxA-6665); 7410-7150 BP (.993) and 7110-7100 BP (.007) (6530±60 bp, OxA6702); 7280-6980 BP (1.000) (6515±65 bp, OxA-6704); 7320-7090 BP (.981) and 7060-7040 BP (.020) (6510±50 bp, OxA-6664); 7300-7000 BP (1.000) (6500±65 bp, OxA-6703); 7240-6970 BP (1.000) (6440±55 bp, OxA6663); 6750-6470 BP (1.000) (6000±60 bp, OxA-6701); 6400-6250 BP (1.000) (5680±50 bp, OxA-6662). Stable carbon isotope analysis indicates that a significant quantity of the protein in the diet was marine-derived so calibration used the mixed atmospheric/marine curve (Schulting and Richards 2001; Stuiver et al. 1998).

42. The six calibrated radiocarbon dates are: 7820-7590 BP (1.000) (7165±60 bp, OxA-6708); 7400-7140 BP (.990) and 7110-7100 BP (.010) (6645±60 bp, OxA6709); 6960-6680 BP (1.000) (6280±60 bp, OxA-6706); 6740-6470 BP (1.000) (6080±60 bp, OxA-6707); 63906170 BP (1.000) (5755±55 bp, OxA-6710); 5660-5450 BP (1.000) (5080±55 bp, OxA-6705). Stable carbon isotope analysis indicates that a significant quantity of the protein in the diet was marine-derived so calibration used the mixed atmospheric/marine curve (Schulting and Richards 2001; Stuiver et al. 1998).

67

Interestingly, Schulting and Richards (2001, also see Schulting 1996:335-6) state that at two standard deviations, a number of dates overlap with the Breton early ‘Neolithic’ ‘passage-tombs’ and the long mounds. Moreover, the arrangement of cists and hearths documented at Téviec is similar to that seen at the long mounds of Tumulus-St-Michel, Le Manio, Kerlescan and Mané Lud (Morbihan, France) (Figures 4.16, 4.17 and 4.18), and the reuse of the graves over centuries is a practice identified throughout the ‘Neolithic’, which will be elaborated on in Chapter 5 (Kirk 1993:212; Scarre 1998:167-9, 2002a; Schulting 1996:347; Schulting and Richards 2001).

Recently, Bahn (1989a:557-8) discussed a burial and occupation context at the Roc del Migdia rock shelter (Cataluña, Spain). Dated to the late Magdalenian, 1387013140 BP (.995), 13070-13050 BP (.005) (11520±150 bp, UGRA-117 RM1a), it contains red deer, ibex and boar bones and snail shells. The rock shelter is particularly noted for its ‘megalithic burial’, consisting of a skeleton surrounded by hearths and large slabs of stone, one that measures approximately 1 by 0.2 metres (Figure 4.21). Bahn (1989a:558) states that a similar burial occurs in an Epi-‘Palaeolithic’ layer of the Abri Cornille (Bouches-duRhône, France), where large upright stones enclosed a skeleton.

In Portugal, excavations at Poças de São Bento (BaixeAlentejo) from the late 1950s and the 1980s display a similar pattern as that seen at Téviec and Hoëdic, with burials in a shell midden with two marine shells radiocarbon dated to the eighth millennium BP44 and charcoal also radiocarbon dated to the eighth millennium BP45 (Figure 4.19) (Morais Arnaud 1989). Two features identified in the shell midden are of interest to this discussion. First, the post-holes, as timber structures are characteristic of the early earthen long mounds of Scandinavia and Britain. Second, a number of ‘large stones’ have been found in the Poças de São Bento shell midden, which Larsson postulates may have been grave structures (Larsson 1996:130).

Moving to mainland Britain, there are examples of the use of stone for the construction of pavement or circles at Dunford Bridge A and B (South Yorkshire), Barsalloch Dumfries and Galloway), Culverwell (Dorset), Lussa Wood 1 (Argyll and Bute), March Hill Carr (West Yorkshire) and Staosnaig (Argyll and Bute).46 Dunford Bridge A and B (South Yorkshire) consists of charcoal, burnt flint and large flat stones that “provided a crude paving” (Radley et al. 1974:7) (Figures 4.22 and 4.23). Excavated in the 1960s, Barsalloch (Dumfries and Galloway) produced a large quantity of flint, evidence of fire, carbonised wood and stone settings radiocarbon dated to the late eighth and the second half of the seventh

In Spain, at the rock shelter at Colombres (Asturias), four stratigraphic levels have been identified, with level D consisting of greasy organic sediments, a large amount of shell, charcoal, animal bones and an Asturian lithic industry, dated to more than 10000 BP. Level D also contained a human skeleton located near the back of the shelter (Figure 4.20). Resting ten centimetres above bedrock, the skeleton was surrounded by “28 unmodified tubular limestone blocks”. The skull rested on a platform of five blocks (no information of their size) with another seven blocks encircling the skull (Clark 1983:32-3).

46. I have not included the flint floors from Golden Ball Hill (Wiltshire), which consisted of smooth flint pebbles, the largest some 10 by 15 metres, which preliminary excavations tentatively identified as ‘Mesolithic’ based on associated artefacts, including pieces of worked flint of ‘Mesolithic’ character (Anon 1999; Denison 1997b). Subsequently, more extensive excavation of the area identified that the ‘floors’ were part of a larger flint surface associated with ‘late Bronze Age’ pottery (Joshua Pollard personal communication). Similarly, Williamson’s Moss (Cumbria) has not been included as the stone pavement from Area B is associated with charcoal dated to 4238-4059 BP (.777) and 4051-3984 BP (.223) (3756±40 bp, BM-1396) which places it at the end of the ‘Neolithic’. While it has been argued on stratigraphic grounds that stone paving from test pits G46 and AZ26 pre-date 5000 BP (Bonsall et al. 1989:187), this has yet to be confirmed by radiocarbon dating.

44. The two calibrated radiocarbon dates are: 7640-7420 BP (1.000) (7040±70 bp, Q-2493); 7480-7240 BP (1.000) (6850±70 bp, Q-2495). Due to the marine context calibration used the marine curve (Stuiver et al. 1998). 45. The calibrated radiocarbon date is: 7740-7560 BP (.937) and 7540-7510 BP (.063) (6780±65 bp, Q-2494).

68

Figure 4.16 – Outline plans of Tumulus-St-Michel (Morbihan, France) (after Bradley 1998a: figure 17).

Figure 4.17 – The tertre tumulaire of Le Manio I (Morbihan, France) (after Scarre 1998: figure 3). 69

Figure 4.18 – Long mounds of the Carmac/Locmariaquer area (Morbihan, France). Note the presence in each of these monuments of an initial circular or oval core (after Scarre 2002a: figure 11).

Figure 4.19 – Compilation of areas excavated in the late 1950s and in the 1980s at Poças de São Bento (Baixo-Alentejo, Portugal) 1, the extent of the shell midden: 2, colouring left by post-holes; 3, pits; 4, large stones. Roman numerals indicate human skeletons (after Larsson 1996: figure 8). 70

Figure 4.20 – Colombres (Asturias, Spain): Burial and associated features (after Clark 1983: figure 3.14).

71

Figure 4.21 – Roc del Migdia (Cataluña, Spain): ‘Megalithic’ burial (after Bahn 1989a: figure 1).

Figure 4.22 – Plan of Dunford Bridge Site B (South Yorkshire). The dashed line marks the limits of the main flint distribution. The blackened area marks the hearth (after Radley et al. 1974: figure 4). 72

Figure 4.23 – Plan of Dunford Bridge Site A (South Yorkshire). The dashed line marks the limits of the main flint distribution. The stippled area was rich in burnt flints and charcoal fragments (after Radley et al. 1974: figure 5).

Figure 4.24 – Barsalloch: Plan of the excavated area with key to grid squares (after Cormack 1970: figure 2) 73

millennium BP47 (Cormack 1970) (Figure 4.24).

late eleventh to the ninth millennium BP.49 They have been compared to the constructions at Téviec (Morbihan, France). On a stratigraphically higher level than the ring constructions was a patch of flat stones. Work at March Hill Carr (West Yorkshire) has identified two well defined stone-lined hearths, the northernmost approximately 50 by 35 centimetres (Figure 4.31). Each hearth has produced two radiocarbon dates obtained from charcoal dating to the first half of the seventh millennium BP (Spikins et al. 1995, 2002).50 Recent excavations in 1994 at Staosnaig (Argyll and Bute) have identified a stone lining of 13 stones (feature F41) (Figures 4.32 and 4.33) (Mithen and Finlay 2000:384-6). The stones used in the construction are quartzite, mica schist and decayed sandstone. To the west of this feature is a narrow gully comprising about nine stones. The feature has yielded two radiocarbon dates acquired on single fragments of charred hazel nut shell from the lower fill and upper fill dating to the ninth millennium BP.51 The feature has been identified as a cist, which could have been used for cooking or storage, and has been compared to the features at Lussa Wood 1 (Argyll and Bute) (Mithen 2000d:615).

A more substantial stone pavement has been excavated on the Isle of Portland at Culverwell (Dorset) (Palmer 1999:Chapter 1.4 and 1.5). A shell midden with burnt and unburnt stone, along with ‘Mesolithic’ artefacts, dominates the midden, which covers the entire 1326 square metres. Finds from the shell midden consist of limestone, charcoal, ochre, bones, hearths and stone tools. The midden has an extensive stone pavement or floor covering 113.73 square metres, with the stones averaging 25-30 square centimetres and 5-6 centimetres thick. A stone “alignment of big, irregularly shaped nodules of both Portland and Purbeck stone” has also been identified (ibid.:21) and a semi-circular or circular feature marked with stones has also been excavated in the southern part of the midden (Figures 4.25, 4.26, 4.27, 4.28 and 4.29). The finds from Culverwell are the most substantial use of stone in the ‘Mesolithic’ of Britain and have ten radiocarbon dates ranging from the late ninth and early seventh millennium BP.48 At Lussa Wood 1 (Argyll and Bute) (Figure 4.30) on the Isle of Jura a set of three stone rings or circles in a scoop have been identified, dated to the ninth millennium BP (J. Mercer 1981; Searight 1984). Excavations have produced stone artefacts (including some 3000 microliths), red ochre, wood charcoal, burnt hazel nut shells, burnt animal bones and marine shell. The three continuous constructions are built mainly of bedrock slabs and charcoal from the base of the constructions dates to the

Richmond (1999:9) suggests that the use of stone found at Culverwell and Lussa Wood “represent early forms of monument amongst hunter-gatherer societies”. These examples, while rare, illustrate that structural elements attributed to the ‘Neolithic’ had foundations in earlier

49. The two calibrated radiocarbon dates are: 1015010140 BP (.001), 10110-10080 BP (.005), 10070-10060 BP (.001), 10030-10020 BP (.002), 10010-9990 BP (.002), 9940-9930 BP (.001) and 9920-8340 BP (.987) (8194±350 bp, SRR-160); 9400-9340 BP (.010) and 9300-8400 BP (.981) (7963±200 bp, SRR-159). 50. The four calibrated radiocarbon dates are: 6745-6733 BP (.016), 6733-6544 BP (.976) and 6509-6503 BP (.008) (5835±35 bp, OxA-6296); 6726-6724 BP (.002), 6724-6696 BP (.142), 6681-6541 BP (.844) and 65116502 BP (.011) (5824±28 bp, UB-4051); 6720-6700 BP (.084), 6670-6544 BP (.908) and 6507-6503 BP (.008) (5813±22 bp, UB-4051); 6717-6703 BP (.033) and 66676495 BP (.967) (5790±35 bp, OxA-6296). 51. The two calibrated radiocarbon dates are: 8700-8690 BP (.003), 8680-8670 BP (.007) and 8650-8410 BP (.990) (7780±55 bp, AA-21625); 8380-8180 BP (1.000) (7480±55 bp, AA-21626).

47. The calibrated radiocarbon date is: 7170-7160 BP (.281), 7110-7100 BP (.295), 6620-6610 BP (.247) and 6600-6590 BP (.178) (6000±110 bp, GaK-1601). 48. The ten calibrated radiocarbon dates are: 8110-7850 BP (1.000) (7525±60 bp, AA-28218); 7860-7620 BP (1.000) (7285±60 bp, AA-28217); 8280-8270 BP (.003) and 8200-7680 BP (.997) (7150±135 bp, BM-473); 77407490 BP (1.000) (7145±70 bp, AA-28216); 8150-8140 BP (.007), 8130-8120 BP (.001), 8110-8090 BP (.035) and 8060-7690 BP (.957) (7101±97 bp, BM-960); 75007240 BP (1.000) (6855±75 bp, AA-28220); 7420-7220 BP (1.000) (6800±60 bp, AA-28213); 7370-7150 BP (1.000) (6730±55 bp, AA-28214); 7180-6880 BP (1.000) (6525±60 bp, AA-28219); 7010-6740 BP (1.000) (6410±55 bp, AA-28215).

74

Figure 4.25 – Culverwell (Dorset): Plan of the floor and hearth 4, Area A and B (part) (after Palmer 1999: figure 6(1)).

Figure 4.26 – Culverwell (Dorset): Plan of the floor and features in Areas A and B (after Palmer 1999: figure 6(2)). 75

Figure 4.27 – Culverwell (Dorset): Plan of the floor in part of Trench 3 indicating a straight line of limestone slabs which, possibly, could have provided support for the footings of a superstructure, such as a shelter or ‘hut’ (after Palmer 1999: figure 10).

Figure 4.28 – Culverwell (Dorset): Parts of Area A and B (after Palmer 1999: figure 6(5)). 76

Figure 4.29 – Culverwell (Dorset): View of part of the floor in Area A with two straight lines of limestone slabs next to a square gap in the floor which could possibly mark the position of a superstructure (after Palmer 1999: plate 7). cases the wooden constructions took on a massive scale, similar to those of stone examples, as at Haddenham (Cambridgeshire) (Hodder 1992; Thomas 1991a). In other contexts, such as Street House (Cleveland) (Figures 4.34 and 4.35) (Vyner 1984) and Tustrup (Jutland, Denmark) (Joussaume 1988:41), both wood and stone were used.

periods. While the ‘Neolithic’ witnessed an explosion in the number and scale of such constructions, they can no longer be viewed as a distinctive attribute of the ‘Neolithic’ package.

Diversity through Continuity Contact between communities will have brought new ideas and influences, but the ways in which these were accepted or rejected will have varied. This is particularly evident in the monuments themselves; no two are exactly alike. (Scarre cited in Cullen 1995:382, also see Scarre 2002a)

Continuity in the Use of Large Constructions and Stone As the above survey indicates, the pre-‘Neolithic’ use of stone, though not common, was widespread. The identification of ‘megalithic’ structures seems to have little to do with stone size, and much to do with whether it dates to the presumed ‘Neolithic’. The search for a special ‘origin’ of the ‘megalithic’ in the ‘Neolithic’ is therefore inherently problematic. There is the question of what constitutes a ‘megalithic’ structure. Is its identification based on size of stones, as one might expect, or is it determined by its chronological context? Bahn’s (1989a:557-8) reluctance to label the use of stone at Roc del Migdia (Cataluña, Spain) as ‘megalithic’, was not so much based on the size of the stones used, but on the basis that the date of the structure was too early, as they are associated with a date of 13870-13140 BP (.995) and 13070-13050 BP (.005) (11520±150 bp, UGRA-117 RM1a). It is worth noting that Burenhult’s (1980, 1984, 2001) identification of the early ‘megaliths’ at Carrowmore (Co. Sligo), is made in relation to stones which are half the size of those identified by Bahn (Figure 4.36). Further, Carrowmore and Street House (Cleveland) (Figures 4.34 and 4.35) use stones that are smaller than many found in ‘Mesolithic’ contexts.

The long history of the use of stone is reflected in the diversity of the ‘Neolithic’ constructions. Throughout western and northern Europe from the ‘Neolithic’, many different constructions were built, not all of which incorporated stone. Many of the earliest constructions in Scandinavia and England consisted of earthen long mounds with or without internal timber structures. In France, the earliest constructions consisted of earthen long mounds with similar structural features (such as stone cists and hearths) to those of the shell middens at Téviec and Höedic (Morbihan). Thomas (1998a:49) has recently discussed the diverse nature of these constructions, ranging from long mounds, ‘megalithic burials’ and long ‘house’ architecture. Cullen points out that many ‘houses’ and ‘tombs’ utilise similar raw materials (wood, earth and stone) along with similar construction methods, remarking that “there is no sudden jump to exclusively megalithic tombs” (1995:382). In Britain, long mounds were constructed over wooden chambers (Case 1969; Thomas 1991a, 1999a), with later stone structures being constructed within similar mounds (Renfrew 1973; Thomas 1991a, 1999a). In some 77

Figure 4.30 – Lussa Wood 1 (Argyll and Bute): a - rings; b - cobbled patch (after J. Mercer 1981: figure 4).

Figure 4.31 – March Hill Carr (West Yorkshire), Trench A, Site 1 excavations in 1994 showing hearth 1 and 2 and the refit patterns (after Spikins et al. 2002: figure 2).

78

Figure 4.32 – Plan and profile of feature F41 at Staosnaig (Argyll and Bute) (after Mithen and Finlay 2000: figure 5.2.35).

Figure 4.33 – Feature F41 at Staosnaig (Argyll and Bute), following removal of fill showing stone lining (after Mithen and Finlay 2000: figure 5.2.37).

79

Figure 4.34 – Plan of the initial phase of the construction at Street House (Cleveland) (after Vyner 1984: figure 2).

Figure 4.35 – Plan of the final form of the cairn at Street House (Cleveland) (after Vyner 1984: figure 7).

Figure 4.36 – ‘Tomb’ No. 4 at Carrowmore (Co. Sligo) (after Burenhult 2001: page 19).

80

Figure 4.37 – The round dolmen at Skredsvik (Bohuslän, Sweden) (after Tilley 1996a: figure 4.23). contexts.52 Carrowmore (Co. Sligo) (Figure 4.36), Street House (Cleveland) (Figures 4.34 and 4.35), Storgård IV (Jutland, Denmark), Hindby Mosse (Skåne, Sweden), Hessland, Skredsvik (Figure 4.37), Lunden (Bohuslän, Sweden), all use stones which are comparable in size to pre-‘Neolithic’ contexts. The use of the term ‘megalithic’ exclusively in these contexts is wholly inappropriate if one looks at the early evidence, and is related to their existence in ‘Neolithic’ contexts.

It is true that the structures referred to in the ‘Upper Palaeolithic’ and ‘Mesolithic’ are not on the scale of the largest stone structures of western and northern Europe, such as Stonehenge, West Kennet and Avebury (Wiltshire), Newgrange and Knowth (Co. Meath), Maes Howe (Orkney Islands), Gavrinis (Morbihan, France), etc. However, this scale of ‘monumentality’ is actually the exception even in ‘Neolithic’ contexts. Those famous examples represent the peak of construction around 5000 BP immediately before the cessation of most similar stone construction (Burenhult 1993:80). These exceptionally large constructions are not the dominant scale of constructions down to the late fifth millennium BP, rather, if the evidence is thoroughly examined, the majority of constructions in post-‘Neolithic’ contexts are more comparable in size to those in pre-‘Neolithic’

52. As we have seen shell middens, particularly those at Téviec and Hoëdic, have been seen as being the inspiration for the earthen mound (Kirk 1993:212; Schulting 1996:347) and perhaps ‘passage-tombs’ (Scarre 1992, 1998, 2002; Scarre et al. 1993). Interestingly, work on Oronsay, Scotland at the turn of the century was carried out in the hope of discovering ‘Neolithic’ or Bronze Age burials in what was thought to be tumuluslike mounds. However, these mounds contained large shell middens (Woodman 1989:3).

81

about 6500 BP onwards was the materialisation of these elements on an increasingly massive scale, apparently as part of the long interaction between the various and differing subsistence groups of western and northern Europe. The varied resulting constructions illustrate the long history and varied influences such structural features had. What is needed, but is beyond the scope of this research, is a detailed examination of the background to ‘Neolithic’ stone, timber and earthen constructions to fully understand the stages that led to their later construction on a massive scale.

Conclusions Rather than seeking to dichotomise stone constructions on a ‘Mesolithic’/‘Neolithic’ basis, we need to note the evidence for development and continuity in the use of stone constructions from the ‘Upper Palaeolithic’ onwards. The constructions that fill the landscape of western and northern Europe from the beginning of the ‘Neolithic’, drew on elements of the earlier cultural assemblages. ‘Mesolithic’ populations were using stone, earth and timber for constructions. What happened from

82

CHAPTER FIVE: LIVING WITH THE DEAD, MAKING ANCESTORS for status display, but that such displays may occur in other contexts (see Bradley 1984a; Chapman 1991; Robb 1994; Whalen 1983).

With the exception of monument construction, the Neolithic is beginning to appear as an exceptionally uneventful period. (Mithen 1999a:478)

The identification of ideological interpretation in the postprocessual approach is problematic as the material remains of ideological processes may leave the same archaeological traces as other processes (McHugh 1999:16). The study of mortuary practices has witnessed varied approaches over the last twenty-five years, many of which have attempted to identify ‘laws’ or crosscultural similarities between cultures separated by time and space. As Whaley (1981:9) notes on attitudes to death and dying: it is futile to look for one common attitude to anything in any given society. The assumption that all human beings will think the same way at any time is one which our own experience of the present should indicate is not a valid one to impose on the past…bearing in mind that there is bound to be an almost infinite diversity both among individuals and also among groups in the same society. Despite this, archaeological explanation is generally phrased in terms of a single widespread attitude held by all individuals in the society (McHugh 1999:18). Generally, no matter which approach is taken, ethnographic analogy is used to interpret the burial practice. Many archaeologists argue that ethnographic evidence is necessary if one wants to interpret mortuary practices (Brown 1995; Chapman 1987; Gräslund 1994; Veit 1993). Amongst other things, they argue that archaeologists inevitably draw parallels from their own experience and knowledge to make sense of the past. In relation to mortuary practice, the danger of using ethnographic analogies is the declarations about what generally represent the ‘ideal’ procedures, which may rarely be closely followed in practice. Further, the ethnographic analogies drawn upon may not cover the variability present in the archaeological evidence. The limited nature of these observations means that they fail to account for the variability in the archaeological evidence (McHugh 1999:17, also see Chapter 3).

Chapters 2 and 4 have identified elements of ‘Neolithic’ human social behaviour that have been generally identified as distinctive to this period. This chapter will continue this reappraisal, focusing on the way that people interacted with dead individuals (human corpses) from the early tenth to the late fifth millennium BP. As with the previous attributes, the discourse generally argues that at the onset of the ‘Neolithic’ at the middle of the seventh millennium BP, attitudes towards the dead changed. This change is associated with a new concern for the ancestor materialised through the construction of stone chambered structures and timber and earthen mounds. The issue of these constructions has already been discussed (see Chapter 4). This chapter will illustrate continuity in treatment and deposition of human skeletal material between the ‘Mesolithic’ and the ‘Neolithic’ in order to argue that there is no empirical evidence to suggest that the human corpse was viewed radically differently during the ‘Neolithic’ and the ‘Mesolithic’, even though there is more evidence for deposits of the dead in the ‘Neolithic’.

Theoretical approaches to the Study of Mortuary Practice There have been various interpretative and theoretical approaches to the archaeology of mortuary studies over the last quarter of the twentieth century (see McHugh 1999:Chapter 1, also see Bartel 1982, Brown 1995). Some of these approaches have sought to devise crosscultural ‘laws’ drawn from a simplistic view of the ethnographic evidence to provide order and meaning to a varied mortuary assemblage through time and space. Mortuary assemblages have been viewed as an accurate reproduction of social organisation (Binford 1971; Saxe 1970, also see Tainter 1975, 1978). More recently, others have used historical contexts for explanation, focusing on ideological processes in mortuary remains (for example Barrett 1990; Bradley 1984a; Shanks and Tilley 1982; Shennan 1982; Whittle 1988a), working on from Hodder’s work (1984). These postprocessual approaches differ from the work of Saxe and Binford, as they do not view mortuary remains as fossilising a particular living social structure. Rather, the mortuary practices are characterised as processes that create and modify social structures. For these researchers, burial is an arena for political dominance through ancestral association, helping to create a particular social structure (see Hodder 1990a; Renfrew 1973; Shennan 1982; Shanks and Tilley 1982; Sillar 1992; Taylor 1989, also see Block 1971). These approaches argue that burial need not be a context

What my analysis of the treatment, patterning and deposition of human skeletal material in Britain and Ireland from the early tenth to the late fifth millennium BP will illustrate is that previous studies of mortuary practice have inherently taken too narrow an approach (see Barrett 1990; Binford 1971; Bloch 1971; Bradley 1984a, 1995b; Bristow 1998; Chapman 1995; Cooney 1992; Criado Boado 1989; Hodder 1982b, 1982c, 1984, 1990a; Lucy 1992; McHugh 1999; Mizoguchi 1993, 1995; O’Shea 1984; Parker Pearson 1982; Saxe 1970; Shanks and Tilley 1982; Shennan 1982; Tainter 1975, 1978; Thomas 1991a, 1999a, 2001a; Veit 1993; Whittle 83

ascribed to the ‘Iron Age’53 (Saville and Hallén 1994). This has serious implications for how we proceed. A number of chambered structures across Scotland and Ireland have had ‘Iron Age’ artefacts excavated from them (see Henshall 1963, 1972a; Hingley 1996, 1999; Shee Twohig 1981). It is nonetheless assumed that these structures were constructed and initially used in the ‘Neolithic’ for the purpose of depositing human skeletal material, as has been demonstrated by radiocarbon dating in some cases. However, the possibility exists that the human skeletal material could have been deposited in later periods.

1988a, 1996a), which has obscured unparalleled forms of human social behaviour.

The Evidence: Human skeletal material and dating Attempts to study human skeletal material in the British and Irish ‘Mesolithic’ just ten years ago would have been limited to a handful of contexts: Gough’s New Cave, Aveline’s Hole (Somerset), the Oronsay (Argyll and Bute) shell middens and perhaps Thatcham III (Berkshire). However, extensive radiocarbon dating projects by Woodman, Chamberlain, Richards and Schulting has identified many more ‘Mesolithic’ contexts with human skeletal material from the early tenth to the middle of the seventh millennium BP (see Figure 5.1). The majority of dates have come from caves and rock shelters where complex formation processes are at work; generally meaning that the chronology of human skeletal material is difficult if based solely on stratigraphy. A cave may contain deposits and assemblages of many different periods, for example from the ‘Mesolithic’, ‘Neolithic’ and ‘Iron Age’ (see Chamberlain and Williams 2000a, 2000b, 2001). Thus, human skeletal material must be radiocarbon dated to determine (as far as that is possible with current methods) that it dates to a certain period.

Because of this and other factors already mentioned (see page 25), all arguments I make about human skeletal material are based only on those contexts that contain radiocarbon dated human skeletal material. In doing this I am assuming that this dates all the human skeletal material from the context unless earlier (pre-tenth millennium BP) or later (post-fifth millennium BP) dates for human skeletal material have been obtained from the context. While this is sometimes only based on one human bone and one radiocarbon date, this assumption is necessary. This practice thus excludes archetypal ‘Neolithic’ contexts, such as Newgrange (Co. Meath), Maes Howe (Orkney Islands), Wayland’s Smithy (Oxfordshire) and Mount Pleasant (Dorset). The decision was taken in order to ensure, as far as one is able, that the contexts discussed had human skeletal material dating from the early tenth to the late fifth millennium BP (Figure 5.2). These contexts will be the basis of discussions relating to arguments about the treatment and deposition of human skeletal material and are presented in full in the Online Database (see Appendix 6).

The danger of dating human skeletal material by association and stratigraphy only, rather than by radiocarbon dating, has been demonstrated in a number of contexts. For example, MacArthur Cave (Argyll and Bute) has been identified as a classic Obanian culture context, with many bone and antler implements being excavated; including an antler biserial barbed point radiocarbon dated to 7670-7460 BP (.952) and 7450-7430 BP (.048) (6700±80 bp, OxA-1949). Human skeletal material from at least four individuals has also been excavated, some of which had been deposited in a marine shell deposit. These individuals were generally assumed to date to the ‘Mesolithic’ due to the associated Obanian artefacts. However, four samples of human skeletal material produced radiocarbon dates placing them in the third millennium BP showing that they should be

Taphonomic Considerations Taphonomic factors are important considerations when dealing with any archaeological assemblage and/or context. This is particularly so when studying skeletal material, especially that of a fragmented nature. The Russian palaeontologist Ivan A. Efremov first termed taphonomy from the Greek taphos (burial) and nomos (laws), and defined it as: 53. The four calibrated radiocarbon dates are: 2710-2630 BP (.263) and 2620-2360 BP (.737) (2460±55 bp, OxA4487); 2710-2630 BP (.169), 2620-2580 BP (.040), 25702560 BP (.003), 2540-2520 BP (.015), 2510-2310 BP (.743), 2240-2210 BP (.024) and 2190-2180 BP (.005) (2365±55 bp, OxA-4486); 2470-2410 BP (.049), 24002390 BP (.015), 2380-2150 BP (.926) and 2140-2120 BP (.010) (2295±60 bp, OxA-4488) and 2330-2040 BP (.975) and 2030-2010 BP (.025) (2170±55 bp, OxA4485).

84

Figure 5.1 – Distribution of radiocarbon dated human skeletal material from British caves (after Chamberlain 1996: figure 1).

85

Figure 5.2 – Map generated from Online Database showing the distribution of all contexts with radiocarbon obtained from samples identified as human skeletal material (excluding loose teeth). The general taphonomy of bone has received considerable attention (for example, Binford and Bertram1977; Denys 2002; Lyman 1987; Maltby 1985; Marshall 1989; Mays 1991, 1994, 1998; Mays and Andersen 1996; Micozzi 1991). Many variables and influences, both natural and cultural, are at work on skeletal material (Evans and O’Connor 1999; Lyman 1987; Maltby 1985; Schiffer 1976, 1987). Natural factors include soil acidity, post-depositional action of roots, biological action, exposure to heat or sunlight, animal damage, fluvial transport or animal transport, as well as random processes (Andrews and Cook 1985; Behrensmeyer 1978; Binford and Bertram 1977; Canti 2003; Davidsson 1998; Denys 2002; Garland and Janaway 1989; Haynes 1990; Henderson 1987; Littleton 2000; Miller 1975; Noe-Nygaard 1987; Waldron 1987).

the study of the transition (in all its details) of animal remains from the biosphere into the lithosphere, i.e. the study of the process in the upshot of which organisms pass out of the different parts of the biosphere and being fossilised become part of the lithosphere. (1940:85) While it was first developed in relation to animal remains, taphonomy can be more broadly defined to cover all organic and inorganic material, as the postmortem relations between the deposited remains and their external environment (Hiscock 1990; Noe-Nygaard 1987:7). In relation to this research, I am concerned with what happened to skeletal material prior to deposition and after being deposited in or on the ground. 86

least part of each body was burnt away, removed from the site, or circulated within the wider, ongoing daily routine. (Fowler 2001:160)

Armour-Chelu and Andrews (1994) have illustrated that a single deep-burrowing earthworm (Lumbricus terrestris) dispersed skeletal material of mice and shrews through a 4.5 litre volume of soil in two years. Cultural processes include human disturbance and excavation (Bennett 1999; Clarke 1978; Garland and Janaway 1989); all contribute to produce archaeological assemblages. While taphonomic processes have generally been portrayed as destructive and resulting in the loss of information (see Gifford 1981:400; Raup and Stanley 1978:296), they can also identify cultural practices as noted above (Hiscock 1990). For example, weathering and gnaw marks on skeletal material can be inferred as indicating exposure. Ultimately, the question is whether the patterning identified in the empirical evidence is due to positive human activity in the past – humans as a depositional agent – or to other taphonomic processes of re-deposition and post-deposition. Unfortunately, taphonomic issues have only just begun to be discussed in some detail for the prehistoric deposition and dispersal of the dead in Britain and Ireland (see Barber 1997; Baxter 1999; Brück 1995; Carr and Knüsel 1997; Hill 1995; West 1996).

Even if one were to include all human skeletal material assumed to belong to the ‘Neolithic’, this would still only amount to a fraction of all those who died in Britain and Ireland between the middle of the seventh and the late fifth millennium BP. This missing component is either yet to be excavated or has not survived, though of course we have to work with what is available to us. Further, the known archaeological contexts must only represent a fraction of the possible locations available in the past for depositing human skeletal material. The dead that are represented in archaeological assemblages from excavated contexts were subject to a variety of treatments, including exposure, burning and disarticulation (Figure 5.3). After death, the living populations had a number of options open to them in the treatment and disposal of a corpse, not all of which would be identifiable in archaeological assemblages. The treatment that a human corpse was subjected to may have been influenced by the cause of death, the age of the corpse (see Liddle 1998:42-3; Thomas and Whittle 1986), the completeness of the corpse, whether or not the population could recover the corpse and the time of year. Very specific, local social rules may well have applied which would preordain what, if any, action was to be carried out at a particular place and time. However, and critically, these social rules cannot be assumed to have been constant across Britain and Ireland and over several thousands of years. But these variables are lost once items of a cultural assemblage leave the systemic context and enter the archaeological context (see Schiffer 1972, 1976). The most common response has been to try to recover this ‘lost’ component of information through the use of ethnographic analogy. However, as outlined throughout this research, drawing analogies between archaeological assemblages and ethnographic evidence is likely to limit the scope of the human social behaviour that can be identified. Rather, archaeologists should seek to identify human social behaviour through examining archaeological assemblages and by identifying patterns in the material first.

The Use of Published Archaeological Data Not all the excavation literature is of the quality one would wish. Excavations from the nineteenth and early twentieth century were carried out when techniques were not ideal for the retention of human skeletal material. Human skeletal material reports are frequently not extensive, generally under three pages, if published at all. There has been a widespread failure in the past to see the archaeological significance of excavated human skeletal material as a source of archaeological information. This is particularly true in the British and Irish ‘Mesolithic’ and ‘Neolithic’ because it generally amounted to disarticulated, fragmentary and incomplete human skeletons (Larsen 1997:1-3; Wysocki and Whittle 2000:591). A geographical bias is also present, with more contexts being identified in certain regions, such as Wessex (which includes the English counties of Somerset, Avon, Dorset, Wiltshire, Hampshire and Berkshire, Cunliffe 1993:1). While this may indicate more contexts in these regions than others, it may also be due to the concentration of high quality archaeological research in these regions. Due to these variables, most of the evidence I use relating to the treatment of human skeletal material will be limited to those excavations that have made detailed analyses of the human skeletal material assemblage.

A number of scenarios will be put forward illustrating the possible processes following death, which can be identified in archaeological assemblages of human skeletal material. While the argument will be made in relation to humans and draw on examples from the Online Database (see Appendix 6), the examples could equally be made for other animals. Throughout this book, contexts and human social behaviour will be identified where there is no empirical reason to differentiate between human and non-human species. This leads to an argument about discussing skeletal material in general rather than differentiating skeletal material based on species type.

Treatment and deposition of human skeletal material in the ‘Neolithic’

Death and Deposition: a many varied thing

But crucially, the ‘body’ was never laid to rest. Bodies and body parts were often added too. At

The vital signs of life have ceased, what next? The 87

Figure 5.3 – An archaeological view of the processes that human and animal skeletal material can undergo (after Kinnes 1992b: figure 2.5.3).

Category of human skeletal material skulls articulated disarticulated overall rates

% weathered

% root-marked

80 71 49 54

100 71 15 60

% heavily weathered 18 21 47 17

% heavily rootmarked 12 14 10 11

Table 5.1 – Percentages of weathered and root-marked human skeletal material within categories of types of deposit after McKinley forthcoming). 88

main enclosure at Hambledon Hill Complex (Dorset) may have been “a gigantic necropolis constructed for the exposure of the cadaveric remains of a large population” (1980:63). The spatial constraint of some deposits, including human skeletal material, along the bottom of the ditch lead Mercer to suggest that they had been placed in a series of bags (ibid.:30).

complete corpse could be placed in a pit, ditch, mound or a stone chambered structure. Examples of fully articulated human skeletons have been identified in chambered structures, which have produced very early dates with human skeletal material dated to the later part of the seventh millennium BP: Hazleton North (Gloucestershire) and Whitwell (Derbyshire). Articulated skeletons have been identified in pits from Pitdrichie Farm 1 (Keabog 1) (Aberdeenshire), Alington Avenue (Dorset), Willington Plantation Quarry (Bedfordshire), Lough Gur C and D (Co. Limerick), Martinstown (Kiltale) (Co. Meath), Gairneybank 3 (Perth and Kinross), Boatbridge Quarry 2 (South Lanarkshire), St. Margaret's-at-Cliffe (Kent), Barrow Hills Pit 4660, Mount Farm (Oxfordshire) and South Dumpton Down (Kent). Articulated skeletons have also been identified in pits beneath mounds from Risby (Suffolk), Shrewton 24 (Wiltshire), Black Patch 3 (East Sussex), Park Farm (Berkshire), Barrow Hills ‘Barrow’ 4A, Barrow Hills ‘Barrow’ 17 (Oxfordshire) and West Cotton 6 F3259 (Northamptonshire) and from a ditch at Willie Howe (Humberside) and from a wet deposit at Castle Balkeney (Gallagh) (Co. Galway). In these cases that have been identified in the literature as articulated human skeletons, small hand and feet bones may be missing. In the literature, this is generally assumed to be due to natural processes. However, the possibility must be entertained that these bones may have been missing at death or removed by humans sometime after death.

Exposure of human skeletal material could have been carried out on a platform, which would have ruled out soil processes modifying the bones and would have limited the animal influences mainly to birds (Figure 5.4) (T. Pollard 1999; Scott 1992). The recent publication of excavations from Parc le Breos Cwm (Swansea) identified weathering on a number of bones using the weathering stage criteria outlined by Lyman (1994) (Whittle and Wysocki 1998:153). Animal modification was also identified at Parc le Breos Cwm, being attributed mainly to large mammalian carnivores/omnivores, and possibly herbivores, with some rodent modification, a pattern also seen at Carsington Pasture Cave (Derbyshire) (Chamberlain 1999). Whittle and Wysocki (1998:157-8) highlight the evidence of a carnivore-scavenged corpse exposed for 22 days in rural woodland near Seattle in the United States of America (Figure 5.5). The evidence shows the quick reduction of the articulated skeleton, which over longer periods is more severe (Figure 5.6). It is further reported that in human skeletons exposed for up to a year, disarticulation is complete except for the bones of the vertebral column (Haglund et al. 1989). Willey and Snyder (1989) report that two 18-27 kilogram deer fawns were consumed in 24 hours by five timber wolves with no identifiable bones remaining. These reported contexts are quite specific and many variables are at work, but they illustrate the effect which animal modification can have in a short time span, and suggest that measures may have been introduced in the past to limit the modification of the corpse by animals.

Alternatively, after death, living individuals could have defleshed the corpse and reduced the corpse to individual skeletal elements of various sizes. Disarticulation could have been achieved by exposing the corpse to natural elements and animals. This could have occurred within a ditch, a pit or in a cave or rock shelter. Evidence from Windmill Hill (Wiltshire) suggests that natural processes of weathering led to the filling of the ditches, with evidence that humans intervened to hasten the process (Whittle et al. 1999c:353-4). At Hambledon Hill Complex (Dorset) fine-grained silt was identified around the skull, indicating that the skulls had lain exposed prior to the ditches being filled (Andersen 1997:253-4). The recent re-examination of the human skeletal material from Hambledon Hill Complex by McKinley identified weathering and root-marks on much of the assemblage (Table 5.1) (McKinley forthcoming). Evidence of weathering of human skeletal material has also been identified at Windmill Hill (Wiltshire) (Smith 1965a:137).

The forthcoming excavation report for Hambledon Hill Complex (Dorset) includes a detailed study of the presences of cut marks on the human skeletal material (McKinley forthcoming). The analysis recorded cut marks from 23 bones, about 7 percent of the assemblages. The cut marks were identified on five articulated skeletons (36% of articulated material), three skulls (21% of skulls) and 15 disarticulated fragments (5% of disarticulated material), with radiocarbon dates being obtained for some of this assemblage (Table 5.2). Cut marks were most commonly recorded in skull (43%) and femur (35%), but also on the ischium, ilium, rib, radius, fibula and tibia. Individuals across all age ranges (neonate to older adult) and both sexes were affected. Human skeletal material with cut marks covered all of the main localities, with no significant concentration apparent within any particular area. The cuts were produced by human action using stone tools. The cut marks were all of a similar form and type, commonly occurring as small groups of parallel or almost parallel lines. Most of the marks were extremely shallow, though there were several slightly deeper incisions. In most instances, relatively few cuts (between 1-15) were evident in one skeletal element,

This process means that human skeletal material had been left in open ditches, pits, caves or rock shelters, exposing them to weathering, soil processes and animals which would have reduced the corpse from an articulated state to disarticulated skeletal elements. The assemblage of human skeletal material from Giants’ Hills 2 (Lincolnshire) showed evidence of breakage, weathering and modifications by animal actions (Evans and Simpson 1991:16), while at the chambered structure at Isbister and Quanterness (Orkney Islands) some human skeletal material showed evidence of weathering (Barber 1988; Richards 1988). Roger Mercer has suggested that the 89

Figure 5.4 – Suggested structure, use and two phases of decay of a three-point support raised platform (after Scott 1992: figure 8.7).

90

Figure 5.5 – Human corpse found in a rural woodland area after being exposed for 22 days (after Haglund et al. 1989: figure 4).

Figure 5.6 – Human corpse found in a rural woodland area after being exposed for 2½ months (after Haglund et al. 1989: figure 5).

91

Find Number HH76 2625

Age/Sex young/younger mature adult male

Type Articulated

ST78 2756

Neonate

Articulated

ST79 2726

older mature adult Articulated male

ST78 2755

older subadult/ young adult male

Articulated

ST80 1875

young adult male

Articulated

Absolute Dates 5594-5449 BP (.841), 5382-5327 BP (.159) (4765±45 bp, OxA-7773); 5584-5501 BP (.394), 5491-5446 BP (.218), 5413-5324 BP (.388) (4725±40 bp, OxA-7774) (replicate of OxA-7773) 5609-5567 BP (.333), 5557-5470 BP (.667) (4815±35 bp, OxA-7101) 5584-5501 BP (.544), 5491-5451 BP (.226), 5379-5329 BP (.230) (4738±28 bp, UB-4242) 5450-5380 BP (.098), 5330-5040 BP (.898), 5010-5000 BP (.004) (4560±55 bp, OxA-7044); 5580-5510 BP (.105), 5490-5280 BP (.843), 5160-5130 BP (.024), 5110-5070 BP (.028) (4645±60 bp, OxA-7045) (replicate of OxA-7044) 5567-5557 BP (.027), 5470-5426 BP (.237), 5424-5318 BP (.737) (4679±27 bp, UB-4243)

Table 5.2 – Radiocarbon dated skeletal material associated with cut marks (after McKinley forthcoming). 5.8, 5.9, 5.10, 5.11, 5.12, 5.13, 5.14, 5.15, 5.16, 5.17 and 5.18). The analysis is undertaken with reference to literature relating to supposed evidence for cannibalism (anthropophagy) in the Americas and Europe and the evidence of butchery of animals. The human skeletal material had evidence of a variety of surface modifications (Table 5.3). Cut marks are found on 57 finds (16.9% of the total assemblage) and occurred throughout the complex and at various phases (Table 5.4). The cut marks occur on a variety of anatomical parts and occur within all age categories (Table 5.5).

however, in excess of 100 cuts were recorded along the length of the anterior shaft of the left femur (HH76 2625). Of particular interest, McKinley states that the cuts were characteristic of those produced in the butchery of animals during defleshing. They are the “shorter, more oblique cuts made to the underside of the exposed bone to free it from the mass of meat and/or sever muscle insertions” (Binford 1981a:129) and are characteristically “very superficial, rarely deep” (ibid.:131). Some of the marks resulted from the action of scraping flesh off the bone as part of a cleaning process. This action, where undertaken by a highly skilled operative may leave no marks on the bone at all. In the majority of cases, the cuts can be identified as the result of either the severing of specific muscle attachments and/or cleaning/stripping the bone of its soft tissues by repeated short scrapes, possibly accompanied by pulling away of the tissue. The identification of cut marks from Hambledon Hill Complex is of interest as it was overlooked in the original analysis. With much of the analysis of human skeletal material assemblages occurring when such analyses were in its infancy, a re-assessment of at least the bestpreserved assemblages seems appropriate and necessary (McKinley forthcoming).

Evans (2000:78) concludes that the position of the cut marks on cattle skeletal material “can be directly compared to the position of cuts associated with butchery practices”. Further to this she states “that all forms of butchery processes were carried out” (2000:80). Comparing the cattle evidence with the human, she concludes that the cut marks “on the human bones have a different objective behind them” related to their totally different position on the skeleton to those on the cattle (2000:82). Evans argues that due to the lack of the six criteria that must be met if cannibalism is to be inferred, the human skeletal material from the Hambledon Hill Complex was not processed for meat (see pages 146-7). Rather, due to the predominance of randomly positioned cut marks over the surface of long bones and the skull, the cut marks are the “result of defleshing and disarticulating the remains as part of Neolithic funerary practices” (2000:83). Further, in support of this conclusion, Evans states that gnawing marks, particularly

Evans (2000) employed various methods, including analysis using a stereo microscope and a Scanning Electron Microscope (SEM), to compare the position and form of cuts on human and cattle skeletal material excavated from Hambledon Hill Complex (Figures 5.7, 92

30 25 20 15 10 5

skel

carpals

sternum

mc

calc

clav

pelvis

talus

patella

scapula

phalanges

tarsals

ulna

rad

vert

max

hum

mt

rib

tib

fib

fem

mand

skull

0

Figure 5.7 – Anatomical distribution of human remains from Hambledon Hill (Dorset), shown as a percentage of the total bones present (after Evans 2000: figure 7.02).

pelvis

scap

ulna

hum

tib

fib

mand

fem

skull

40 35 30 25 20 15 10 5 0

Figure 5.8 – Graph showing the anatomical distribution of both articulated and disarticulated cut human bone, displayed as a percentage of the total number of cut bone (after Evans 2000: figure 7.10).

25 20 15 10 5 ulna

axis

slull

femur

calcaneus

mandible

metacarpal

astragalus

tibia

radius

pelvis

metatarsal

humerus

scapula

0

Figure 5.9 – Anatomical distribution of cut marked cattle bone from Hambledon Hill (Dorset), expressed as a percentage of the total number of bones recorded (after Evans 2000: figure 7.21).

93

Figure 5.10 – Fine cut marks appearing under magnification, parallel to the main cut marks noted during recording (human skull - 682, 725 + 1037, HH77) (after Evans 2000: figure 7.14).

Figure 5.11 – Patterns of parallel cut marks on a human mandible, seen under magnification using the stereo microscope (1916, HH76) (after Evans 2000: figure 7.15).

94

Figure 5.12 – Scanning electron microscope of a cut mark illustrating the weathered ‘rounded’ appearance of the cuts, not clearly visible through the naked eye (human scapula - 2919, ST78) (after Evans 2000: figure 7.16).

Figure 5.13 – Well defined cut marks on a cattle metacarpal (1470, HH75) (after Evans 2000: figure 7.23).

95

Figure 5.14 – Well defined cuts on a cattle skull (1613, HH75) (after Evans 2000: figure 7.24).

Figure 5.15 – Cuts across the distal articulation on a cattle metatarsal (2346, ST79) (after Evans 2000: figure 7.25).

96

Figure 5.16 – Parallel cuts on a cattle mandible, made clear under magnification (708, HH75) (after Evans 2000: figure 7.26).

Figure 5.17 – Scanning electron microscope photograph showing the well defined ‘V’ shaped cut marks on a cattle metatarsal (2346, ST79) (after Evans 2000: figure 7.27).

97

Figure 5.18 – Scanning electron microscope photograph showing well defined ‘V’ shaped cuts on a cattle tibia (266, HH74) (after Evans 2000: figure 7.28).

Surface Modification Good condition Moderate condition Poor condition Gnawed Cut Chopped

Articulated remains 46.2 30.7 23.1 15.4 53.8 0

Disarticulated remains 5.6 84.8 9.6 4.1 16 0

Table 5.3 – Surface modification of the human skeletal material, expressed as a percentage of the total assemblage (after Evans 2000:figure 7.3).

Phase I II II / III III IV V VI VII VIII VIIIa Neo

Main enclosure 21.4 14.3 3.6 25 7.1 28.6 -

Stepleton enclosure 38.5 53.8 7.7 -

Long barrow 100 -

Hanford spur 100 -

Shroton outworks 100 -

EXD 66.7 33.3 -

Table 5.4 – Distribution of cut marked bone by location and phase, expressed as a percentage of the total for each individual location (after Evans 2000:figure 7.9). 98

Age category Neonate Infant Juvenile Sub-adult Sub-adult / adult Adult

Cut bone 1.7 7 19.3 12.3 12.3 47.4

Table 5.5 – Distribution of cut bone within age categories, shown as a percentage of the total number of cut bone (after Evans 2000: figure 7.11). forthcoming). A skull fragment with cut marks in the stone structure at Carrowmore 51 (Co. Sligo) has been radiocarbon dated to the sixth millennium BP54 (Burenhult 2000:181) and a human scapula with cut marks from Charterhouse Warren Farm Swallet (Somerset) has been radiocarbon dated to the second half of the fifth millennium BP55 (Hedges et al. 1989; Levitan et al. 1988). From Eden Walk II (Greater London) occupation debris, including animal bones, worked antler, pottery and worked flint has been excavated. Human skeletal material has also been identified, consisting of disarticulated skull, femur and clavicle bones (Bird et al. 1980; Hedges et al. 1993b; Serjeantson et al. 1992). From examinations of the human femur, which has been radiocarbon dated to the late fifth and the early fourth millennium BP,56 it was concluded that marks “showed a consistent orientation at three points of muscle attachment, suggesting they were the result of defleshing” (Hedges et al. 1993:312).

on articulated skeletons…, and the weathered appearance of the cuts, implies that individuals were left lying on the surface for some time before burial. Such weathering is not as extensive on the animal bones, suggesting that they were deposited into the ground soon after they were finished with. (2000:84) The identification that the processing of the human skeletal assemblages defleshed the skeleton indicates that pieces of human flesh were removed from the skeleton and what happened to this after the defleshing is anyone’s guess. Despite the systematic identification of different procedures in processing the cattle and human skeletons, it does not necessarily follow that the resulting flesh was used for different purposes. The pattern identified is a significant one but not necessarily for the reasons that Evans concludes. The different treatment that the human and cattle skeletons were exposed to does not rule out the possibility that the resulting human and cattle flesh served similar functions. One is not in the position, with the evidence available, to identify categorically that cattle flesh served one purpose, namely subsistence, and the human another, namely funerary/mortuary. One could, with the evidence for gnawing marks, argue that the human skeletal material, both articulated and disarticulated, was carelessly and haphazardly disposed of, being left to scavenging animals, while the other animals were buried soon after it was defleshed and disarticulated.

Recent excavations from Carsington Pasture Cave (Derbyshire) unearthed a large quantity of skeletal material, including human, cattle, pig, dog, horse, sheep, goat and aurochs, along with a bone pin, a fragment of worked antler and a flint flake. The human skeletal material represented a minimum number of 20 individuals, with radiocarbon dating of two human bone samples indicating that one calibrated to the fifth millennium BP and another to the third millennium BP

Harman and Evans (1991) noted one possible cut mark on a fragment of humerus shaft from the Giants’ Hills 2 (Lincolnshire), but describe neither the cut nor its location any further. A fragment of right distal humerus from Haddenham (Cambridgeshire) had four groups (415 cuts) of short, parallel cut marks in the anterior, medial and posterior surfaces in the areas of the brachialis and triceps attachments. One other cut was noted in this assemblage, a long cut of about 5 mm in a fragment of zygoma (Baxter 1999:7; McKinley

54. The calibrated radiocarbon date is: 5580-5530 BP (.052), 5490-5260 BP (.801), 5190-5120 BP (.075) and 5110-5050 BP (.071) (4625±60 bp, Ua-11581). 55. The calibrated radiocarbon date is: 4400-4390 BP (.009), 4380-4370 BP (.018) and 4360-3990 BP (.973) (3790±60 bp, OxA-1559). 56. The calibrated radiocarbon date is: 4090-4030 BP (.058) and 4020-3630 BP (.942) (3550±85 bp, OxA2924).

99

(Chamberlain 2001b).57 The human bone assemblage was almost entirely devoid of smaller bones, with only a foot bone and a few vertebrae being recovered, a pattern that is also seen in the animal bone assemblage. Due to the excellent preservation of material excavated, this could indicate that the disarticulation process occurred outside the cave or in as-yet unexcavated regions of the cave system (Chamberlain 1999). Further, Chamberlain argues that movement of fragments of human skeletal material occur within the cave system, “demonstrated by the spatial distribution of parts of a single skull, which were separated by a distance of 3.5 m within the second chamber” (Chamberlain 1999:Online). Multiple v-shaped cut marks have been identified on two femurs from an adult individual, with their location consistent with disarticulation at the knee joint. A similar pattern of cut marks has been identified on the animal bone assemblage, which has been identified as associated with “butchery or other processes of food preparation or secondary product utilisation” (ibid.). This interpretative issue of identifying butchery and disarticulation will be discussed further below (see pages 146-7).

locations or be subjected to further processes. The remains could have been broken down into smaller skeletal elements or been burnt or cremated. Cremated human skeletal material, to varying degree, has been identified at Bath Gate (Gloucestershire), Barford Farm, Flagstones House, the Hambledon Hill Complex (Dorset), Gravelly Guy, Ascott-under-Wychwood, Dorchester Cursus, Barrow Hills Pit 4700, Barrow Hills Pit 919, 950, Barrow Hills Pond ‘Barrow’ 611, Barrow Hills Pond ‘Barrow’ 4866, Barrow Hills ‘Barrow’ 12 (Oxfordshire), Crarae (Argyll and Bute), Loanleven, Almondbank, North Mains Henge (Perth and Kinross), Llandegai A, Llandegai B (Gwynedd), Durrington Down ‘Barrow’ 7, Millbarrow, West Kennet, Stonehenge (Wiltshire), Annagh (Co. Limerick), Straid (Co. Derry), Kilgreany (Co. Waterford), Mill Road Cist (West Lothian), Sand Fiold, Isbister (Orkney Islands), Oldtown 1,2,3 (Co. Kildare), Keenoge (Co. Meath), Carrowmore ‘Tomb’ 51, Primrose Grange ‘Court Tomb’ 1 (Co. Sligo), Embo, Tulach an t’Sionnaich (Highland), Dillonsdown (Co. Wicklow), Jerpoint West (Co. Kilkenny), Knockmaree (Phoenix Park) (Co. Dublin), Grange 2 and 3 (Co. Roscommon), Moneen (Co. Cork), Poulawack (Co. Clare), Parc le Breos Cwm (Swansea), Leuchar Brae (Skene) (Aberdeenshire) and Collessie (Fife).58 Of all the various ways that a corpse can be treated, disarticulated is by far the dominant form of human skeletal material in the British and Irish ‘Neolithic’ (see the Online Database and Appendix 6), a pattern identified by Thorpe (1984) in relation to long mounds in Wessex (Figures 5.19, 5.20 and 5.21).

Disarticulated elements of one body may have been deposited with other skeletal material from another individual. Fowler (2000, 2001) has recently highlighted a number of interesting deposits of different human skeletal material involving 2 or more human skeletons. Within chamber seven at Ascott-under-Wychwood (Oxfordshire), an almost complete human skeleton has been found with the atlas vertebra from a second human skeleton inserted into the backbone. As Fowler states “essentially this skeleton was given an ‘extra’ vertebra” (2001:143). In chamber one, a vertebra was placed within a smashed skull, ulnae were placed in articulation with radii from other bodies, and intact humeri and femurs were broken and then put back together (Chesterman 1977:26). At the chambered structure at Lanhill (Wiltshire) at least two skulls were found articulated with a mandible that did not belong to the skulls (Pollard 1998, 2001:143; Keiller and Piggott 1938:125, 127). At Ballaharra (Isle of Man) excavations seem to reveal a crushed crouched human skeleton, however, it “proved on laboratory examination to consist of representative parts of three individuals – a fully adult male, a youth of about 19 years, and a boy of 10-11 years” (Cregeen 1978:148, also see Pollard 1998, 2001:151).

Location of human skeletal material Human skeletal material radiocarbon dated and calibrated to between the middle of the seventh and the late fifth millennium BP can be found in virtually all archaeological contexts identified in the archaeological literature, including pits (Figure 5.22), ditches (Figure 5.23), cursuses, enclosures (Figure 5.24), causewayed enclosures (Figure 5.25), occupation debris (Figure 5.26), henges (Figure 5.27), caves (Figure 5.28), rock shelters

58. A new AMS dating technique allows accurate and precise dates from the bio-apatite in cremated bone. The work has been carried out at the Groningen laboratory on over 400 determinations on material from the Netherlands, Ireland, Britain, Germany and Belgium (Lanting and Brindley 1998, 2000; Sheridan 2002). With this technique being in its infancy, few contexts exist in the study area with radiocarbon dated cremated human skeletal material. Most of these contexts are included as they have dated articulated or disarticulated samples.

The human skeletal material left behind after the various actions above were carried out, which could be devoid of small bones, might then be deposited in a variety of 57. The two calibrated radiocarbon dates are: 4783-4768 BP (.009), 4612-4597 BP (.010) and 4574-4240 BP (.980) (3980±60 bp, OxA-9930); 2711-2627 BP (.253), 2620-2556 BP (.142) and 2550-2350 BP (.604) (2435±55 bp, OxA-9806).

100

Figure 5.19 – State of human skeletal material at long mounds in Wessex (after Thorpe 1984: figure 4.6).

Figure 5.20 – State of human skeletal material identified as adults at long mounds in Wessex (after Thorpe 1984: figure 4.11).

101

Figure 5.21 – State of human skeletal material identified as children at long mounds in Wessex (after Thorpe 1984: figure 4.12). (Figure 5.29), wet contexts (Figure 5.30),59 shell middens (Figure 5.31), stone chambered structures (Figure 5.32), cists (Figure 5.33), long mounds (Figure 5.34), round mounds (Figure 5.35) and cairns (Figure 5.36). Archaeological excavation can identify human skeletal material in specific contexts, but cannot identify what contexts they have entered and how long the process took from death to their final deposited context. The form of deposition and the location of deposits indicate that in the majority of cases human skeletal material would be accessible and could be recovered after deposition. These then could be subject to new sets of natural and cultural processes.

that the population circulated human skeletal material (Figures 5.39 and 5.40) (Ashbee 1970:83; Bradley 1998a:Chapter 4, 2000a:122; Bradley and Edmonds 1993:26; Lucas 1997; Smith 1965a; Thomas 1991a, 1999a, 2000a; Thomas and Whittle 1986:148; Thorpe 1984, 1996:Chapter 7; Whittle et al. 1999a:Chapter 17, 18). Smith (1965a:137) argued that human skeletal material from the chambered structure at West Kennet (Wiltshire) was re-deposited into the ditches of the causewayed enclosure at Windmill Hill (Wiltshire), three kilometres away. Examination of the assemblage of human skeletal material from West Kennet (Wiltshire) revealed too many mandibles compared with skulls and far too few long bones (Piggott 1962:23-4 and 79-89). The movement of human skeletal material between West Kennet and Windmill Hill is supported further by the near-identical radiocarbon dates obtained from both these contexts

The predominance of disarticulated, fragmentary and incomplete human skeletal material, with some individuals represented by only a few fragments, has led to arguments that there was a general practice of movement, circulation and arrangement of human skeletal material in the British and Irish ‘Neolithic’ (Figures 5.37 and 5.38). Unevenly represented elements of human skeletons were identified as early as the 1860s (Greenwell 1865:107; Thurnam 1869:185-6). Disarticulated assemblages of human skeletal material from stone chambered structures and long mounds from Yorkshire and Wessex have been presented as evidence 59. Here it is worth commenting on the different interpretations offered for human crania from riverine deposits in the British Isles. These deposits have been viewed as ritual acts (Bradley 1995a; Bradley and Gordon 1988; West 1996). However, other scholars have offered alternative interpretations based on demographic and taphonomic arguments (Knüsel and Carr 1995, 1996; Turner et al. 2002).

102

Figure 5.22– Plan and section of the central pit at Willington Plantation Quarry (Bedfordshire) (after Dawson 1996: figure 3).

103

Figure 5.23– Alington Avenue (Dorset): Phase 1, earlier prehistoric (after Davies et al. 1985: figure 2).

104

Figure 5.24– Plan and section of Chilbolton (Leckford Estate) (Hampshire), including details of central features (after Russel 1990: figure 2).

Figure 5.25– Child skeleton at base of ditch at Flagstones House (Dorset) (after Smith et al. 1997: figure 28).

105

Figure 5.26– A human skull fragment found on the clays at Carrigdirty Rock (Co. Limerick) (after O'Sullivan 1997a: page 15).

Figure 5.27– Distribution of human skeletal material at Stonehenge (after Cleal et al. 1995: figure 250).

106

Figure 5.28– Longitudinal section of Kilgreany (Co. Waterford) (after Movius 1935: figure 3).

Figure 5.29– Excavation of a skeleton in cist 2 at An Corran (Highland) (after Downes and Badcock 1998: figure 28).

107

Figure 5.30 – Bog body found at Castle Balkeney (Gallagh) (Co. Galway) (after Ó Floinn 1995a: figure 66).

Figure 5.31 – Schematic drawing of the excavated deposits at Carding Mill Bay (Argyll and Bute), showing sections (A-B, C-D, E-F) and a summary of the phasing of contexts (after Connock et al. 1991-2: figure 2).

108

Figure 5.32 – Sketch plan and section of Quanterness (Orkney Islands), with the later roundhouse to the east (after Renfrew et al. 1976: figure 1).

109

Figure 5.33 – Plan and sections of Grange (Co. Roscommon) (after Ó Ríordáin 1997: figure 2).

110

Figure 5.34 – Photograph of the human skeletal material at Giants’ Hills 2 (Lincolnshire), from the west. 30 centimetre scale (after Evans and Simpson 1991: plate viia).

Figure 5.35 – Photograph of the primary articulated teenage male skeleton with Food Vessel at Tallington 16 (Lincolnshire) (after W.G. Simpson 1976: plate 24 upper).

111

Figure 5.36 – Ground plan and section of Collessie (Fife) (after Anderson 1888: figure 3 and 4). (Whittle 1993:Table 1 and also the Online Database).60 Missing human skeletal categories have also been identified at the causewayed enclosure at Hambledon Hill Complex (Dorset), with skulls nearly always missing the lower jaw (Mercer 1980), and from the chambered structure at Hazleton North (Gloucestershire), which has too few long bones and probably too few skulls (Saville 1990). Missing human skeletal elements were also identified at Lanhill (Wiltshire), where one set of arms, one spinal column and a pelvis and one set of legs and a skull, were absent from three respective

skeletons (Fowler 2001:143; Keiller and Piggott 1938). Work on the many chambered structures on the Orkney Islands by A. Jones (1997, 1998), Reilly (2003), Richards (1988) and Barber (1988) has identified the lack of specific human skeletal elements in some contexts, such as at Quanterness and Isbister. This has been taken as evidence that the population circulated and rearranged human skeletal material assemblages. Jones identified that the level of disarticulation within chambered structures was related to the topography of the context. He argued that full articulation of human skeletons occurred in lowland locations and human skulls were identified in higher locations (1998:318) (Figure 5.41). Extensive taphonomic analysis by Baxter on the chambered structure at Ascott-under-Wychwood (Oxfordshire) and earthen long mound at Haddenham (Cambridgeshire) has identified that human skeletal material was circulated and moved while still “partdecayed; rotting, but still hanging together with ligaments, tendons and possibly muscles intact” (1999:6). Ascott-under-Wychwood (Oxfordshire) contains 46-49 individuals and has two human bone samples radiocarbon

60. Human skeletal material from West Kennet has produced the calibrated radiocarbon dates of 5740-5440 BP (.842) and 5420-5320 BP (.158) (4825±90 bp, OxA449); 5710-5700 BP (.004) and 5660-5310 BP (.996) (4780±90 bp, OxA-563); 5710-5700 BP (.004) and 56605310 BP (.996) (4780±90 bp, OxA-451); 5600-5290 BP (.989), 5160-5140 BP (.006) and 5100-5090 BP (.005) (4700±80 bp, OxA-450). Human skeletal material from Windmill Hill has produced calibrated radiocarbon dates of 5598-5434 BP (.680) and 5424-5319 BP (.320); (4750±70 bp, OxA-2399) and 5600-5440 BP (.665) and 5420-5320 BP (.335) (4745±70 bp, OxA-2403).

112

Figure 5.37 – General plan of Fussell’s Lodge long mound (Wiltshire) as excavated (after Ashbee 1966a: figure 2).

113

Figure 5.38 – Fussell’s Lodge (Wiltshire): Distribution of human anatomical parts (after Thomas 1999a: figure 6.5).

114

Figure 5.39 – Contexts of burials in early ‘Neolithic’ Wessex for children and adults (after Thorpe 1996: figure 7.6).

Figure 5.40 – Contexts of deposition of body parts in early ‘Neolithic’ Wessex (after Thorpe 1996: figure 7.7). dated from the sixth to the early fourth millennium BP.61 With much of the remains disarticulated and with a general lack of small bones of the hands and feet, Baxter argues that initial treatment of the corpse took place somewhere other than inside the structure. However, the possibility exists that bones may have disappeared due to taphonomic processes within the chambered structures. The literature argues that human skeletal material was circulated and deposited in a variety of contexts, as

Thomas (2000a:660, also see Darvill and Thomas 2001:15) states: “We might talk, then, of a general economy of human remains”. The empirical evidence illustrates that skeletal material and artefacts were circulated across the landscape in similar ways. Thus we may need to question other dichotomies which we assume to be universal, such as culture:nature (see Barnatt and Edmonds 2002; Bradley 2000a:122; Brück and Goodman 1999:8-10; Cummings 2002; Evans et al. 1999; Gojda 2001; Higginbottom et al. 2001; Ingold 1995; Johnston 1999a, 1999b; Scarre 2002b; Thomas 1999a:68, 226, 228) as we may be missing a significant behaviour characteristic of prehistoric populations in Britain and Ireland.

61. The two calibrated radiocarbon dates are: 5910-5570 BP (.905) and 5560-5470 BP (.095) (4930±100 bp, BM1976R) and 4530-3980 BP (1.000) (3870±100 bp, BM1975R).

The empirical evidence above should make it inappropriate and misleading to label certain contexts as ‘tombs’ or ‘barrows’. This argument will be developed throughout, but suffice it to say that the patterning of the dead across the landscape is quite different in the 115

Figure 5.41 – A schematic diagram illustrating the relationship between the deposition of human skeletal material and the landscape (after Jones 1998: figure 4). such as ‘tomb’ and ‘house’ are not appropriate. However, I am not arguing that contexts did not exist with only human skeletal material.

‘Neolithic’ than in our own current Western/modern society, so much so that notions of ‘tombs’ and ‘barrow’ may be wholly inappropriate in discussions of the treatment and deposition of this human skeletal material. The labels, drawn from Western/modern society and observed and documented human social behaviour, are incapable of conveying the complex and intermeshed pattern of human social behaviour which occurred in prehistoric Britain and Ireland, not just in relation to human skeletal material, but also the relationship of that material to the occupation locations and other built locations (see Chapter 6).

Lane’s case study of Mount Pleasant Farm (Rhondda, Cynon, Taff) identifies the apparent change of function of a context from ‘domestic’ to ‘funerary’, a pattern which has been identified across western and northern Europe (for example, Alexander et al. 1960; Andersen and Johansen 1990; Asingh 1987; Eogan 1963, 1974, 1986, 1991; Eogan and Roche 1997; Gibson 1996; Gramsch 1995; Grogan 1996; Grøn and Skaarup 1991; Hibbs 1984; Hodder 1990a; Johansen 1985; C.S. Jones 1997; O’Kelly 1989; Madsen 1979; Manby 1976; Nielsen and Nielsen 1991; Price 1987; Robertson-Mackey 1980; Saville 1989; Sharples 1986; Skaarup 1982; Smith and Simpson 1966; Whittle et at. 1993). However, the data I have gathered in the British and Irish ‘Neolithic’ poses a far more complex interpretative problem for the current labels being employed. The Online Database (see Appendix 6) documents 25 locations which contain radiocarbon dated human skeletal material and other archaeological assemblages classified as occupation debris, suggesting that ‘domestic’ and ‘funerary’ activities are not necessarily performed in distinct localities across the landscape.

Living with the dead Paul Lane’s 1986 paper questioned the archaeological employment of ‘ritual’ and ‘domestic’ in describing archaeological assemblages and contexts. He argued: their use introduces a set of largely ethnocentric, and frequently androcentric, assumptions, which serve to reinforce and reproduce an appearance of mutual exclusiveness and opposition between these two aspects of human action. (1986:181) Similarly, Trevor Kirk questions the assumed dichotomy between ‘ritual’ and ‘domestic’ activity (also see Brück 1999a, 1999b; Holtorf 2001). Kirk (1998:104-5), in his discussion of the Paris Basin, describes the interaction between the living and the dead as a “routinised social practice just as is the making of a cup of tea or the production and use of pottery or flint tools”. He describes a mortuary process in which ‘tombs’ were only one of many cultural realms that the dead entered. Drawing from these arguments, I would like to express similar problems with the terms ‘funerary’ and ‘domestic’, that is, ‘places for the dead’ and ‘places for the living’. The empirical evidence from the British and Irish ‘Neolithic’ clearly shows that human skeletal material was not confined to certain privileged contexts. Rather, they occupied a variety of contexts, some of which have produced evidence of prolonged human occupation. This means that modern categories that dichotomise such activities,

The language of social labels employed throughout the history of archaeological discourse in Britain and Ireland has created a false landscape where the living and dead are separated. Labelling contexts as ‘tombs’ or ‘barrows’ identifies them as distinguishable locations where human skeletal material is found. Currently, the interpretative process is such, that a context like Beckhampton Road (Wiltshire) (Figure 5.42), which contains no human skeletal material but only animal bones and artefacts, is identified as a ‘barrow’, and a context like Hambledon Hill Complex (Dorset), which contains a mass of human skeletal material, is identified as a causewayed enclosure. Labelling a certain class of structure a ‘tomb’ or ‘barrow’ leads to a situation whereby, first, human skeletal material is expected and second, its interpretation starts 116

Reed 1974; Sherratt 1990, 1993, 1995).62 Hodder (1984:59) suggests eight points of similarity between the long ‘houses’ of central Europe and the long mounds of western and northern Europe. These include their form, internal organisation, external ditches and alignment. Not all of these features are characteristic of all structures, but it is claimed that recognition of a general relationship is warranted, suggesting some kind of transformation of ‘domestic’ architecture into ‘funerary’ architecture (Marshall 1981:110; Sherratt 1990). For Sherratt (1990:150), this transformation indicates “the need for more permanent monuments (just as the church in a medieval village was often the only building of stone)”. In this view, the structural dichotomy leads to the correlation that stone was related with the dead, and timber with the living (see Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Whittle 1997c). But such a simple dichotomy fails to encapsulate all the empirical evidence, even if we wish to maintain the assumed dichotomy between the living and the dead.63 An interesting case is provided by the structural and depositional events on the Orkney Islands.

with the assumption that human skeletal material should be present and thus any absence then has to be explained. I have documented 73 contexts identified as ‘Neolithic’ ‘tombs’ and ‘barrows’ that contain no human skeletal material (see the Online Database). The interpretation of these contexts needs to be revisited. Unfavourable soil conditions for the preservation of bone have been a popular environmental reason for its absence, however, this reason is doubtful when animal bones are also recovered. Cultural interpretations used to describe the apparent lack of human skeletal material, have been robbing of the structure or that the structure was a cenotaph. Identifying a structure as a cenotaph has been a widespread interpretive ploy when discussing British chambered ‘tombs’ and ‘barrows’ since the beginning of their scientific exploration in the nineteenth century (Coffey 1896; Greenwell 1877). Currently, the discussion of cenotaphs creates an indefensible position; an archaeological interpretation without any artefactual proof. Thus, the interpretations are derived only from the social conventions being used and not from the data. The lack of human skeletal material in identified ‘tombs’ and ‘barrows’ still poses an interpretive problem for current scholars (Bradley 1998a; Edmonds 1999a; Lucas 1997; Thomas 1991a, 1999a; Thorpe 1996). Much discussion has focused on two long ‘barrows’ in Wiltshire; South Street and Beckhampton Road (Wiltshire) (Ashbee et al. 1979). At the Beckhampton Road long mound, three cattle skulls and various long bones were located along the axis of the mound, the area where human skeletal material is usually located (Figures 4.10 and 5.42) (Ashbee et al. 1979:247; Thorpe 1984:51; Whittle et al. 1999c:359). Elaborate social stories have been formulated to explain this lack of human skeletal material based on symbolism – replacing human subjects with animal ones (Edmonds 1999a:61; Thorpe 1984, 1996:163; Whittle et al. 1999c:359, see below).

In the Orkney Islands, both types of structures, that is those interpreted as ‘houses’ and those regarded as ‘tombs’, are constructed from stone due to the lack of timber in the region. Such instances suggest that a simplistic dichotomy of raw materials and function is dubious. Moreover, the existence of timber ‘mortuary’ structures under earthen mounds in Britain further shows that the claimed dichotomy between structures for the dead (stone) and structures for the living (timber) is only a ‘reality’ in archaeological discourse, not in the actual archaeological contexts. This indicates the problematic practice of associating structural remains with discrete familiar ‘social’ functions. For example, from the evidence from the Orkney Islands, in particular Skara Brae (Figure 5.45), Richards (1992a, also see A. Jones 1998) has identified structural similarities between ‘houses’ and ‘tombs’. The ‘houses’/‘village’ at Skara Brae, along with the ‘house’ at Barnhouse (Orkney Islands), has rock art on the walls of the structure similar to those on ‘tombs’ (Bradley et al. 1999, 2001; Shee Twohig 1981) (Figures 5.46, 5.47, 5.48, 5.49, 5.50, 5.51, 5.52, 5.53 and 5.54) and also human skeletal material, including two articulated female skeletons in a cist below the wall of ‘hut’ 7, and disarticulated human skeletal material. Based on the current discourse one could argue that the Skara Brae ‘village’ was in fact a chambered ‘tomb’, with analogy for the structural and artefactual

The situation is complicated further by the structural similarities that have long been identified between ‘houses’ and ‘barrows in continental western and northern Europe (see Chapter 4). In the 1860s, Sven Nilsson, who initiated one of the first scientific studies of ‘megalithic tombs’ in Sweden (1868), compared the ground plans of Eskimo huts in Greenland and North America with those of Swedish ‘passage-tombs’, identifying seven morphological similarities between the two (Figure 5.43) (Nilsson 1868:139). More recently, in this tradition similar arguments have been made in relation to long ‘houses’ and long ‘barrows’ in western and northern Europe (Figure 5.44) (Bradley 1998a; Hodder 1984, 1990a, 1994; Madsen 1979; Marshall 1981;

62. Gordon Childe (1949b) also identified a link between the plans of long ‘houses’ and long ‘barrows’ in western and northern Europe. 63. Chris Scarre (2002:47) has recently identified that long ‘houses’ and long ‘barrows’ do overlap at Balloy (Burgundy, France) (Mordant 1997).

117

Figure 5.42 – Phases of Beckhampton Road (Bishops Cannings 76) (Wiltshire) (after Pollard 1993: figure 2.36).

118

Figure 5.43 – Reproduction of plate XIV from Nilsson’s The Primitive Inhabitants of Scandinavia showing ground plans of Swedish passage graves and a reconstruction diagram of an Eskimo house (after Tilley 1998: figure 4).

119

Figure 5.44 – Comparison of ‘tombs’ (stalled cairns - A, B, C) and ‘houses’ (D) in the Orkney ‘Neolithic’. A. Kierfea Hill. B. Knowe of Craie. C. Knowe of Yarso. D. Knap of Howar (after Hodder 1984: figure 7).

120

Figure 5.45 – Plan of Skara Brae (Orkney Islands) (after Shee Twohig 1981: figure 286). 121

Figure 5.46 – Abstract design at the Maeshowe (Orkney Islands) chambered ‘tomb’ (left) overlain by runic inscription (right) (after Bradley et al. 2001: figure 6).

122

Figure 5.47 – Incised motifs on the Orkney Islands at the Cuween Hill chambered ‘tomb’ (1-3), the Holm of Papa Westray South chambered ‘tomb’ (4) and the Quoyness chambered ‘tomb’ (5) (after Bradley et al. 2001: figure 8).

123

Figure 5.48 – Incised motifs on the Orkney Islands at the Quoyness chambered ‘tomb’ (6 and 7) and the Wideford Hill chambered ‘tomb’ (8) (after Bradley et al. 2001: figure 9).

124

Figure 5.49 – Incised motifs on the Orkney Islands at the Wideford Hill chambered ‘tomb’ (after Bradley et al. 2001: figure 10).

Figure 5.50 – Incised motifs from the chambered ‘tomb’ at Maeshowe, the ‘house’ Barnhouse and a cist slab from Brodgar from the Orkney Islands (after Bradley et al. 2001: figure 13). 125

Figure 5.51 – Skara Brae (Orkney Islands). Decorated stones from ‘Hut’ 7 and 8 (after Shee Twohig 1981: figure 287).

126

Figure 5.52 – Skara Brae (Orkney Islands). Decorated stones from the Passages (after Shee Twohig 1981: figure 288).

127

Figure 5.53 – Skara Brae (Orkney Islands). Stones 42-5 (after Shee Twohig 1981: figure 289).

128

Figure 5.54 – Skara Brae (Orkney Islands). Decorated stones from the passage and miscellaneous locations (after Shee Twohig 1981: figure 290). The recent excavation report of the Windmill Hill (Wiltshire) causewayed enclosure has identified a number of contexts with parallel treatment and deposition of humans and animals, in particular cattle. Both articulated and disarticulated human and animal skeletal material is present (Whittle et al. 1999c). From Trench F 630 (inner ditch), a human child femur was found lodged within an ox humerus (Figure 5.55) (ibid.:357). From context 117 in Trench A, a human infant cranium was excavated placed adjacent to an ox frontal (ibid.:362). While there are no spatial or taphonomic reasons to view the human and animal skeletal material as ‘different’, the Western/modern conceptual dichotomy of animal:human is maintained, illustrated by this statement: One of the most striking patterns to emerge is the parallel treatment sometimes afforded to human and animal bones…Domestic animals were

evidence from other structures identified as ‘tombs’. Animal and human skeletal material: conflicting views At the beginning of the chapter I argued that the kinds of processing and deposition applied to human corpses and human skeletal material could also be carried out on animals and animal skeletal material. Similar treatment is widely afforded to both humans and animals across Britain and Ireland in the ‘Neolithic’. Both can be deposited as entire articulated skeletons or can be reduced into the smallest skeletal elements located in similar contexts. The similar treatment, deposition and location of both animal and human corpses have long been identified in the British and Irish ‘Neolithic’ (Leaf 1935; Mortimer 1905; Richmond 1999:17-8; Thomas 1988b; Thurnam 1869; Whittle et al. 1999c). 129

Figure 5.55 – Windmill Hill (Wiltshire) Context 630, Trench F. Early ‘Neolithic’. The distal half of the humerus of an ox, with the shaft of the femur of a child inserted into the marrow cavity (after Grigson 1999: figure 161). At Beckhampton Road (Wiltshire) long mound the three cattle skulls and various long bones were located along the axis of the mound, the location normally reserved for human skeletal material (Figure 5.42) (Ashbee et al. 1979:247; Thorpe 1984:51; Whittle et al. 1999c:359). Thorpe (1984:51) explains that this illustrates an animal hierarchy with the possible identification of cattle “as a homologue to the aristocratic clan”. Whittle et al. (1999c:359 and 362) argue that the cattle are taking the place of the human with: metaphorical connections perhaps being drawn between the transformation of cattle through butchery and consumption and the transformation of the human corpse in mortuary ritual. The presence of animal bones, particularly cattle, is documented at many ‘barrows’ and chambered ‘tombs’ in Britain and Ireland. From Amesbury 14 and Amesbury 42 (Wiltshire) long ‘barrows’, cattle skulls and feet were excavated. While at Heytesbury (Wiltshire) long ‘barrow’, the head and horn of at least seven cattle have been excavated and from Tilshead 5 (Wiltshire) long ‘barrow’, cattle skulls have also been found (Ashbee 1970). At Fussell’s Lodge (Wiltshire) long ‘barrow’ an ox hide with skull and feet was located within the structure, with other cattle bones placed along the axis of the flint cairn (Ashbee 1966a). Irthlingborough 1 (Northamptonshire) round ‘barrow’ has produced 184 cattle skulls and a few aurochs skulls covering a central pit with disarticulated human skeletal material (Davis and Payne 1993; Halpin 1987). Deposits of cattle bones can be found at the ‘barrows’ and ‘tombs’ at Willerby Wold (North Yorkshire), Wor Barrow (Dorset), Mourne Park (Co. Down), Ballyalton (Co. Antrim), Annaghmare (Co. Armagh), Lough Gur (Co. Limerick), Giants’ Hills 2 (Lincolnshire), Lambourn (Berkshire), Abingdon (Oxfordshire), Kilham (Humberside), Crosby Garrett (Cumbria), Quanterness, Isbister (Orkney Islands) and Nutbane (Hampshire) (Ashbee 1970; Barber 1988; Bradley et al. 1984; Grigson 1981; A. Jones 1997, 1998; McCormack 1985-6; Schulting 2000; Wymer 1966).

unlikely to have been invested with precisely the same rights and qualities as people. They were after all eaten. (ibid.) This comment illustrates the preconception which is prevalent throughout the discourse, an aversion to look at the empirical evidence in a different way in order to identify unparalleled forms of human behaviour not present in our Western/modern social life. The similar treatment is explained as the possible metaphor that animal bones may have had for human existence, linked with birth and death (ibid.:385). Thus a similar explanation is pursued as that outlined in relation to the deposition of cattle skulls at the Beckhampton Road (Wiltshire) long mound (see below). At Ascott-under-Wychwood (Oxfordshire), Chesterman stated that animal bones, including cattle, sheep/goat, pig and dog, pottery, charcoal and flint, were mixed with human skeletal material in the passage and the chamber (Chesterman 1977:24; Fowler 2001:143). At Holm of Papa Westray, North (Orkney Islands), deer and human skulls were deposited along with deer tines (Fowler 2001:144; Richards 1988:53-4). From the central chamber at Cuween Hill (Orkney Islands), 25 dog skulls were intermixed with five human skulls and seven fully articulated dog skeletons were “each deposited with between two and three human skulls in separate stalls and side cells” (Jones 1998:311, 2002, also see Fowler 2001:144). From Ballaharra (Isle of Man), two mixed deposits have been identified. The first deposit contained skeletal material from a sheep/goat, a dog, a ‘pheasantsized bird’ and the cremated remains of 33-40 humans, including skull fragments and vertebrae from children and long bones from adults, as well as four arrowheads, other flint implements, slate and pottery. The second deposit contained the cremated remains of three to five humans, including mainly skulls and long bones, dog shins and ankle bones and vertebrae from an unidentified small mammal (Cregeen 1978:146-9, also see Fowler 2001:151-2). 130

to determine whether the absence of cut marks on animal bones is a feature common throughout the ‘Neolithic’ of Britain, due to the lack of comment on cut marks in excavation reports, such as the chambered structure at Hazleton North (Gloucestershire) and the occupation context at Knap of Howar (Orkney Islands). He concludes, based on the current evidence, “that butchery marks [cut marks] do not occur on any Neolithic bone assemblages” (ibid.).

Thomas’ study of deposition at Cotswold-Severn chambered structures states: In chambered contexts cattle bones are most frequently recovered and appear to have been given similar treatment to the human bones in the same contexts – burnt where human bones are burnt, articulated where humans are articulated, disarticulated where humans are disarticulated (1988b:549) Recent re-examinations of the excavated human and animal assemblages from West Kennet (Wiltshire) have identified similar treatment of the human and animal (Liddle 1998; Thomas and Whittle 1986). Within the five chambers of the stone structure there is an ordering of animal species, an ordering we also see for the human assemblage based on age and sex at West Kennet (Wiltshire), Burn Ground (Gloucestershire) and Notgrove (Gloucestershire) (Figure 5.56) (Thomas 1988b). For example, dog bones are only present in the northeast and northwest chambers and the northeast chamber contains a mass of cattle bones (Figure 5.57) (Thomas and Whittle 1986:147). Interpreting this, Thomas and Whittle cite Thorpe’s (1984:51) ‘hierarchy of animal’ argument for Beckhampton Road (Wiltshire). Jane Liddle’s (1998) recent examination of 765 animal bones from West Kennet (67% of the total assemblage) has identified a number of patterns. Firstly, it is interesting to note that during the course of her analyses a number of human bones were identified, including scapula, mandibles and foetal bones, illustrating the problems of identifying species from bone fragments (Liddle 1998:5). Gnawing by canids was identified on 91 bones (12% of the assemblage), mainly visible on cattle and pig bones. A further 12 bones (1.6% of the assemblage) had evidence of rodent gnawing (ibid.:11). Seven bones (0.9% of the assemblage) had evidence of burning, with 103 bones (13% of the assemblage) having butchery marks (ibid.:12). Also, a near-complete articulated goat skeleton was found in the second phase of the northwest chamber. West Kennet and other ‘tombs’ and ‘barrows’ have other animal species, suggesting that the practice was not only confined to cattle bones (see Herity 1987; A. Jones 1997, 1998; McCormack 1985-6; Richmond 1999:18).

Just like human skeletal material, animal bones may have been circulated and exposed to a variety of treatment (see Darvill and Thomas 2001:15). Animal bones have been found in many of the contexts discussed in the Online Database (see Appendix 6). Radiocarbon dating evidence from Stonehenge I (Wiltshire) indicates that cattle bones were 300 years old when they were deposited in the ditch. From Parc le Breos Cwm (Swansea), a badger bone radiocarbon dated to the ninth millennium BP64 and a large ungulate bone radiocarbon dated to the thirteenth millennium BP,65 have been found in a chambered structure constructed in the ‘Neolithic’. If these bones were deposited with the bulk of the human skeletal material, they would have been over 3000 and 7000 years old respectively (Whittle and Wysocki 1998). However, archives recently discovered at the Natural History Museum in London by Wysocki, seem to suggest that the animal bones from the passage which have given the very old radiocarbon dates were in fact excavated from the nearby Cat Hole Cave (Swansea) in the late nineteenth century and subsequently reburied in the chambered structure by Lubbock and Vivian in 1869 when the structure was first discovered and excavated (Michael Wysocki personal communication).

Human skeletal material in the ‘Mesolithic’ I will now turn to human skeletal material in the ‘Mesolithic’, from the early tenth to the middle of the seventh millennium BP. I will return to the ‘Neolithic’ in the later section of this chapter. A general assumption is that there is a lot more human skeletal material in the ‘Neolithic’ than the ‘Mesolithic’. This is correct if all human skeletal material which is assumed to be ‘Mesolithic’ and ‘Neolithic’ is included. However, if only

Generally for stone chambered ‘tombs’ and earthen and timber ‘barrows’, animal bone assemblages are interpreted as sacrificial offerings (Clutton-Brock 1979), funerary feasting (Edmonds 1999a; Grinsell 1958; Hedges 1983, 1984; Mount 1994; Renfrew 1979; Thomas 1991a, 1999a) and/or as evidence of totemic practices (Fraser 1983; Hedges 1983, 1984). The animals are always subservient or else embody or symbolise human skeletal material, maintaining our Western/modern human:animal dichotomy. However, there is no taphonomic or depositional reason to do so. Barber states that none of the animal bones from Quanterness and Isbister (Orkney Islands), which were used by Renfrew and Hedges to argue for funerary feasts, exhibited butchery marks, and he goes on to say that “this seems typical of animal bone assemblages from chambered tombs” (Barber 1997:67). He states that this is not simply a consequence of past excavation standards in bone retrieval and analysis. Barber points out that it is difficult

64. The calibrated radiocarbon date is: 8590-8570 BP (.046) and 8560-8370 BP (.954) (7665±65 bp, OxA6499). 65. The calibrated radiocarbon date is: 12960-12580 BP (.679), 12520-12320 BP (.309), 12220-12180 BP (.013) (10625±80 bp, OxA-6500).

131

Figure 5.56– Predominant character of human skeletal material in the chambers of three chambered structures: 1) Burn Ground (Gloucestershire); 2) Notgrove (Gloucestershire); 3) West Kennet (Wiltshire) (after Thomas 1988b: figure 6).

132

Figure 5.57– Relative representation of human bones, animal bones and pottery vessels in different phases of the secondary deposit at West Kennet (after Thomas and Whittle 1986: figure 6).

133

Point (Daylight Rock Cave) (Pembrokeshire), Paviland Cave (Goat’s Hole and Fox Hole), Worm’s Head Cave (Swansea), Badger Hole (Somerset), Kent’s Cavern, Oreston Third Bone Cave (Breakwater Quarry) (Devon), Killuragh (Co. Limerick), Rockmarshall Midden 111 (Co. Louth) and Ferriter’s Cove (Co. Kerry) (see the Online Database and Appendix 6). It is of note that three contexts have disarticulated human skeletal material radiocarbon dated to both the ‘Mesolithic’ and the ‘Neolithic’: Pontnewydd Cave (Denbighshire), Paviland Cave (Goat’s Hole and Fox Hole) (Swansea) and Kent’s Cavern (Devon).66

human skeletal material directly dated by radiocarbon dating is included, the picture is quite different. Recent work by Woodman, Chamberlain, Richards and Schulting have identified many more contexts with human skeletal material dated to the ‘Mesolithic’. Again, only using contexts with radiocarbon dated human skeletal material, I will document the treatment, deposition and location of human skeletal material during the ‘Mesolithic’. I will argue that there is continuity with assemblages of human skeletal material dated to the ‘Neolithic’ (see Bradley 1998a:61). What is distinctive in the ‘Neolithic’ is the new structural components which proliferated across the landscape, including henges, enclosures, mounds and chambered structures and the practice of cremation, as identified in the archaeological excavation literature.

The majority of interest in the taphonomic analysis of human skeletal material assemblages has come from research on Oronsay, a small island off the west coast of Scotland, which has produced two shell middens with human skeletal material radiocarbon dated to the ‘Mesolithic’: Cnoc Coig and Caisteal nan Gillean II (Argyll and Bute) (Meiklejohn and Denston 1987; Mellars 1987; Nolan 1986, also see Pollard 2000b). Cnoc Coig contains 49 human skeletal fragments, including skulls, long bone shafts, hands, feet, vertebrae, ribs and pelvis, representing a minimum of four individuals in four deposits in close proximity to hearths (Figures 5.59 and 5.60). Importantly, the vast majority of the human skeletal material from Cnoc Coig (95.7%) was recorded in situ, with 60% coming from extremities (wrists, hands,

Treatment and deposition of human skeletal material in the ‘Mesolithic’ The various options available to the population when an individual died have been outlined above. From the early tenth to the middle of the seventh millennium BP, one can assume that similar options would have been available to the occupants of Britain and Ireland. However, cremation is not known in British and Irish contexts. This is a curious issue because the practice of cremation in the ‘Mesolithic’ is documented throughout continental Europe (see Schulting 1998b:215). Both articulation and disarticulation are documented, with disarticulation the dominant condition of the skeleton (see the Online Database and Appendix 6). Detailed discussions of radiocarbon dated human skeletal material assemblages have been limited to four contexts, Gough’s New Cave (Somerset) and Aveline’s Hole (Somerset) (Barton 1999; Chamberlain and Williams 2001; Smith 1992) and two shell midden contexts on Oronsay (Argyll and Bute); Cnoc Coig and Caisteal nan Gillean II (Meiklejohn and Denston 1987; Mellars 1987, Nolan 1986). However, I will attempt to expand the discussion with reference to the Online Database (see Appendix 6). At Gough’s New Cave and Aveline’s Hole (Somerset) both disarticulation and articulation occur. There has been less investigation of the assemblage at Aveline’s Hole due to the destruction of the human skeletal material and the excavation records, during the Second World War. From Gough’s New Cave, microscopic analysis of cut marks and scrapes found on human skulls, human mandibles and other human skeletal material indicates that the marks are post-mortem, produced with the sharp edge of a flint knife. Detailed study of the positioning of the cut marks shows that the human corpse was skinned and the joints dismembered soon after death. The tongue was removed before detaching the jaw from the rest of the head (Figure 5.58) (Barton 1999:22-3; Cook 1991). Gough’s New Cave also produced burnt and highly fragmented human skeletal material. Here I want to concentrate on disarticulation of human skeletal material in the ‘Mesolithic’ due to the implications this treatment has for the ‘Neolithic’ contexts (see below). Disarticulated human skeletal material has been excavated from Pontnewydd Cave (Denbighshire), Ogofyr-Ychen (Cave of the Oxen), Potter’s Cave, Small Ord

Figure 5.58– Cut marks on a human lower jaw bone from Gough’s New Cave (Somerset) (after Barton 1999: figure 2.5).

66. Interpretation of these contexts could be significant in light of the finds from Stonehenge I (Wiltshire) indicating the circulation and deposition of old animal bones (see page 131).

134

the twelfth and the eleventh millennium BP,67 was contained within the structure while the rest was distributed throughout the cave (Figure 5.62). The absence of certain skeletal elements, cut marks on one skull, and the presences of ochre within the structure, have been identified as evidence of secondary burial (Cauwe 2001:151-2).

ankles and feet) (Nolan 1986:254-6). No cut marks have been found on the bones, suggesting to Meiklejohn and Denston (1987) that they were interred elsewhere with certain skeletal elements being intentionally removed and brought to Cnoc Coig. However, this human skeletal material assemblage could be interpreted as having been exposed on the shell midden at Cnoc Coig, with the larger skeletal elements being moved away from the shell midden, and smaller elements being left behind (Bradley 1997a:14-5; Pollard 1996:204; Telford 2002:295-7). From Caisteal nan Gillean II a similar, though smaller, assemblage of five human skeletal fragments has been identified, including vertebrae, hand and foot bones (Meiklejohn and Denston 1987). It is interesting to note that in the ‘Neolithic’, the lack of such skeletal elements in chambered structures has been seen as evidence that initial treatment of the corpse took place somewhere outside the structure. Further, recent work by Schulting and Richards (2001:324), argues that the placing of successive human interments in ‘tombs’ and the manipulation of human skeletal material seen throughout ‘Mesolithic’ Europe, particularly in northern France and Belgium, may provide the antecedence for the practice in ‘Neolithic’ contexts (Table 5.6) (also see Case 1976; Childe 1957:266; Cullen 1995; Scarre 1992, 1998, 2002a; Telford 2002:297). As Scarre (2002a:45) points out, “It would be hardly surprising if Neolithic burial practices had drawn upon this long-established and widespread Mesolithic tradition”. Further, the dominant practice at Bandkeramik contexts is for single burials, with double burial rare and only 3 examples of triple burials out of some 2500 examples. There are also only a few examples having been documented of post-mortem disturbance or manipulation of human skeletal material (Jeunesse 1997), ‘Mesolithic’ input seems likely (Scarre 2002a:45-7). It is important to note that this view has a long tradition, with Childe (1957:266) claiming that it “can hardly represent the unifying idea, since collective burial in natural caves was practised even in Mesolithic Palestine”.

Abri des Autours (Namur) held two human skeletal assemblages. The older is a female skeleton within a small pit covered by ochre. The second is the cremated and disarticulated human skeletal assemblage and four non-retouched flint bladelets (Figure 5.63). The human skeletal material contained at least five adults and six children and has been radiocarbon dated to the second half of the twelfth and the first half of the eleventh millennium BP (Cauwe 2001:154-7).68 While this evidence may not provide the origin for the practice identified from the seventh millennium BP throughout western and northern Europe, it does illustrate that the disarticulation, cremation, manipulation and circulation of human skeletal material is not a solely ‘Neolithic’ practice. Further, these ‘collective burials’ identified in Belgium, which contain the human skeletal material of a number of individuals, is compatible with a number of cave contexts in Britain and Ireland from the early tenth to the middle of the seventh millennium BP (see the Online Database and Appendix 6). As with ‘Neolithic’ contexts, parallel treatment of human and animal skeletal material has been documented in ‘Mesolithic’ contexts. While there are the famous examples of dog burials from continental Europe at Skateholm, Sweden (Figure 5.64) (Larsson 1990a) and the Iron Gates Gorge, Yugoslavia (Radovanoviü 1996, 1999), most contexts in Britain and Ireland have produced disarticulated animal bones. Bradley (1997a:14) states that in Europe, ‘Mesolithic’ beads were occasionally fashioned out of human teeth, which also

Nicolas Cauwe (1996, 2001) has recently argued that the tradition of collective burials, involving movement and manipulation of human skeletal material, may extend back into ‘Mesolithic’ and ‘Upper Palaeolithic’ contexts. His argument is based on two early ‘Mesolithic’ collective burials excavated in southern Belgium caves, Grotte Margaux and Abri des Autours (Cauwe 2001). Grotte Margaux (Namur) contains a small pit partly surrounded by a dry-stone wall and a pavement covered by a stone roof (Figure 5.61). Some of the female disarticulated human skeletal material, which has been radiocarbon dated from six samples to the second half of

67. The six calibrated radiocarbon dates are: 1119110667 BP (.968), 10661-10638 BP (.019), 10611-10596 BP (.010) and 10588-10583 BP (.003) (9590±110 bp, Gif-92354); 11173-10546 BP (.988), 10536-10533 BP (.001) and 10523-10504 BP (.010) (9530±120 bp, Gif92355); 11173-10546 BP (.988), 10536-10533 BP (.001) and 10523-10504 BP (.010) (9530±120 bp, OxA-3533); 11068-10940 BP (.088), 10858-10825 BP (.015), 1080810796 BP (.005) and 10792-10237 BP (.891) (9350±120 bp, OxA-3534); 11037-11027 BP (.003), 10978-10973 BP (.001) and 10748-10189 BP (.996) (9260±120 bp, Gif-92362); 10665-10663 BP (.001), 10637-10613 BP (.015), 10594-10594 BP (.000) and 10581-10186 BP (.984) (9190±100 bp, Lv-1709). 68. The calibrated radiocarbon date is: 11107-11097 BP (.007), 11094-10576 BP (.988) and 10569-10562 BP (.005) (9500±75 bp, OxA-4917).

135

Figure 5.59 – Cnoc Coig (Argyll and Bute); (a) schematic section (note vertical exaggeration); (b) plan showing structures within midden and hearth areas (after Smith 1992: figure 8.8).

136

Figure 5.60 – Horizontal plot showing the distribution of all human skeletal material in the levels at Cnoc Coig (Argyll and Bute) (after Nolan 1986: figure 63).

137

138

Collective double+single double+single; single; triple single+double

8350±105 bp

8100±90 bp

9590±110 bp 9190±100 bp (6 dates)

9215±65 bp

9075±65 bp

9070±70 bp

7960±100 bp

9020±100 bp

8715±310 bp

Auneau

Grotte des Perrats

Grotte Margaux

La Vergne 10

La Vergne 3

La Vergne 7

Noyen-sur-Seine

Petit-Marais

Varennes

Upper layer contained a single articulated human skeleton; lower layer contained parts of 2 semi-articulated human skeletons.

Disarticulated human skeletal material, with many skeletal elements missing, carefully arranged.

Human skulls and other disarticulated human skeletal material of 4 individuals with traces of cut marks.

Two human skeletons plus an infant at the base of a pit, with cremated adult remains scattered over the infant

Three skeletons, 2 of which were seated inside a pit, at a higher level was incomplete and disarticulated human skeletal material, higher still was 3 articulated skeletons

Two human skeletons and an articulated skeleton at a higher level.

Disarticulated human skeletal material of at least 9 individuals in a pit at the back of the cave and on adjacent stone paving.

Disarticulated human skeletal material of 8 individuals inside the cave, with numerous cut marks indicating defleshing, plus human teeth marks.

An articulated skeleton in a sitting position at the base of a stone-lined pit.

Disarticulated human skeletal material of at least 8 individuals placed along the rear wall of the cave with a number of large stones at the entrance of the cave.

Description

Billard et al. 2001

Ducrocq et al. 1996

Auboire 1991

Duday and Courtaud 1998

Duday and Courtaud 1998

Duday and Courtaud 1998

Cauwe 1998b

Boulestin 1999

Verjux and Dubois 1996

Cauwe 1995, 1998a

Reference

Table 5.6 – Deposition of human skeletal material documented by recent investigations at ‘Mesolithic’ funerary contexts in northern France and Belgium (after Scarre 2002a: table 1)

single+double

Single

collective?

Collective

Single

Collective

9090±140 bp

Abri des Autours

Type of burial

Radiocarbon dates

Context

Figure 5.61 – Grotte Margaux: Plan of the early ‘Mesolithic’ cave (after Cauwe 2001: figure 6).

139

Figure 5.62 – Grotte Margaux: Distribution of human skeletal material in the cave (after Cauwe 2001: figure 7).

140

Figure 5.63 – Abri des Autours: Plan of the ‘Mesolithic’ cave (spelling from original)

141

‘Mesolithic’ come from occupation debris. It could be argued then that the dead were circulated among the living, in much the same way as outlined earlier for the ‘Neolithic’ (Bradley 1997a:14; Meiklejohn and Denston 1987; Meiklejohn et al. 1998; Schulting 1998b; Tilley 1996a; Whittle 1996a:202). Meiklejohn material was “intermingled with other features. A burial is just another feature of a mesolithic habitation site”. Further they conclude; “Far from segregating the dead in a specialized site or structure, people were clearly living between and on top of their recently deceased”. Tilley’s (1996a:109) work on the Scandinavian ‘Mesolithic’ comes to a similar conclusion, that there was little spatial division between the living and the dead. The evidence from the British and Irish ‘Mesolithic’, especially the detailed work from Cnoc Coig (Argyll and Bute), would suggest that a similar practice occurred. I would argue that a similar practice occurs in the ‘Neolithic’ but that this has generally been obscured due to a disparity between the methods and prior assumptions applied to the mortuary practices of these ‘two periods’, and the identification of privileged locations, namely ‘tombs’ and ‘barrows’, in the ‘Neolithic’ landscape.

identifies parallel treatment of human and animal material. The detailed examination by Nolan (1986) of the human and animal skeletal material from Cnoc Coig (Argyll and Bute) has identified apparently identical treatment of the two assemblages. He has identified similar depositional practices in relation to group 3, flipper bones of a seal, and group 2, containing mostly human hand bones, concluding: “In terms of formation processes, this apparently identical treatment of human and seal bones is curious” (1986:255). The analysis indicated that the predominance of human extremities (wrists, hands, ankles and feet) skeletal elements is reminiscent of the bias of skeletal extremities of red deer and pig. (after Cauwe 2001: figure 11). Nolan’s interpretation and conclusion deserves to be quoted in full: In short, the human bone assemblage from Cnoc Coig is an anomaly which cannot be readily explained at present. Indeed, were these bones identified as belonging to some other mammal, one would almost certainly conclude that they indicate an exploitation pattern similar to that of red deer and pig in which animals were killed and butchered elsewhere and selected portions of them brought to the site! Such a conclusion can scarcely be accepted unless one is willing to indulge in speculation about cannibalism in the late Mesolithic of western Scotland; and Meiklejohn and Denston [1987]…point out that the total absence of cut marks on the human bones from all the Oronsay shell middens (which totals 58 bones) argues against such an interpretation. (ibid.:256-7)

The disarticulated assemblages of human skeletal material, which are identified as evidence for a complex and multi-staged mortuary process involving exposure, dismemberment and circulating the dead amongst the living and through the landscape in the ‘Neolithic’, are seen as evidence of possible cannibalism in the ‘Mesolithic’. This is linked to the deep conceptual separation between the scholarly traditions of ‘Mesolithic’ and ‘Neolithic’ studies discussed in the Introduction. It is worth quoting Bradley’s satiric declaration again: in the literature as a whole, successful farmers [‘Neolithic’ populations] have social relations with one another, while hunter-gatherers [‘Mesolithic’ populations] have ecological relations with hazelnuts. (1984a:11) Recently, Bradley (1998a:21) has gone further, arguing that scholars have felt it possible to reconstruct ‘Neolithic’ ideology from assemblages of human skeletal material (Barrett 1994, 1996; Cooney 2000a; Edmonds 1999a; Lucas 1997; Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Thomas 1991a, 1996b, 1999a, 2000a; Thomas and Whittle 1986; Thorpe 1984, 1996; Whittle 1996a). He emphasises that little has been made of such concerns in the ‘Mesolithic’ where similar assemblages occur. However, since this review, Pollard (2000b) has argued that ‘Mesolithic’ occupation contexts with lengthy occupation could be termed ‘persistent places’ (Barton et al. 1995). He examines the contexts at Star Carr (North Yorkshire) and the middens on Oronsay (Argyll and Bute), identifying that the long histories of these contexts could have developed relationships between places and ancestry. Similar phenomenological approaches in relation to ‘Mesolithic’ contexts have been taken by Cummings (2000) and Warren (2000).

The cannibalism issue will be outlined and discussed later in the chapter, however it is interesting to note that in North and Central America, and some European contexts, similarities between the processing of human skeletal material and animal bones representing food refuse has long been identified as a diagnostic for recognising cannibalism (see Degusta 2000; Hurlbut 2000; Pepper 1920; Villa 1992; Villa et al. 1986; White 1992). Location of human skeletal material Human skeletal material radiocarbon dated and calibrated to between the beginning of the tenth and the middle of the seventh millennium BP can be found in caves (Figures 5.65), wet contexts, occupation debris and shell middens (Figures 5.59, 5.60 and 5.66). The predominance of disarticulated, fragmentary and incomplete human skeletal material, with some individuals represented by only a few fragments, could suggest that the arguments proposed for the general practice of movement, circulation and arrangement of human skeletal material in the British and Irish ‘Neolithic’ could occur in the ‘Mesolithic’. Living with the dead Many of the contexts mentioned as dated to the 142

Figure 5.64 – Skateholm II. Grave XXI with a buried dog. A deer antler lay along the back of the dog and three flint knives were found on its stomach (above). A decorated hammer of red deer antler was placed on the chest of the dog (below) (after Larsson 1989a: figure 7).

143

Figure 5.65 – Section and deposits of Ogof-yr-Ychen (Cave of the Oxen) (Pembrokeshire) (after Davies 1989a: figure 7.4).

144

Figure 5.66 – Feature location plan of the northern area at Ferriter’s Cove (Co. Kerry) (after Woodman et al. 1999: figure 2.1). become an accepted and supposedly demonstrable ‘fact’ over the last ten to twenty years, essentially tied up with the identification of chambered ‘tombs’ and earthen and timber ‘barrows’ as ‘houses of the ancestors’ (Barrett 1994, 1996; Bradley 1984a; Clarke et al. 1985; Cooney 2000a; Edmonds 1999a; Lucas 1997; Malone 1994; Mizoguchi 1995; Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Richards 1988; Shanks and Tilley 1982; Thomas 1991a, 1999a, 2000a; Thomas and Whittle 1986; Thorpe 1984, 1996; Whittle 1996a).

Mortuary archaeology in the ‘Neolithic’: The Creation of Ancestors The treatment and patterning of human skeletal material in the British and Irish ‘Neolithic’ is interpreted as an extensive mortuary process tied up with a concern for the ancestors and ritual. Ancestors, and the ritual behaviour associated with their identification, have increasingly 145

chambered mounds and earthen mounds as the medium that converts the dead, as articulated material, to the disarticulated material of the ancestor (also see Tilley 1996a:334, pages 53-5).

Despite the dispersed nature of human skeletal material and the circulation of these assemblages around the landscape, most scholars have allocated a unique significance to the assemblages in chambered ‘tombs’ and earthen and timber ‘barrows’. From Childe to Thomas, the practice has been to identify these contexts as the proper or special repositories for human skeletal material, as if they were mausoleum-like features in the ‘Neolithic’ landscape (King 2001:325). For example, when confronted with the empirical evidence that human skeletal material exists in both earthen long ‘barrows’ and causewayed enclosures in southern Britain, Childe (1949a:40 and 77) attempted to explain this patterning in terms of social status. Childe contrasted communal ‘burials’ of small chiefly elites in earthen long mounds with the unceremonial dumping of ordinary bodies into the ditches of causewayed enclosures. Thorpe (1984:43) comments that it was: Childe’s class-based view of society, which led him to expect there to be clear social divisions between a ruling class and the dispossessed, and to suppose that these would be reflected in the archaeological record.

This logic, derived from an anthropological perspective, has been used to characterise the transformation of the individual into the community of ancestors in Yorkshire (Lucas 1997), Britain as a whole (Barrett 1994, 1996; Thomas 2000a) and Ireland (Cooney 2000a). It is generally assumed that the construction of ‘monumental’ architecture automatically requires, and relates to, a belief system that encompasses ancestor worship (see Chapter 4). The archaeological literature concerning ‘Neolithic’ Britain and Ireland makes a potentially fallacious link between the practice of ancestor worship in small-scale societies in the ethnographic record, and the ideology of those societies of prehistoric Britain and Ireland who built stone, timber and earthen constructions. The archaeological literature draws heavily on anthropological literature identifying the importance of the disarticulation for the creation of the ancestor out of individual human corpses (see especially Cooney 2000a, 2000b; Thomas 1991a, 1999a, 2000a). However, in ‘Mesolithic’ contexts where disarticulation is the dominant form of processing the human corpse, such anthropological evidence and arguments are not developed. Currently only certain types of substantive analogies are drawn on, which pertain to ancestor worship. The disarticulated human skeletal material assemblages widely identified in both ‘Mesolithic’ and ‘Neolithic’ contexts could draw on other anthropological and ethnographic analogies as a starting point for the possible human social behaviour that may have produced them (see Parker Pearson and Ramilisonina 1998a:309; Tilley 1996a:1-6), one of which will be developed below. However, this is only an alternative explanation of equal validity. Many others are possible. My aim is to illustrate the potential variation of explanations of a given archaeological assemblage when different anthropological analogies are employed.

Although Childe’s particular interpretation has been challenged (see Piggott 1954), the prevailing discourse continues to pursue a similar kind of logic.69 While recent approaches have used a variety of interpretive and theoretical tools, a similar argument for specific kinds of social action and verbalised social meaning prevails (see Barrett 1994, 1996; Cooney 2000a; Edmonds 1999a; Lucas 1997; Malone 1994; Mizoguchi 1995; Parker Pearson 2000; Parker Pearson and Ramilisonina 1998a, 1998b; Richards 1988; Shanks and Tilley 1982; Thomas 1991a, 1999a, 2000a; Thomas and Whittle 1986; Thorpe 1984, 1996; Whittle 1996a). Thomas (1991a:185) seeks to differentiate between ‘ancestors’ and the ‘dead’. He argues that the ‘megalithic’ chambered mounds and earthen mounds converted the dead to the realm of the ancestors. He states that the ancestors could be removed from these structures and circulated around the landscape like “the relics of Christian saints” (1999a:136; also see Bradley 1998a:53-4; Edmonds 1997:106, 1999a:58-9; Gresham 1972; Woodward 1993). Thomas’ logic characterises the chambered ‘tombs’ and earthen and timber ‘barrows’ as the ‘proper’ or ‘privileged’ location of the dead in the landscape. In his discussion of mortuary practices in southern Britain, Thomas (1991a, 1999a) offers anthropological evidence for the importance of secondary burial, stating that the defleshing of the dead must occur “before the individual can join the community of the ancestors” (1991a:112; derived from Bloch 1982; Douglas 1966; Hertz 1960). Thomas (1991a:119-20) identifies the megalithic

Cooking up the Ancestors! Since the nineteenth century and the beginning of archaeological inquiry into the prehistoric inhabitants of Britain, archaeology has served to document the ‘barbarian’, ‘uncivilised’ progression towards the civilised ideal of the present. Cannibalism and human sacrifice were common Victorian explanations for the mass of fragmentary and disarticulated human skeletal material excavated from the many chambered ‘tombs’ and earthen and timber ‘barrows’ emptied by the Victorian era ‘barrow’ diggers (Pollard 1993:3). Thurnam identified cannibalism as present in archaeological assemblages, which became less prevalent as time progressed, “the older the date of any sepulchral monuments, the more likely are we to find in them traces of this practice” (1869:186). William Greenwell identified disarticulated human skeletal material as evidence of cannibalistic feasts (1865:107, 1877:544).

69. James Whitley (2002) argues that ancestors have replaced chiefs as the defining entity of prehistory.

146

identification of cannibalism. Their paper offers alternative interpretations to explain disarticulated and broken human skeletal assemblages, a recurring feature of the Pueblos during the twelfth century AD.

The identification of cannibalism in the ‘Neolithic’ continued into the early parts of the twentieth century, generally drawing on the recurring association between disarticulated and fragmentary human skeletal material (some with cut-marks), animal bones and occupation debris at contexts, such as at causewayed enclosures and in pits (Brothwell 1961; Childe 1949a:99; Curwen 1937:81; Lethbridge 1935; Mortimer 1905:xxiv, lxv; Piggott 1954:48; Tratman 1929; Wheeler 1943:21). For example, in relation to occupation debris and human skeletal material from Fifty Farm (Suffolk), Lethbridge (1935:126) states: It was interesting to notice that the flint scrapers were almost always found in little groups of three or four as if they had been just used for one operation and then thrown down. One cannot help thinking that this represents the clearing of a single skin on each occasion. Beyond any doubt they were cannibals. The human jaw and tibia turned up, as we scraped, in just the same way as the bones of pig or ox scattered on the occupied area. It is interesting to note that the victim, in spite of all we hear of the wonderful teeth of primitive man, suffered not only from pyorrhoea but caries also. Why did they eat such a weedy individual? However, such arguments are rarely considered currently in the ancestor-centric ‘Neolithic’ which dominates British archaeological discourse (see Whitley 2002).

Disarticulated, broken and sometimes burnt human skeletal material is a common feature of both ‘Neolithic’ contexts in southern Britain and southwest American Pueblo contexts during the twelfth century AD. While much variation occurs within the assemblages of these regions, distinctly different conclusions have been drawn about the prehistoric human social behaviour that led to their formation. However, distinguishing archaeologically between disarticulation for the veneration of the ancestor (excluding eating the corpse) and disarticulation (butchery) of a corpse for the purpose of food is problematic if not impossible, apart from possibly the recent human coprolite evidence from 5MT10010 (Billman et al. 2000; Lambert et al. 2000a), which has not been without its critics (Dongoske et al. 2000). Behaviourally, these acts are not very different and they may even fall into the same class of human social behaviour. One way forward would be to examine any archaeological differences between the assemblages, but the problem remains how to demonstrate how the attitudes and opinions correlated with the human action. We must see that the current notions of verbal meaning that prevail in the majority of archaeological discourse need not bind archaeological interpretation. Both the ancestor and the cannibalism conclusions draw heavily on ethnographic and ethno-historic analogies. It is argued here that the current use of substantive uniformitarianism to generate interpretations in relation to social action is problematic and allows for an array of unsubstantiated conclusions. The very nature of social analogy drawn from ethnographies and ethno-histories confines archaeological interpretation to human social behaviour that has been observed or documented. However, archaeology’s chronological scope is such that attempting to find parallels with human social behaviour that has been observed or historically documented, drastically reduces and underestimates the potential of archaeology. Moreover, the current postprocessual discourse argues for the contextual integrity of human cultures, while the interpretive tools currently used by postprocessualists infringes on the contextuality they advocate.

On the other side of the Atlantic the situation is quite different, with the archaeological identification of cannibalism at the centre of archaeological discussion in North American contexts (see Billman et al. 2000; Dongoske et al. 2000; Hurlbut 2000; Lambert et al. 2000a, 2000b; Novak and Kollmann 2000; Olsen and Shipman 1994; Pijoan Aguadé and Lory 1997). Central to this issue is the possible archaeological interpretations that can be drawn from an assemblage of disarticulated and broken human skeletal material. Billman et al.’s (2000) claim of cannibalism at 5MT10010, a small early Pueblo III context in southwestern Colorado in the United States of America, is based on the context and treatment of the human skeletal assemblage. The human skeletal material is described as “broken and disarticulated remains from a minimum of seven individuals...recovered from non-burial contexts in two of the three pithouses” (Billman et al. 2000:157). This assemblage of human skeletal material was found in association with stone tools and other occupation debris, including pottery sherds, animal bones and hearths (Figures 5.67, 5.68 and 5.69). The osteological evidence for cannibalism at 5MT10010 is based on the criteria set out by Turner and Turner (1995), which includes cut marks (Figure 5.70). An interesting additional avenue of research has been the study and identification of human myoglobin in the human coprolites found in the hearth of Feature 15 (Billman et al. 2000:166-7; Lambert et al. 2000a:402-4). Billman et al. (2000:167) state that this “is the first direct evidence of the consumption of human flesh in the prehistoric Southwest” (however see Dongoske et al. 2000:184-5). Dongoske et al.’s (2000) critique of Billman et al.’s report addresses the issue of the archaeological

Conclusions The study of the mortuary process automatically places the emphasis and focus on human skeletal material. The accompanying artefacts (pottery, stone tools, etc.) and the animal bones are thus only props, vital but subservient in relation to the primary concern in the mortuary process – human skeletal material. An attempt should be made to move away from identifying the mortuary process in its traditional sense, as it is too narrow an approach. Archaeological assemblages present a far more complex and problematic record of human social behaviour. The specific level and elements of human social behaviour 147

Figure 5.67 – Plan of 5MT10010 (Colorado, U.S.A.) (after Billman et al. 2000: figure 2).

148

Figure 5.68 – Plan of pithouse Feature 3, 5MT10010 (Colorado, U.S.A.) (after Billman et al. 2000: figure 3).

149

Figure 5.69 – Plan of pithouse Feature 13, 5MT10010 (Colorado, U.S.A.) (after Billman et al. 2000: figure 4).

150

Figure 5.70 – Cut marks on an adolescent femur from 5MT10010 (Colorado, U.S.A.) (after Billman et al. 2000: figure 7). drew reference to Classical and ethnographic sources for analogy (Pollard 1993:3). Such methods of interpretation have deeply affected the discourse, resulting in a situation whereby observations and declarations drawn from anthropological literature have become accepted facts, which underpin much of the discipline. The position that the corpse and death holds in the current discipline draws heavily on our Western/modern perceptions, fears and views of death, which are reinforced as a ‘universal phenomenon’ through the employment of anthropological literature. Thus, the interpretations are derived only from the social conventions being used and not from the data.

that are currently being asked of the assemblages of human skeletal material in Britain and Ireland cannot be answered by the empirical evidence, and require the use of secondary evidence drawn from anthropological sources. The practice is further complicated by the parallel treatment and deposition of human and animal skeletal material, particularly in relation to structures identified in the literature as ‘tombs’ and ‘barrows’. This practice has dominated and defined the British and Irish ‘Neolithic’ discourse over the last two hundred years. Victorian ‘barrow’ diggers, such as Thurnam (1869), Greenwell (1865, 1877) and Mortimer (1905),

151

CHAPTER SIX: MOVING AND LIVING AROUND THE LANDSCAPE ‘Neolithic’. This coupled with recent reappraisals of the settlement patterns of the ‘Neolithic’ in many areas of Britain and Ireland, has led to arguments that the ‘Neolithic’ populations were involved in a mobile residential system of occupation (Edmonds 1999a, 1999b; Evans et al. 1999; Harding 1995; Pollard 1999a, 2000a; Schofield 1995; Thomas 1991a, 1999a; Thorpe 1996; Whittle 1996c, 1997a, 1999a, Whittle et al. 2000, also see Kelly 1992). Whittle (1999a:63) has recently described the “transition from a mobile Mesolithic to a still mobile Neolithic”. Thomas (1999a:29) has argued that both lithic and faunal evidence suggests a high degree of mobility amongst the ‘Neolithic’ population of lowland southern Britain. However, it has rightly been argued that it is important not to attempt a generalised occupation model for the entirety of Britain and Ireland (Cooney 1997, 2000a:38, 2001; Pollard 1999a, 2000a).

At one level, archaeologists employ the term ‘settlement’ to characterize particular types of site, while at another it is used to describe the process by which a particular group of people inhabit or colonize a region. In both cases, ethnocentric notions of human behaviour are often uncritically projected into the past. (Brück and Goodman 1999:2) The previous chapter identified the ‘Mesolithic’ and ‘Neolithic’ movement and circulation of human and animal skeletal material. This chapter is concerned with evidence for a mobile residential system of occupation in the ‘Neolithic’. It will discuss issues of occupation, settlement and mobility. Artefact scatters will then be examined as possible signatures of a mobile residential system. It will be argued that identifying an artefact scatter as serving an ‘industrial’, ‘ceremonial’ or ‘domestic’ function assumes that such isolated practices will automatically be identifiable in archaeological assemblages and contexts. This may in fact be obscuring the actual patterning of human social behaviour across the landscape. Following this, a new procedure combining lead and strontium isotope analysis of archaeological human dental tissues, which may demonstrate the occurrence of residential mobility, will be discussed. Finally, the movement and circulation of artefacts across the landscape will be discussed, which may provide us with a window on the movement of human and animal skeletal material and the living population around the landscape. While this artefact movement does not necessarily mean that the majority of the population was mobile; it does indicate that the population was, to varying degrees, moving across the landscape. Throughout this chapter and the remainder of the book, the term mobility will refer to ‘residential’ mobility, people moving across the landscape, probably at different times of the year for subsistence and other reasons. If another form of mobility is being discussed I will indicate this.

In Britain, and to a lesser extent Ireland, investigation into artefact scatters has become the major focus in the study of settlement archaeology. While artefact scatters have been investigated since the beginning of the nineteenth century, if not earlier, they have not received the theoretical and methodological development of other aspects of archaeological assemblages and contexts, such as pottery, lithics or human skeletal material. Thousands of lithic assemblages have been collected across Britain and Ireland, many with poor provenance and have been deposited in museum collections with varying degrees of care, some even later discarded (Schofield 2000:46). The wealth of such material in museum collections has led to attempts to appraise the value of such assemblages and to incorporate them into occupation and land use patterns of the ‘Neolithic’ (Gardiner 1984, 1987a; Healy 1984; Holgate 1988a, 1988b; Schofield 1994c; Young 1987). The first significant archaeological investigation of artefact scatters occurred in the 1970s and 1980s (see Bradley 1995a:14; Schofield 2000:Table 5.1). Currently, artefact scatters are identified as the archaeological signature of a mobile ‘Neolithic’ population utilising various locations at different times during the annual economic and residential cycle (Pollard 1999a; Thomas 1997a; Whittle 1997a; but see Edmonds 1987, 1995:812).

Residential patterns in the ‘Neolithic’ From the late 1980s onwards, there has emerged a general acceptance that the lack of substantial structures, interpreted as being used for occupation, in most of the British ‘Neolithic’, and to a lesser extent in Ireland (Cooney 1997), is an empirical ‘reality’, rather than a product of taphonomic bias (see Edmonds 1993a Thomas 1991a, 1999a; Whittle 1997a; see below). Consequently, archaeological investigation has turned its attention to ‘alternative’ evidence for ‘Neolithic’ occupation. As outlined in Chapter 2, over the last thirty years evidence has been identified which illustrates continuity in subsistence practices from the ‘Mesolithic’ to the

The systematic incorporation of artefact scatters into the archaeological discourse has only just begun (see Austin and Sydes 1998; Barrett et al. 1991; Gillings 2000; Hey 1998; Holgate 1987a; Schofield 1988, 1991a, 1991b), with the issue of the degree of compatibility of the evidence from artefact scatters remaining a major hurdle (Schofield 1995:105, 2000:48). Thus, while their potential has not yet been fully realised or even understood, the opportunity exists for further integration of the mobile residential system of ‘Neolithic’ Britain and Ireland into a holistic approach to the activities of people 152

evidence at some of these sites for rebuilding of the structures on the same spot.

on the landscape (Schofield 2000). However, I would argue further that other aspects of human social behaviour, namely the patterning of skeletal material (both human and non-human), across the landscape must also be reinterpreted in the light of this new acceptance of a mobility model for the ‘Neolithic’ (see Chapter 5). The evidence outlined for the movement of material across vast distances and the existence of artefact scatters both indicate a general movement of artefacts across the landscape and a model of dispersed land use. This dispersed land use may also be evidenced by the continued utilisation and limited woodland clearances in the ‘Neolithic’ of Britain and Ireland (see Chapter 2). However, this evidence does not necessarily indicate mobility, as these patterns could occur in a situation of sedentary residency. This is part of a larger problem of identifying the different archaeological signatures left by sedentism and mobility (see Binford 1980, 1982; Close 2000; P.C. Edwards 1989; Kelly 1992). Even within a sedentary society, not all people are sedentary. They may be mobile throughout the year or only at certain times of the year. All this suggests that such arguments should be termed in relation to the degrees of mobility.

Archaeologically the identification of the level of mobility in a society or population is difficult. However, new scientific techniques based on the chemical composition of human skeletal material may provide an empirical way of identifying mobility in individuals, which may serve as a guide to more general patterns of mobility (see below). It is interesting to note the level of evidence required in ‘Neolithic’ contexts for the archaeological discourse to accept that some degree of mobility occurred. For so long ‘Neolithic’ populations were identified as farmers living in solidly built houses arranged in villages, tending to the crops and flocks, within our own idealised farming landscape (Gijn and Zvelebil 1997:3; Whittle 1997a). Such arguments were accepted on blind faith despite the lack of substantially built houses. ‘Residential’ mobility, on the level I will argue for ‘Neolithic’ contexts is a generally accepted principle for ‘Mesolithic’ contexts. Zvelebil and RowleyConwy (Rowley-Conwy 1984; Zvelebil 1981; Zvelebil and Rowley-Conwy 1984) have argued for a system of varied residential locations, including some sedentary communities dispersed across the landscape. These claims have been made in relation to the large Ertebølle shell middens of southern Scandinavia. However, by this, Zvelebil and Rowley-Conwy mean no more than the sedentism of the North American North Coast indigenous populations, as it has been pointed out that this is actually a form of mobility (Fletcher 1995:166, Friedman 1976). I will present more evidence than is generally offered in ‘Mesolithic’ contexts. Again the different approaches of the two discourses are highlighted, which has continued to insinuate difference while playing down continuity.

The archaeological identification of mobility …none but the largest !Kung dry-season camps would be found by the future archaeologist. (Yellen 1977:80) In the last few decades, archaeologists have discussed issues of mobility in relation to the prehistoric past. Mobility in the archaeological literature has little to do with the ability to move, rather it relates to the act of moving (see Close 2000:50; Kelly 1992:44). However, how this ‘moving’ translates archaeologically has long been an issue of discussion. Moving usually leaves no material trace archaeologically, thus, as Kelly (1992:54) comments, “it is difficult to study mobility archaeologically”. Acknowledging this, Close (2000:50) has argued that this leads to the reconstruction of “mobility strategies” or “patterns of mobility”, rather than the identification of individual cases of mobility in the past.

What do we mean by ‘settlement’ and ‘occupation’? The term ‘settlement’ conjures up images of permanently bound spaces with connotations of sedentism and permanency (Pollard 2000a:363), a view which I do not subscribe to and which will be discussed in detail. It leads to terms such as ‘house’, ‘domestic practice’ and ‘household’, creating a domesticity in the past drawn directly from our own social reality. Brück (1999a) states that the distinct domestic sphere in modern British (Western) society is a result of industrialisation and secularisation, among others things. This reality is viewed as a cross-cultural, universal form of human social behaviour that will necessarily be identifiable in archaeological assemblages and contexts (Brück 1999a:61; Brück and Goodman 1999:1; Thomas 1996a:3). Cooney (2000a:52) argues that the concept of house or dwelling is a cross-culturally identified entity (Kent 1990; MacEachern et al. 1989; Samson 1990; Wilson 1988), but that the concept of ‘home’ differs between societies (Kent 1995; Rapoport 1995; Stea 1995). Domestic activities, such as food preparation and consumption, biological reproduction and childcare, are viewed as the primary feature of settlement, and

P.C. Edwards (1989) has argued that the many categories of archaeological evidence proposed as explicit signatures of sedentism (limited mobility) may be inclusive of medium to highly mobile populations. For example, he identifies that mobile populations invest energy and time constructing large structures, which may or may not be used for occupation (see page 55). However, such a view has yet to infiltrate the ‘Neolithic’ archaeological discourse. For example, Monk (2000:80) has recently argued: While building structures per se do not indicate sedentism, the size of these buildings [in the British and Ireland ‘Neolithic’] and the building timbers used, mainly oak, suggests a level of investment in energy unlikely to be expended by a nomadic or mobile society. In addition, there is 153

ritualistic, connected to the mortuary process.70 The situation identified by Brück (1999a) in relation to the early ‘Bronze Age’, also occurs in the ‘Neolithic’. The definition of ‘settlement’ that occurs throughout the literature cannot be systematically identified archaeologically as it relates to our particular social reality. Identifying a context as a ‘house’ or ‘domestic’ is problematic and assumes that discrete entities in our social world should be immediately recognisable in past human social behaviour. Thus, as Brück and Goodman (1999:3) argue, “The recognition of settlement sites is therefore dependent on our ability to identify domestic practice in the archaeological record.” The core problem is that the postprocessual (contextual) approach was not sufficiently contextual and too subject to its practitioner’s sociality.

conversely settlements are the place where domestic activity is performed. Thus, settlements are “generally considered to be a spatially and functionally distinct type of site…distinguishable from categories of locale whose major roles lie in other realms, such as ritual, political or economic practice” (Brück 1999a:55, original emphasis, also see Carman 1999). While most archaeologists use the term ‘settlement’ and have an understanding of what it means, the situation is quite complicated and problematic (see Lane 1986 and below). Part of the confusion is that ‘settlement’ is used when discussing occupation/domestic activities, while the reality is that settlement locations encapsulate every activity of human society. Settlement tends to be equated with houses, drawing on our Western/modern notion of domestic dwellings. We identified structures in the past as houses which are comparable to our domestic architectures (Richmond 1999:58) and marginalise that which does not fit into our perception and understanding of domestic architecture. Carman (1999) has identified the inherent bias in the interpretation of evidence related to occupation in prehistoric Britain. He cites the Ordnance Survey (1973) publication, Field Archaeology in Britain and the English Heritage (1991) publication, Exploring our Past, where ‘settlements’ are not discussed “until there are ‘established farming communities’ in the Neolithic” (1999:24) (however see Chapter 2). Carman states that the term ‘occupation’, rather than ‘settlement’, is used in earlier periods.

The assumption that a discrete set of ‘domestic practices’ should be identifiable archaeologically relies on the universal nature of such assemblages. Identifying a ‘house’ or ‘domestic space’ in the prehistoric archaeological contexts relates to a Western/modern ideological construct, the ‘home’ and the importance of the hearth in this construct (Brück and Goodman 1999). A way forward is to identify assemblages that are the result of human occupation regardless of their context, producing a general and preliminary picture of the occupation patterns of prehistoric Britain and Ireland. This method will allow equal weight to similar assemblages of artefacts. While this must account for the deposition process, especially secondary deposition in the past, I will argue that many contexts which have produced occupation debris have not been considered part of the occupation history of the population, as they are thought of as being out of context. This has produced a false picture of ‘Neolithic’ life, which has aimed to characterise their human social behaviour as comparable to a familiar (modern) ordering of human activity. The identification of mobility amongst populations is not a problem in a global perspective (see Kelly 1992), but there are issues of how the British want to see their past and their ancestors (see Chapter 1). Thus the discourse has ‘missed’ unparalleled forms of human social behaviour for so long because they are not a reflection of our Western/modern human social behaviour, be that historically documented, observed or assumed/imagined.

The situation is made more problematic due to the current interpretive framework about different human social behaviour across the landscape, which occurs throughout the ‘Neolithic’ discourse in Britain and Ireland (Brück 1999a; Lane 1986). For example, let us start with an assemblage of pottery sherds, lithic material, animal bones and charcoal, which I would identify as occupation debris, that is, material that would be produced by human occupation of a given area over time. The problem is that interpretation of what specific form of human social behaviour produced this in ‘Neolithic’ contexts is governed by the location of the assemblage. Excavating such an assemblage from the remains of a timber structure, such as Lismore Field (Derbyshire) or Tankardstown 2 (Co. Limerick), would lead to the identification of the structure as a ‘house’ and the assemblage as occupation debris or settlement debris. Similar material identified from within chambered ‘tombs’, such as West Kennet (Wiltshire) or Isbister (Orkney Islands), would not result in the structure being identified as a house, rather the remains would be seen as

70. The situation at Skara Brae (Orkney) provides an interesting case, which has been outlined earlier (see page 117).

154

towards so-called ‘non-domestic’ constructions, such as henges, long mounds and causewayed enclosures (Figure 6.1) (Darvill 1996a:82; Thomas 1996a:2). However, while occupation debris is found across the landscape such as from causewayed enclosures (Andersen 1997:242-67; Edmonds 1993a Oswald et al. 2001:124-6) and henges (Clare 1986; Harding and Lee 1987), this is not deposited in structures that are identifiable ‘houses’, even though pits and post-holes have been identified. Thus, this evidence has been viewed as indicative of seasonal and temporary activity (Oswald et al. 2001:1301), suggesting that the ‘permanent occupation’ settlements were yet to be found (however, see Bradley 1982; Wainwright 1975). The preoccupation with the search for settlement structures is linked to the standard view of the ‘Neolithic’ discussed previously (Bradley 1998a:Chapter 1; Thomas 1993, 1996a) – almost as if archaeological investigation was searching for an English village (see Brück 1999a; King 2001:327-8).

Born Dead?: Settlement structures and occupation in Britain and Ireland …our knowledge of prehistoric and early people is derived chiefly from their funeral deposits, and for all we know of their mode of life, excepting such information as has been obtained from lakedwellings, and crannogs, they might as well have been born dead. (General Pitt Rivers, 1892, cited in Clark 1937:468. Clark’s emphasis) The study of occupation in Britain and Ireland during the middle of the seventh to the late fifth millennium BP has been influenced and hindered by our perceptions, expectations, underdeveloped theorisation and the elusive nature of the evidence (Pollard 1999a:78, 2000a:363; Thomas 1996a). The archaeological literature in Britain, particularly in southern Britain, is dominated by the view that much more is known about the dead than the living, encapsulated by Pitt Rivers’ above statement. Prior to the 1960s, the lack of settlement evidence was viewed by many archaeologists as indicating that the population were nomadic pastoralists living in temporary structures that left little or no archaeological trace (see Childe 1949a). During the 1960s, 1970s and into the 1980s, more evidence came to light which was interpreted as the cultivation of cereal crops, such as plough marks and grain impressions on pottery and pollen. While the British and Irish ‘Neolithic’ was identified as a sedentary mixed farming system there was still a general lack of the substantial occupation structures (‘houses’), which were presumed to accompany agriculture (for example Burgess 1980; Grinsell 1963; Megaw and Simpson 1979). This led to the belief that the permanent settlements (‘houses’) lay buried under colluvium and alluvium just waiting to be excavated (Pollard 1999a:78, also see Bradley 1998a:Chapter 1): it is not surprising that few domestic sites have been recovered in view of the enormous accumulation of hill wash and other material that would have built up over the last 5000 years in the valley bottoms over such settlement sites. (Megaw and Simpson 1979:86) However, as noted by Thomas (1996a:3), large settlement structures are known from the middle ‘Bronze Age’ (midfourth millennium BP). If such depositional factors affected the second half of the seventh, the sixth and the fifth millennium BP, it is curious that middle ‘Bronze Age’ contexts were unaffected. However, there is geoarchaeological sedimentary evidence for considerable landscape changes due to erosion and depositional processes in the chalkland and limestone regions of southern Britain since the ‘Neolithic’ (Allen 1988, 1992; Bell 1982, 1983; Bell and Boardman 1992, also see Monk 2002).

Across Britain and Ireland, timber and stone structures have been excavated that have been identified as ‘houses’ (Barclay 1996; Cooney 2000a; Darvill 1996; Darvill and Thomas 1996; Gibson 1996; Grogan 1996; Hey 2001; Richmond 1999). In a recent study of ‘Neolithic’ ‘houses’ across northwest Europe, Barclay (1996:61) chose to “use the term “building” rather than “house” because of the difficulties of drawing distinction between some domestic structures and buildings which may have had different functions”. Despite this, they are generally identified as ‘houses’ despite the inconsistencies already pointed out. Scottish examples of ‘houses’ are Balbridie (Aberdeenshire) (Fairweather and Ralston 1993) (Figure A3.7), Barnhouse (Figures A3.21, A3.22 and A3.23) (Barclay 1996; Richards 1992b), Skara Brae (Figure 5.45), Rinyo (Figure A3.170), Links of Noltland (Figure A3.121), Knap of Howar (Figure A3.112) (Orkney Islands) (Clarke and Sharples 1985; Kinnes 1985; Ritchie 1983, 1985a; Simpson 1971), “Benie Hoose” (Figure 6.2), Standing Stones of Yoxie (Figure 6.3), Gruting School (Figure 6.4), Scord of Brouster (Figure 6.5), Stanydale ‘Temple’ (Figure 6.6), Ness of Gruting (Figure 6.7) (Shetland Islands) (Barclay 1996; Whittle 1985a; Whittle et al. 1986), Eilean Domhnuill (Western Isles) (Figure A3.79 and A3.80) (Armit 1991; Barclay 1996) and a recent find at Claish Farm, near Callander (Perth and Kinross), associated with more than 200 early ‘Neolithic’ round-bottom pottery sherds (Figure 6.8) (Barclay et al. 2002; Denison 2001; Rowley-Conwy 2002). In the same volume, Grogan (1996) stated that 50 ‘houses’ have been identified in Ireland. More recent discoveries has lead Grogan (2002:517) to state that in excess of “90 structures have been identified as probable houses dating to the Neolithic period in Ireland”. Examples of these are Tralee, Cloghers (Co. Kerry), Corbally (Co. Kildare), Coolfore (Co. Louth), Enagh (Co. Derry), Inch (Co. Down), Newtown (Figure A3.144), Knowth structures 6 and 8 (Figure A3.114) (Co. Meath), Tankardstown 1 and 2 (Figure A3.185), Lough Gur B (Figure 6.9), Lough Gur K2 (Figure 6.10), Lough Gur C (Figures 6.11 and 6.12), Slieve Breagh (Figure 6.13) (Co.

Substantial wooden ‘residential’ structures are not totally absent from southern Britain in the ‘Neolithic’ (see below). A 1970s research project in the United Kingdom aiming to unearth buried occupation structures, found a number of buildings which were identified as serving a ‘domestic’ purpose, but also an even greater imbalance 155

Figure 6.1 – Bar chart showing the discovery rate of ‘Neolithic’ buildings in England, Wales and the Isle of Man (after Darvill 1996: figure 6.2).

Figure 6.2 – Ground plan of the building at “Benie Hoose” (Shetland Islands) (after Barclay 1996: figure 5.2).

156

Figure 6.3 – Ground plan of the ‘Neolithic’ building at Standing Stones of Yoxie (Shetland Islands) (the black marks are upright stones and the darker-toned area is a later alteration) (after Barclay 1996: figure 5.1).

Figure 6.4 – Ground plan of the ‘Neolithic’ building at Gruting School (Shetland Islands) (black spots are post-holes, post-settings or post-supports) (after Barclay 1996: figure 5.1).

157

Figure 6.5 – Ground plan of the ‘Neolithic’ building at Scord of Brouster 3 (Shetland Islands) (the cross-hatched area is a hearth and the black spots are post-holes, post-settings or post-supports) (after Barclay 1996: figure 5.1).

Figure 6.6 – Ground plan of the ‘Neolithic’ building at Stanydale ‘Temple’ (Shetland Islands) (the black marks are upright stones, except where marked ‘p’ which are post-holes) (after Barclay 1996: figure 5.1).

158

Figure 6.7 – Ground plan of the ‘Neolithic’ building at Ness of Gruting (Shetland Islands) (the cross-hatched area is a hearth) (after Barclay 1996: figure 5.1).

Figure 6.8 – Plan of the structure at Claish Farm (Perth and Kinross) (after Barclay et al. 2002: figure 3).

159

Figure 6.9 – Plan of the structure at Lough Gur B (Co. Limerick) (after Cooney 2000a: figure 3.1).

Figure 6.10 – Plan of the structure at Lough Gur K2 (Co. Limerick) (after Cooney 2000a: figure 3.1).

160

Figure 6.11 – Plan and section of Lough Gur C (Co. Limerick) (after Cooney 2000a: figure 3.3).

161

Figure 6.12 – Plan of the Lough Gur C complex (Co. Limerick) (after Cooney 2000a: figure 3.4).

162

Figure 6.13 – Plan of the structure at Slieve Breagh 1 (Co. Meath) (after Cooney 2000a: figure 3.2).

Figure 6.14 – Plan and activity phases of Ballyharry (Co. Antrim) (after Cooney 2000a: figure 3.5).

Figure 6.15 – Plan of the structure at Ballygalley 1 (Co. Antrim) (after Cooney 2000a: figure 3.1). Darvill (1996) has identified 64 ‘house’/buildings in England, Wales and the Isle of Man, with some more recent examples also being excavated. Examples of these included Fengate (Cambridgeshire) (Darvill and Thomas 1996; Pryor 1978), Lismore Fields (Derbyshire) (Garton 1987), Rucklers Lane (Hertfordshire) (McDonald 1993), Barford structure C (Warwickshire) (Figure A3.16),

Limerick), Ballyharry (Figure 6.14), Ballygalley 1 (Figure 6.15) (Co. Antrim), Ballynagilly (Co. Tryone) (Figure A3.15) and Ballyglass 1 (Co. Mayo) (Figure A3.14) (Anon 2000; Cooney 1999a, 2000a; Crothers 1996; Eogan and Roche 1997; Gowen 1987; Grogan 1996, 2002; Grogan and Eogan 1987; Ó Ríordáin 1954).

163

Figure 6.16 – Plan of the ‘Neolithic’ structure at White Horse Stone (Kent) (after Oxford Archaeological Unit 2000: page 451). (Figure 6.24) (Coles 1971; Megaw and Simpson 1979:669); Deepcar (South Yorkshire) (Figure 6.25) (Radley and Mellars 1964); Cnoc Sligeagh and Cnoc Coig (Argyll and Bute) (Mellars 1987). Recent excavations at Dunbar (East Lothian) and Howick (Northumberland) have unearthed circular buildings. At Dunbar (East Lothian) a nearcircular building marked by deep post-holes has been identified along with an ember pit or hearth, a few fragments of burnt animal bones, including one from a bird, large quantities of ‘Mesolithic’ flints, such as crescent-shaped microliths, scrapers and points (Denison 2003). On the Northumberland coast at Howick a circular, sunken-floored building marked by post-holes and stake-holes has been discovered. Finds include hearths, charred hazelnut shells, burnt animal bones, consisting of wild pig, fox, birds, dog/wolf and possibly a bear, and over 16000 lithics (some burnt), including scalene triangles and backed blades, scrapers, awls, cores and flakes. Structural evidence from the building indicates that it was completely rebuilt on the same site twice and 21 radiocarbon dates have been obtained between the late eleventh sixth and the tenth millennium

White Horse Stone (Kent) (Figure 6.16) (Oxford Archaeological Unit. 2000; Rowley-Conwy 2002), Hembury (Devon) (Figure A3.95), Gwernvale (Figure A3.89), Trelystan (Powys) (Britnell 1982) and Yarnton (Oxfordshire) where a two-roomed structure with a hearth has been excavated (Hey 2001; Rowley-Conwy 2002). One of the examples identified was a ‘Mesolithic’ example from Bowman’s Farm (Hampshire) (Green 1991, 1996). From Bowman’s Farm, three subrectangular structures and one sub-circular structure were excavated with associated artefacts identified as ‘Mesolithic’ and a sample of oak has been radiocarbon dated to 6950-6500 BP (1.000) (5910±90 bp, OxA-263) (Figure 6.17). Many identified structural remains, especially those in England, Wales and the Isle of Man, consist of post-holes and stake-holes arranged in a circular or linear form. These are generally interpreted as windbreaks or shortterm structures (Healy 1988), with such features and contexts also being identified in ‘Mesolithic’ contexts. ‘Mesolithic’ examples of post-hole and stake-hole structures can be found at Mount Sandel (Co. Derry) (Figure 6.18) (O’Kelly 1989:17-22; Woodman 1985a), Culverwell (Dorset) (Palmer 1990, 1999), Llyn Aled Isaf (Denbigshire) (David 1990:193-4), Frainslake (Pembrokeshire) (Gordon-Williams 1926); Broomhead Moor V (South Yorkshire) (Figure 6.19) (Radley et al. 1974); Lismore Fields (Derbyshire) (Garton 1987; Myers 2001); Abinger (Surrey) (Leakey 1951); Bolsay Farm (Argyll and Bute) (Figures 6.20 and 6.21) (Mithen et al. 2000a, 2000b); Kindrochid (Argyll and Bute) (Figure 6.22) (Marshall and Mithen 2000); Broom Hill (Hampshire) (Figure 6.23) (O’Malley 1976, 1978); Downton (Wiltshire) (Higgs 1959, Pollard 1999a); Monk Moors (Cumbria) (Bonsall 1976); Morton on Tay (Fife) 164

BP

71

et al. 2003).

(Dension 2003; Howick Project 2003; Waddington

These structures and structural remains generally contain assemblages that would be identified as the result of human occupation. However, similar assemblages are found throughout other contexts, including henges, causewayed enclosures, enclosures, pits, ditches, open contexts, cursus, stone chamber structures (‘tombs’) and earth and timber mounds (‘barrows’) (see the Online Database and Appendix 6). Many chambered ‘tombs’ contain assemblages which, if discovered in more acceptable contexts for occupation, such as in a rectangular structure which resembled a house, would be identified as occupation debris. It is important to note that human skeletal material is found with other occupation debris, suggesting that the two are not mutually exclusive. As far as I can discern, ‘tombs’ are not seen as suitable for living in because they were cramped, have little ventilation and generally contain human skeletal material. However, it is interesting to note that it is claimed that during the Scottish ‘Iron Age’, ‘Neolithic’ chambered ‘tombs’ were re-modelled and re-used for domestic purposes (Figure 6.26) (Hingley 1996, 1999), which demonstrates that they could feasibly be used for occupation. However, occupation debris in ‘tombs’ and ‘barrows’ is not systematically identified as such. Rather, it is interpreted as the result of ritual practice, which is generally seen as separate from, and exclusive of, occupation or domestic human social behaviour (Brück 1999a, 1999b; Lane 1986).

71. The twenty one calibrated radiocarbon dates are: 10185-9889 BP (0.945), 9878-9868 BP (0.012), 98469822 BP (0.021) and 9808-9785 BP (0.022) (8890±45 bp, OxA-11829); 10147-10135 BP (0.011), 10112-10076 BP (0.080), 10071-10058 BP (0.009), 10031-10018 BP (0.014), 10010-9993 BP (0.018), 9954-9680 BP (0.862) and 9640-9632 BP (0.006) (8802±38 bp, OxA-11804); 10147-10135 BP (0.009), 10112-10076 BP (0.062), 10071-10058 BP (0.008), 10031-10018 BP (0.012), 10010-9993 BP (0.015), 9954-9656 BP (0.856), 96529626 BP (0.026) and 9619-9601 BP (0.011) (8790±45 bp, OxA-11853); 10146-10137 BP (0.006), 10111-10077 BP (0.055), 10069-10061 BP (0.005), 10030-10019 BP (0.009), 10009-9994 BP (0.012), 9947-9927 BP (0.011), 9923-9625 BP (0.882), 9623-9600 BP (0.017), 95869582 BP (0.002) and 9566-9563 BP (0.002) (8785±45 bp, OxA-11828); 10146-10137 BP (0.006), 10111-10077 BP (0.055), 10069-10061 BP (0.005), 10030-10019 BP (0.009), 10009-9994 BP (0.012), 9947-9927 BP (0.011), 9923-9625 BP (0.882), 9623-9600 BP (0.017), 95869582 BP (0.002) and 9566-9563 BP (0.002) (8785±45 bp, OxA-11856); 10146-10139 BP (0.004), 10111-10078 BP (0.047), 10068-10063 BP (0.003), 10029-10020 BP (0.007), 10008-9996 BP (0.009), 9935-9929 BP (0.003), 9922-9598 BP (0.917), 9588-9578 BP (0.005) and 95699560 BP (0.005) (8780±45 bp, OxA-11832); 1010510095 BP (0.012), 9913-9625 BP (0.951), 9623-9600 BP (0.030), 9586-9582 BP (0.004) and 9566-9563 BP (0.004) (8763±38 bp, OxA-11803); 9910-9896 BP (0.023), 9893-9599 BP (0.955), 9588-9579 BP (0.011) and 9569-9560 BP (0.011) (8754±38 bp, OxA-11802); 10104-10103 BP (0.001), 9911-9895 BP (0.025), 98949596 BP (0.921) and 9591-9556 BP (0.052) (8750±45 bp, OxA-11857); 9890-9877 BP (0.033), 9869-9845 BP (0.074), 9838-9805 BP (0.074), 9803-9596 BP (0.742) and 9592-9556 BP (0.077) (8734±37 bp, OxA-11801); 9889-9878 BP (0.031), 9869-9845 BP (0.068), 98369805 BP (0.067) and 9802-9554 BP (0.835) (8730±40 bp, Beta-153650); 9890-9877 BP (0.025), 9869-9845 BP (0.055), 9838-9805 BP (0.059) and 9802-9549 BP (0.861) (8715±50 bp, OxA-11830); 9889-9878 BP (0.023), 9869-9845 BP (0.049), 9835-9831 BP (0.003), 9824-9805 BP (0.039) and 9801-9549 BP (0.885) (8715±45 bp, OxA-11831); 9889-9878 BP (0.019), 98689846 BP (0.044), 9824-9805 BP (0.035) and 9799-9548 BP (0.902) (8710±45 bp, OxA-11854); 9888-9880 BP (0.013), 9866-9848 BP (0.031), 9822-9807 BP (0.027), 9794-9794 BP (0.001), 9788-9785 BP (0.004) and 97839547 BP (0.925) (8700±45 bp, OxA-11827); 9815-9813 BP (0.001) and 9749-9531 BP (0.999) (8650±45 bp, OxA-11855); 9701-9529 BP (1.000) (8630±40 bp, OxA11826); 9682-9642 BP (0.053), 9631-9468 BP (0.925)

In relation to Irish ‘court-tombs’, in addition to material that could be attributed to occupation, hearths have identified in forecourts and evidence of in situ fires in the chambers has also been noted (Case 1973; C.S. Jones

and 9449-9433 BP (0.022) (8555±60 bp, AA-41788); 9470-9446 BP (0.075), 9436-9434 BP (0.003), 94349254 BP (0.870) and 9170-9145 BP (0.052) (8324±37 bp, OxA-11805); 9407-9403 BP (0.128), 9336-9331 BP (0.130), 9232-9221 BP (0.278), 9186-9181 BP (0.153) and 9131-9095 BP (0.311) (8278±35 bp, OxA-11806) and 9399-9378 BP (0.045), 9372-9357 BP (0.032), 93509341 BP (0.016), 9325-9312 BP (0.013), 9303-9085 BP (0.852), 9077-9069 BP (0.007) and 9050-9031 BP (0.035) (8233±36 bp, OxA-11807).

165

Figure 6.17 – Pre-excavation ground plans for ‘Mesolithic’/‘Neolithic’ structures at Bowman’s Farm (Hampshire) (after Green 1996: figure 7.3).

Figure 6.18 – Mount Sandel Upper (Co. Derry): ‘Hut’ A showing hearths a, b, c and d; pits 1 and 2; postholes in solid black (after O’Kelly 1989: figure 9). 166

Figure 6.19 – Plan of Broomhead Moor Site 5 (South Yorkshire), showing stake-holes (black) and hearth areas (stippled) (after Radley et al. 1974: figure 2).

Figure 6.20 – Location, plan and profiles of stake-holes in trench 1, Bolsay Farm (Argyll and Bute) (after Mithen et al. 2001a: figure 4.9.23).

167

Figure 6.21 – Plan and profile of stake-holes F40 and F31, trench II, Bolsay Farm (Argyll and Bute) (after Mithen et al. 2001b: figure 4.10.19).

Figure 6.22 – Kindrochid (Argyll and Bute): Plans and profiles of features in areas 1 and 2 (after Marshall and Mithen 2001: figure 4.7.8). 168

Figure 6.23 – Plan of the ‘Mesolithic’ structure at Broom Hill (Hampshire) (after O’Malley 1978: page 118).

Figure 6.24 – Morton on Tay (Fife): Plan of occupation II (after Megaw and Simpson 1979: figure 2.28). 169

Figure 6.25 – Deepcar (South Yorkshire): The distribution of local and foreign stones; quartzites (black), river-worn grits (stippled), local flags (plain) and three possible hearths are indicated (hatched). The section, the flint horizon (lightly stippled) and disturbed areas (shaded) are recorded (after Radley and Mellars 1964: figure 2).

Figure 6.26 – ‘Neolithic’ chambered structures rebuilt in the ‘Iron Age’: a) Quanterness (Orkney Islands); b) Howe Phase 3/4 (Orkney Islands); c) Howe Phase 5 (Orkney Islands); d) Clettraval (Western Isles); e) Unival (Western Isles) (after Hingley 1996: figure 2). 170

1997:75). Furthermore, C.S. Jones (1997), in relation to Irish ‘wedge-tombs’, states that material from these structures is similar to habitation contexts. This is particularly apparent at Lough Gur (Co. Limerick) where it is stated “part of the chamber had been used as a habitation” (Brindley and Lanting 1991-92a:24). In attempting to explain the similarity, C.S. Jones, proceeds, stating: There may have been a conscious effort by the tomb-builders to incorporate the material refuse of daily living as well as bones of the dead into the very fabric of the tomb. We should not be surprised then if some of the rituals performed at the tombs perhaps mimicked the activities carried out in habitation sites. (1997:97, emphasis mine) Here, Scarre’s (1983:266) comment on assemblages from French contexts is also appropriate: First, there are those [settlements] which consist of surface scatters of flint and potsherds. These are difficult to understand and evaluate, since without excavation the precise nature of the occupation is uncertain and some, represented only by material in old collections, may in reality have been burial monuments. Such logic is present throughout the literature based on the assumed dichotomy between ‘ritual’ and ‘domestic’ (see Brück 1999a, 1999b; Lane 1986) and on the

liberally applied to monuments, while not always applied to clusters of artefacts. Dunnell is not attempting to question the ontological validity of sites (ibid.:25-6), however, he indicates that the existence of the term “imparts a serious and unredeemable systematic error in the recording and management programs” (ibid.:36). Thus, the Western/modern notions of a ‘site’ as a distinct and spatially bounded entity does not necessarily translate into archaeological contexts.

presumed appropriate arena for humans to conduct domestic/occupation practice. The occupational component of the artefact assemblages from ‘tombs’ is marginalised. In the archaeological literature, the evidence of occupation debris in ‘tombs’ and ‘barrows’ is identified not as occupation debris but as the debris of ritual (Ashbee 1976; Case 1969, 1973; Herity 1987; C.S. Jones 1997:76).

The archaeological investigation of artefact scatters is founded on models of ‘off-site’ or ‘non-site’ archaeology. This is based on the assumption that human activity is not confined to points in the landscape, but is rather spatially continuous (see especially Dunnell and Dancey 1983; Foley 1981a, 1981b, 1981c). However, ‘non-site’ archaeology argues that intensity and tasks/activities will vary across the landscape, with archaeological signatures that can identify these different tasks identified on a regional distribution. Thus this approach argues that different human activities, such as ‘domestic’, ‘industrial’ and ‘ceremonial’/‘ritual’, will themselves be differentiated across the landscape. This model has been the basis of artefact scatter studies in Britain and Ireland due to the view that if a regional perspective is taken, one may ‘read off’ different human activities at different locations (Bates 1998; Ford 1987a, 1987b; Light et al. 1994; Schofield 1991a, 1991b, 1994a; Shennan 1985; Young 1987, also see Pollard 1998).

The concerns of Dunnell, Foley and others in relation to ‘off-site’/‘non-site’ archaeology have recently been addressed through the analysis of the artefact scatters in Britain, and to a lesser extent in Ireland, which have attempted to examine the ‘gaps’ between the visible monuments on a systematic basis (Barrett 1994; Bradley 1993; Green and Zvelebil 1990; Schofield 1995; Thomas 1991a, 1999a; Tolan-Smith 1996; Zvelebil et al. 1987, 1992). Clark and Schofield (1991:104) argue that in southern England, particularly in the river valleys, artefact scatters are one continuous scatter (a palimpsest effect). While artefact density varies, identifying ‘sites’ per se is problematic, as human activity “is distributed, more-or-less, continuously, across the landscape and simply varies in intensity from one location to another” (Tolan-Smith 1996:11). Thus, the tendency in archaeology to identify ‘sites’ may in fact miss regional patterns of occupation.

‘Filling in the Gaps’: Artefact scatters in Britain and Ireland Much depends on the status of flint scatters, still a major factor in the record. Lacking the context of time depth and of function as they invariably do, they remain no more than a simple record of presence/absence although attempts have been made to elevate them to the level of true settlement patterns. (Kinnes 1994b:95)

Kinnes (1994b) and others (Cooney 1997) have argued, quite correctly at this stage of systematic investigation into artefact scatters, that the information they provide is necessarily coarse. However, with the general lack of settlement structures throughout most of Britain, and particularly southern Britain, the coarse nature of artefact scatters becomes “invaluable for providing information on settlement, both as an aspect of human activity and its physical manifestation” (Schofield 2000:45). Since 1993, an English Heritage project, The Surface Lithic Scatter Sites and Stray Finds Project, has sought to gauge the scale and potential of existing records of artefact scatters (English Heritage 2000; Humble and Schofield 1996; Lisk et al. 1998; Schofield 1994a, 1995, 1997, 2000;

Much has been written regarding the archaeological examination of landscapes, the limitations of site-oriented archaeology and the practice of identifying prehistoric ‘sites’ as distinct and spatially bounded points within the landscape (Brück and Goodman 1999; Dunnell 1992; Dunnell and Dancey 1983; Foley 1981a, 1981b, 1981c; Gojda 2001; Ingold 1993, 1995; Johnston 1998a, 1998b; Knapp and Ashmore 1999; Thomas 1975; Zvelebil and Beneš 1997). Dunnell (1992:21-2) has described the concept of the ‘site’ as “defective, even deleterious to archaeology”. He states that the term ‘site’ has been 171

near Bolam Lake (Northumberland). The flint blade forms recovered would generally be identified as late ‘Mesolithic’ if collected during fieldwalking. He states that these particular blades were recovered from two pits which contained Grimston pottery, charred hazel nuts and two radiocarbon dates within the sixth millennium BP,72 all of which would suggest an early ‘Neolithic’ date. However, as we have seen in Chapter 5 and will be outlined in this chapter, movement and circulation of

Schofield and Humble 1995). Being the first of its kind, the initial results have been quite general, but the work has identified the enormity of the task at hand. The study examined four of the 50 English counties; Oxfordshire, Cornwall, West Yorkshire and Buckinghamshire (as defined prior to local government reorganisation in the mid-1990s). It recorded a total of 3300 surface scatters and 1750 stray finds. Taking this as an average, 41000 surface lithic scatters and nearly 22000 stray finds may exist across England. The question is whether this quantity of data renders artefact scatters “potentially useful but not necessarily usable” (Schofield 2000:48).

artefacts and human and animal skeletal material occurred in the ‘Mesolithic’ and the ‘Neolithic’. These could have been artefacts manufactured in the late ‘Mesolithic’, which were then incorporated into the later pits. Or alternatively, the area may have been continually visited over many centuries forming a palimpsest effect.

The project also aims to provide details to go with the numbers, including degree of survival, documentation, integrity (whether the artefact scatter is known to be discrete or not), function, chronology, size of artefact scatter, scale of collections and associations (Figures 6.27, 6.28 and 6.29) (ibid.:49). Of interest here are the preliminary findings related to the chronology. Just over 30 percent of all artefact scatters were identified as belonging to unknown periods and just over 25 percent belong to two or more periods, ‘Palaeolithic’, ‘Mesolithic’, ‘Neolithic’ and ‘Bronze Age’. There are similar figures for the number of artefact scatters identified as exclusively ‘Mesolithic’ or ‘Neolithic’, with approximately 27 percent identified as ‘Mesolithic’ and approximately 26 percent as ‘Neolithic’. In the ‘Bronze Age’ there is a significant drop to about 16 percent (ibid.:50). While there are difficulties in identifying the data of artefact scatters (see below), this suggests that similar settlement and land use strategies may have been in use in both the ‘Mesolithic’ and ‘Neolithic’ periods, which will be explored later in the chapter. Issues of chronology, taphonomy and function will now be discussed as relatively ill-defined areas of artefact scatter research; issues that must be rectified if artefact scatters are to make a significant contribution to our understanding of human social behaviour in prehistory.

The difficulty of dating artefact scatters is generally assumed to be a factor of the quality of the evidence (see Barrett et al. 1991:27-35; Whittle 1990b:104). However, R.W. Smith (1984:118) suggests that, in the Avebury region, the fault lies in the assumption that human activity dates either to the ‘Mesolithic’ or ‘Neolithic’ (see Ford 1987a; Young 1989; Young and Kay 1998). There is a long tradition of identifying distinct attributes which can assign an assemblage to the ‘Mesolithic’, the ‘Neolithic’ or the ‘Bronze Age’ (Ford 1987a, 1987b; Mace 1959; Pitts 1978a, 1978b; Pitts and Jacobi 1979; Smith 1965a:89-91). This has particularly been true in southern Britain through the study of flake and blade morphology for over forty years. Metrical analysis has been based on the premise that due to technological changes, “flakes evolved from being blade-like to short and squat between the Mesolithic and Bronze Age periods” (Schofield 1994a:91). However, it has been noted that various factors other than chronology could affect the shape and size of lithics, such as distance of occupation context from lithic sources (ibid.). The crux of the problem is that when ‘Mesolithic’ and ‘Neolithic’ populations use the same locations it makes separating material into these ‘periods’ rather difficult (Clay 2001; R.W. Smith 1984).

Chronology In lowland Britain, however, we face the paradoxical situation that we can distinguish between the technologies of the Mesolithic and Neolithic periods, but have arrived at that knowledge through little more than a desire for objective documentation. (Bradley 1987a:181)

Microliths are identified as diagnostic of the ‘Mesolithic’ in Britain, with geometric or scalene triangle microliths identified as indicative of the later ‘Mesolithic’, along with scrapers and burins. In Ireland, the late ‘Mesolithic’ 72. The two calibrated radiocarbon dates are: 5890-5810 BP (.071), 5760-5570 BP (.854) and 5540-5470 BP (.074) (4910±70 bp, Beta-117290); 5880-5830 BP (.043), 5750-5460 BP (.932) and 5370-5330 BP (.025) (4880±80 bp, Beta-117291).

Assessing the relative chronological range of artefact scatters based on lithic artefacts alone is a major difficulty affecting their integration into the wider archaeological discussion (Ford 1987a, 1991; Schofield 1991, 1994a; Whittle 1990b; Zvelebil et al. 1992). While dating may be achieved through other diagnostic artefacts, like pottery, or through the employment of radiocarbon dating, these do not necessarily date the lithic artefacts. Other artefacts, such as pottery, can also be just as vague as chronological markers (Gibson and Kinnes 1997; Thomas 1999a: Chapter 5) (Figure 6.30). For example, Clive Waddington (2000a:39) mentions results from a late ‘Mesolithic’/early ‘Neolithic’ artefact scatter 172

Figure 6.27 – Above: Bar chart of artefact scatters by period. PA = ‘Palaeolithic’; ME = ‘Mesolithic’; NE = ‘Neolithic’; BA = ‘Bronze Age’; UN = Undated. Below: Bar chart of multiperiod artefact scatters in which the numbers refer to the number of these periods represented in the multiperiod artefact scatter (after Schofield 2000: figure 5.2).

173

Figure 6.28 – Bar chart showing some of the results of the pilot study, arranged by county. Non-disc. Unkn. = nondiscrete unknown; part disc. = partially discrete: minor = minor alterations only to original scale of artefact scatter: excav. = excavated (after Schofield 2000: figure 5.3).

174

Figure 6.29 – Some results from the pilot study undertaken in 1994-5 (after English Heritage 2000: page 6). generally only constitute less than one percent of an artefact scatter (Ford 1987a:102). Thus, mixed artefact scatters can be identified “by no more than a handful of tools” (Edmonds 1997:104). However, when diagnostic artefacts are recovered, it may be difficult to identify whether or not they are associated with the scatter within which they were recovered. As Clark and Schofield acknowledge: “That artefacts were found together does not mean that they were necessarily used together” (1991:104, original emphasis). Despite all this, presumed chronological differences have been overplayed at the expense of continuity in lithic assemblages, both in form and production techniques (Bradley 1987a; Edmonds 1995, 1997; Pitts and Jacobi 1979). While microliths appear to have fallen out of use by the middle of the seventh millennium BP, core working patterns in the second half of the seventh and the sixth millennium BP continue to be concerned with the production of narrow parallel-sided blades and flakes. There is also a continued preference towards raw material selection (Edmonds 1987, 1995, 1997:104; Waddington 2000a:39). This has been taken as evidence for continued mobility into the second half of the seventh and the sixth millennium BP (Bradley 1987a; Edmonds 1987). It has also been suggested that anomalous ‘broad flake’ assemblages in the ‘Neolithic’ may reflect preliminary treatment of nodules at quarry sites (Ford 1987a). Despite this, Edmonds (1997:104) has argued “that we cannot always compare like with like”. On closer inspection, ‘Neolithic’ assemblages produced varied waste, sometimes due to the production of different artefacts (Burton 1980). There is also greater variability within what is produced through

is associated with large blade tools and core tools, such as picks and borers. In the ‘Neolithic’, flake core industries are identified, with particular forms of axes, cores, scrapers, arrowheads and other implements taken as chronologically diagnostic of the earlier and the later ‘Neolithic’ (Bradley and Holgate 1984; Clark 1932, 1934; Clark et al. 1960; Clay 2001; Edmonds et al. 1999; Finlay 2000; Gardiner 1984; Holgate 1988a:51-63, 1988b; Jacobi 1976; Kinnes 1979; Mellars 1976a; Myers 2001; Peterson 1990; Pitts and Jacobi 1979; Richards 1990:Table 6; Thomas 1999a:17; Woodman 1994, 2000; Woodman and Anderson 1990; see Table 6.1 and 6.2). Despite these assumptions, Cooney (1997:27) argues that identifying the chronological extent of diagnostic artefacts has not always been successful. He points out that leaf-shaped arrowheads have been viewed as diagnostic of the early ‘Neolithic’, however they are found in contexts throughout the entire chronological sequence of the ‘Neolithic’ (see Kinnes 1979). Similarly, using blade technology as indicative of late ‘Mesolithic’ activity is misleading since this technology survives well into the early ‘Neolithic’ in some regions of Britain (Clay 2001; Pitts and Jacobi 1979; Young 1989). In northeast England, Young (1989:167-8) has identified spatial coincidence between late ‘Mesolithic’ and early ‘Neolithic’ assemblages; such ‘mixed’ artefact scatters are identified elsewhere in Britain (Edmonds 1997:104) and pose a serious problem for the current methodological and theoretical approaches. Those lithic artefacts that may provide an indication of the chronology, if we accept the established approach, 175

1988a:Chapter 4; Newell 1981; Odell and Cowan 1987; Rick 1976; Roper 1976; Saville 1980; Schofield 1994a; Spikins et al. 1995, 2002; Stafford 1995; Waddington 1999:85-95; Zvelebil et al. 1992). The Surface Lithic Scatter Sites and Stray Finds Project has identified that 67 percent of artefact scatters identified in Buckinghamshire, Cornwall, Oxfordshire and West Yorkshire had no recognisable boundaries (English Heritage 2000:6).

core working and the knapping routines identified at this time; they were also less structured than the ‘Mesolithic’ (Edmonds 1997:104). R.W. Smith (1984:118) points out there is no reason to suggest that one could identify an artefact scatter as solely ‘Mesolithic’ or solely ‘Neolithic’ (see Ford 1987a; Young 1989; Young and Kay 1998). As outlined above, this was confirmed by the preliminary results from The Surface Lithic Scatter Sites and Stray Finds Project, in which over 25 percent of scatters were identified as belonging to more than one period. Moreover, 30 percent of scatters cannot be attributed to any period (Figure 6.27) (English Heritage 2000:6; Schofield 2000:50). This problem will continue to inhibit the incorporation of artefact scatters into further archaeological research as long we continue to impose notions of strict linear change on to archaeological assemblages and contexts.

Taphonomy: Identifying Spatial Entities

A number of these factors generate interesting relationships between the surface and subsurface assemblages, and this has been identified as an important consideration when investigating and interpreting artefact scatters (Boismier 1997; Bowden et al. 1991; Clark and Schofield 1991; English Heritage 2000; Waddington 1999:94-5; Zvelebil et al. 1992). Studies have identified the tendency for large lithics, such as cores, to be overrepresented on the surface, and for smaller lithics, such as chips, to be under-represented (Boismier 1997; Waddington 1999:94-5). Thus, there will be a bias towards larger lithics being recovered during surface collection. Waddington (1999:94) argues that this could explain why only 2 percent of the assemblage from surface collection at Milfield Basin (Northumberland) consisted of debitage and waste material. Healy (1984, 1987) has highlighted the problem of recovering earlier ‘Neolithic’ lithic assemblages when later ‘Neolithic’/earlier ‘Bronze Age’ activity has followed. Healy identified a number of contexts, including Spong Hill (Norfolk) (Figure 6.31) and Tattershall Thorpe (Lincolnshire), where surface collection identified predominantly later ‘Neolithic’/‘Bronze Age’ lithic material, while excavation uncovered earlier ‘Neolithic’ lithics. This study illustrates the discrepancies between finds from surface collection and subsequent excavation. Thus the lack of earlier ‘Neolithic’ material compared to later material, for instance, may not be a real absence, rather a function of sufficient depth of ploughsoil. Healy concludes, “In these circumstances, fieldwalking is likely to give an accurate impression of the location and intensity of Later Neolithic and Bronze Age…activity across the landscape, but is unlikely to do the same for the Earlier Neolithic” (1987:15).

Artefact scatters are affected by many taphonomic influences that make it difficult to identify them as spatial entities. Factors such as disturbances from later prehistoric and historic human social behaviour, including continual utilisation of a certain area, create a palimpsest effect. The burial and movement of assemblages by colluviation, alluviation and slope processes, as well as biological processes, such as animal disturbance (including burrowing), animal and human trampling, plant root disturbance, tree throw, erosion, plough damage and intensive modern agricultural practices, all contribute to the loss of the integrity of artefact scatters (Allen 1991; Ammerman 1985; Barton 1987; Bates 1998; Boardman 1992; Boismier 1997; Bowden et al. 1991; Clark and Schofield 1991; Cowan 1999; Dunnell and Simak 1995; English Heritage 2000; Evans and O’Connor 1999; Fritz 1984; Hinchliffe and Schadla Hall 1980; Hiscock 1990; Holgate 1985a,

Clark and Schofield (1991) have recently illustrated the elusive nature of the artefact scatters in the Ewerne Valley (Dorset). For every 330 artefacts found by excavation within the ploughsoil, only one would be identified on the total surface by fieldwalking (a recovery rate of 0.3 percent). At Haldon (Devon) and Fengate (Cambridgeshire) the amount visible on the total surface is approximately two percent (Schofield 1990:346). From Milfield Basin (Northumberland) only 3.3 percent, and from Park Farm (Wiltshire) only 3.5 percent, of the ploughsoil lithic assemblage was represented on the surface (Clark and Schofield 1991:100; Tingle 1987; Waddington 1999:94). This would indicate the possibility that surface collection would underestimate artefact scatter density or miss scatters altogether. However, this is not always the case, in the Stonehenge region, the presence of artefact scatters on the surface did not automatically indicate the presence of substantial

Lithic tools are generally dated by stratigraphic or other archaeological inferential techniques, however the possibility exists to directly date lithic material, such as the TL dating of burnt lithic material from Culverwell (Dorset) (Palmer 1999), Little Hoyle Cave (Pembrokeshire) (Anon 1991) and Avebury cuttings E, J, U (Wiltshire) (Evans et al. 1993). The absolute dating of lithics has been available for quite some time. However, these methods have yet to be employed widely in the British or Irish archaeological discourse, though they may provide a solution to the problem (Debenham 1994; English Heritage 2000; Hedges and Freeman 1994). There is also the possibility of directly dating lithic tool use by dating the carbon content of organic residues adhering to the surface of the lithic tool (Nelson et al. 1986; Saville 1986). The two examples cited from British Columbia, Canada, involved the radiocarbon dating of animal and human blood on lithic tools. However, if such organic residues are to be dated, the recovery and postexcavation analysis of lithic material must avoid handling or cleaning the pieces, meaning that much, if not all, of the current corpus of lithic material is unsuitable for radiocarbon dating.

176

177

Mainly soft hammer-struck Diffuse bulbs Abraded proximal ends

Technological characteristics of blades/flakes

Burins Geometric microliths Microburins Tranchet axes Tranchet axe-sharpening flakes

Knives Ovates Combination tools Ground-edged knives Lozenge-shaped and transverse arrowheads Ground-edge (thin butted) axes and chisels

Wide butts

Narrow butts Parallel ridges on dorsal surface left from previous removals Knives Ovates Leaf-shaped arrowheads Ground (thick-butted) axes

Hard hammer-struck Prominent bulbs Overhangs and occasional mis-hits on proximal ends

Rough flakes cores and discoidal cores Predominantly flakes

Any flint, regardless of quality, was flaked

Later ‘Neolithic’

Soft and hard hammer-struck Diffuse bulbs Abraded proximal ends

Cube-shaped blade cores Core rejuvenation flakes Predominantly blades

Good quality flint selected for flaking

Earlier ‘Neolithic’

Table 6.1 – Summary of raw material, technological and typological characteristics of later ‘Mesolithic, earlier ‘Neolithic’ and later ‘Neolithic’ flint assemblages (after Holgate 1988a: table 5.7).

Diagnostic implements

Two opposed platform cores Core tablets and crested blades Blades and bladelets

Debitage

Minimal butts Parallel ridges on dorsal surface left from previous removals

Good quality flint selected for flaking

Quality of raw material

Later ‘Mesolithic’

178 – – –

Polished axes

Polished edge knives

Other implements





Thick and pointed butted forms.

Present

Present

Leaf-shaped







‘Crested’ blades and flakes.

Narrow blades common, but not exclusive. Hard and soft hammers are used. Butts are frequently abraded.

Prepared for blades and flake removals.

Earlier ‘Neolithic’

Beaker

Present –

Bifacially flaked daggers.

Thick and a variety of thin-butted forms.

Present







Barbed-and-tanged.

Variety of invasively retouched forms.

Lozenge-shaped and a variety of transverse forms.







Attempted cores.

Broad flakes common, though not exclusive. Hard hammers are used. Butts are large and seldom abraded.

Variety of multiplatform and discoidal cores.

Later ‘Neolithic’

Table 6.2 – Chronologically diagnostic traits of late ‘Mesolithic’ to later ‘Neolithic’ flintwork in the upper Thames Valley (after Bradley and Holgate 1984: table 1).



Present

Burins

Microdenticulates

Present

Microburins



Geometric forms predominant.

Microliths

Knives

Core tablets and crested bladelets.

Other debitage



Dominance of bladelets often detached with a soft hammer. Butts are small and frequently abraded.

Removals

Arrowheads

Prepared for the removal of bladelets, single or opposed double platform cores.

Cores

Late ‘Mesolithic’

Figure 6.30 – Radiocarbon dates for Peterborough Wares, Grooved Wares and Beakers (uncalibrated dates bp) (after Thomas 1999a: figure 5.10).

Figure 6.31 – Spong Hill (Norfolk). A) Comparison of material from unstratified and stratified contexts, B) Hypothetical reconstruction of depositional history of artefacts (after Healy 1987: figure 2.1). 179

Figure 6.32 – Maddle Farm and Vale of the White Horse Fieldwalking Survey: Results of fieldwalking the same field in consecutive years (after Tingle 1987: figure 7.2).

180

isolate a discrete form of human social behaviour depending on the types of artefacts recovered, drawing on the findings of ethnoarchaeology (Binford 1980; Bradley 1995a; Gamble 1991; Holgate 1985a, 1985b). Two functions that are consistently sought for artefact scatters have been the domestic and the industrial. Schofield’s (1991a, 1991b, 1994a; also see Light et al. 1994) areaintensive surface collection survey in the upper Meon valley on the Hampshire chalklands, has proposed ways to distinguish the presumed ‘unique’ archaeological signatures of domestic and industrial zones.

additional artefacts in the subsoil (Richards 1990; Whittle 1993:42). The Maddle Farm and Vale of the White Horse Fieldwalking Survey identified that in 1984 five times more lithic artefacts were brought to the surface by ploughing than had been previously by a cultivator in the previous two years (Figure 6.32) (Tingle 1987). All this suggests that there is not one uniform relationship between surface and sub-surface deposits across Britain and Ireland. Clark and Schofield (1991:94-5) suggest that the relationship between surface/sub-surface finds will vary from region to region due to variation of environment, climate, soil types, occupational history and the nature and intensity of agricultural activity. However, generally the current research suggests that, if we are to identify density and composition of artefact scatters, there needs to be a fundamental shift towards the excavation of artefact scatters, rather than relying solely on surface collection.

According to Schofield, the upper Meon valley survey indicates that the range of artefact types is the necessary category of information (Schofield 1991a:119). He has sought to identify specific aspects of the core-reduction sequence, which is reductive and sequential. He argues that settlement and industrial/quarry contexts can be differentiated based on the two distinctive archaeological signatures that are produced (1991b:163, 1994a:91). Industrial or quarry contexts tend to be high density clusters, dominated by the initial stages of the core reduction, namely broken flakes, waste nobbles, preparation flakes and primary waste material, which produces more manageable lumps for transportation to settlement contexts (cf. Gould 1980). On the other hand, settlement contexts generally have lower densities than industrial/quarry contexts and contain tertiary waste material, with cores and retouched lithics occurring in higher numbers (1994a:91). Schofield (1990:346) does not start “with the premise ‘wider range of implements = settlement’ but rather ‘specific implement types (for example scrapers) = settlement’” (Table 6.3, Figure 6.33). However, as we have seen above, taphonomic process and the continual utilisation of a certain area (the palimpsest effect) can affect the density of artefact scatters and the lithic-types found.

Function Ethnoarchaeology…an excellent means of getting an exotic adventure holiday in a remote location…After figuring out what you think is going on with the use and discard of objects (you should never stay around long enough to master the language) you return to your desk and use these brief studies to make sweeping generalisations about what people in the past and in totally different environments must have done. (Bahn 1989b:52-3) The inability to systematically identify the function of artefact scatters has been viewed as an obstacle to their integration into studies of land use. Function was investigated as part of The Surface Lithic Scatter Sites and Stray Finds Project, with Schofield (2000:50) commenting that: Alarmingly, 93% of all scatters have no known function. So, of the 1173 scatters in one of the counties (Cornwall), only around 60 could be attributed a function, even defined in broad terms as settlement, industrial or ceremonial. Some have gone so far as to say, ‘If that is the case, what is the point?’ While the identification of function would seem to be an acceptable aim, such interpretations generally attempt to

Recently Martin Bates (1998) stated that the function of a ‘site’ could be identified by the size of artefact scatter: Artefact scatters between 1-10 metres in diameter are identified as tent/hut, butchery sites, while scatters between 10-100 metres in diameter are classified as settlements. However, such a simplistic approach fails to acknowledge the complex nature of artefact scatters and the varied and long-term nature of their formation. We cannot assume that artefact scatters can be classified as ‘domestic’, ‘industrial’, ‘ceremonial’, or as fitting any of our Western/modern categories. The reality of domestic

Density

Primary waste

Tools

Cores

Settlement

Low

Low

High

High

Industrial

High

High

Low

Low

Table 6.3 – Expected assemblage characteristics for ‘domestic’ and ‘industrial’ area assuming a policy of extra homerange production (after Schofield 1991a: table 10.1).

181

Figure 6.33 – Core reduction sequence and its archaeological correlates (after Schofield 1991b: figure 4).

182

life in small-scale societies is that the practicalities of daily life and production are often tightly associated (Cribb 1991; Kent 1984, 1990). While these neat categories, attested to in Western/modern society, help us to give order to the complex mass of artefact scatters, they do so at the expense of identifying past human social behaviour. Identifying a scatter as ‘industrial’ does not in itself add anything to our knowledge of the artefact scatter. Moreover such categories are misleading due to the tacit nature of both daily life and artefact scatters, where such activities are not necessarily performed in distinct localities across the landscape.

(1998:69, also see 2000a) notes, that while not generally stated, spatially amorphous and compositionally mixed scatters are considered ‘bad’ data because they are less amenable to functional categorisation. Therefore the focus is upon delineating patterns of landscape exploitation and adaptation that are general and repetitive, rather than variable, contingent and historically constituted. Unduly segregating artefact scatters into strict categories of function does not allow the unparalleled to be discovered. Such conceptual barriers underestimate the tacit, entangled and complex nature of daily community life and may obscure the broad study of the obvious. Further, it leads to the ‘strange’ situation whereby the majority of artefact scatters are identified as having no function (Schofield 2000:50). Of course they had some sort of function, as human activity occurred at the locations, but with deposition occurring over long periods of time, we cannot assume that an artefact scatter represents one depositional event associated with one discrete function. What the artefact scatters may represent is more a general and diverse form of human social behaviour than the specific functions that we might want to ascribe to them.

Order Out of Chaos?: The Future for Artefact Scatters Studies As so often, the development of an appropriate body of theory lagged behind the collection of data. (Bradley 1995a:14) Since the 1980s, artefact scatters have been a major component of archaeological investigation in Britain and to a lesser extent in Ireland. The enormity of the task was soon realised. Beyond the immediate realisation of the sheer mass of data to be dealt with,73 there was the issue of what artefact scatters actually represent in relation to human social behaviour. Despite 200 years of investigation into artefact scatters, it has only been in the last twenty years that issues of theory and methodology have been discussed. Cooney (1997) has recently criticised the archaeological value of the study of artefact scatters, characterising their investigation as rather problematic and unproductive. Such criticisms are premature since no archaeological theory has been set up to analyse them and no standard methodology has yet been established. A theoretical perspective is required to determine what patterns of human activity can be identified and to what extent they are represented by artefact scatters. The work in The Surface Lithic Scatter Sites and Stray Finds Project will be a fundamental aid for these aims, but it must not continue to seek to assume that human social behaviour will be separated in terms of Western/modern functional categories.

Those artefact scatters which are of interest and are unfamiliar, that is those which are “spatially amorphous and compositionally mixed” (Pollard 1998:69), are deemed problematic and placed in the ‘too hard basket’. This is a recurring theme throughout the discourse of prehistoric archaeology in Britain and Ireland. Anomalies and unfamiliar aspects of the data are either ignored or forced to conform to standard functional labels of our social reality – missing the wood for the trees. These elements of the data should be highlighted, their importance recognised as the true value of prehistoric archaeology – the possibility of identifying uncertain, unusual and unexpected forms of human social behaviour. Artefact scatters should be identified as locations where human activity occurred. Their value lies in the fact they indicate that human activity was not confined to the highly visible constructions in the landscape or to ‘villages’.

Reconstructing Lifetime Movements

We cannot attempt to pigeonhole artefact scatters’ complex story into traditional categories. Pollard

Scientific techniques that combine lead and strontium isotope analysis of archaeological human dental tissues to reconstruct the residence patterns of the individuals, have produced evidence which can identify the lifetime

73. With more material being added every year, especially since the implementation by the Department of Environment in 1990 of the Planning Policy Guidance note 16 Archaeology and Planning (PPG16).

183

the other two individuals, B and D, were probably brother and sister but unrelated to the woman, individual C, and the younger child, individual A (see Bronk Ramsey et al. 2000b; Green 2000a, 2000b; Montgomery et al. 2000; J. Richards 1999, 2000, for details of excavation and dating also see the Online Database and Appendix 6).

movements of prehistoric and historic populations in Britain and elsewhere (for example, Barreiro et al. 1997; Budd et al. 1997a, 1997b, 2001; Cox et al. 2001; Ezzo and Price 2002; Green 2000a, 2000b; Montgomery et al. 2000; Price et al. 1994a, 1994b, 1998, 2000, 2001, 2002; Sealy et al. 1995). The techniques offer an important new way of studying mobility and migration of both human and animal populations.74

Montgomery and Budd from the University of Bradford and Evans from the NERC Isotope Geosciences Laboratory, carried out the work, and have concluded that individual C did not grow up in the chalkland in which she was buried, but probably approximately 80 kilometres to the northwest in the Mendips orefield. Individual A was born before the adult woman C moved back to the chalkland. Individual B and D appear to have been born on the chalkland and moved to a different area in their early years. The procedures, methods, results and conclusions of this work will now be summarised and discussed.

Price et al. (2002) have identified two minor problems with such an investigation. First, that levels of strontium isotope ratios in local bedrock, soil, water, plant and animals are variable. Second, a range of values in human bone and enamel data makes it difficult to distinguish some migrants from locals. However, these two problems can be addressed if studies incorporate the analysis of small animal samples (preferably from tooth enamel of fossil animals) for comparative purposes whenever possible as they generally have small home ranges. This will provide a robust measure of local strontium isotope ratios and a reliable, if conservative, means for determining confidence limits for distinguishing migrants. I will now document recent work carried out in Britain. While the current sample of humans used is small, consisting of 4 individuals, the methods offer an empirical way of identifying mobility archaeologically.

Procedures and Methods Procedures and methods are discussed in Montgomery et al. (2000:370-4) are summarised below. There is a systematic variation within the nature and localities of stable and radiogenic isotopes, which become incorporated into both bone and teeth by way of diet (Faure 1986). Recently, work has supplemented this data with additional information from lead (Budd et al. 1997a, 1999). Combining lead, stable and radiogenic isotope “measurements relies on the integrity of the trace elements signal preserved within the tissue of interest and the ability to relate it to a particular period of life of the individual” (Montgomery et al. 2000:370). Strontium in archaeological skeletal tissue has been analysed as a potential indicator of residence patterns. Recent studies have highlighted the benefit of combining strontiumbased studies and lead isotopes to enhancing the study of residency and mobility (Barreiro et al. 1997; Budd et al. 1997a, 1997b, 1998, 1999). This method has been used to identify the source of modern rhinoceros horn to specific national parks in southern Africa (Hall-Martin et al. 1993).

Recently this procedure was carried out on four human skeletons, a young to middle-aged adult female (C) and three juveniles aged five, eight or nine, and nine or ten years (A, B and D, respectively), from excavations by Martin Green in 1997 at Monkton-up-Wimbourne (Down Farm Shaft) (Dorset). The four human skeletons, buried in a crouched position, came from a single, oval pit cut into the chalk bedrock and backfilled with chalk blocks and rubble. The deposition of the four skeletons was interpreted as one event, with a radiocarbon date obtained from skeleton C giving a calibrated date range of within the second half of the sixth millennium BP.75 DNA analysis carried out on the four skeletons has indicated that individual A was the daughter of individual C, and 74. Recently sulphur isotope variation has been measured to identify mobility. Measurements of į34S values complement collagen į13C and į15N measurements and can provide corroboratory palaeodietary insights or new locality information. As collagen į34S values reflect local environment į34S values, they can be used to identify the region where an individual normally resides, and therefore identify migratory individuals (Richards et al. 2001). 75. The calibrated radiocarbon date is: 5454-5374 BP (.239), 5373-5362 BP (.008), 5331-5209 BP (.366) and 5191-5048 BP (.387) (4585±50 bp, OxA-8035).

In pre-industrial societies, it is highly likely that lead was incorporated into the body mainly, or solely, from diet, and lead levels in an individual would primarily reflect those of the soil and thus the underlying geology. In studying diet, teeth are particularly advantageous as different dental tissue preserves lead intake at specific stages of life. Deciduous teeth are formed in the developing foetus within 14-19 weeks of fertilisation and complete enamel mineralisation occurs within a year of birth. There is no remodelling, so lead levels are a reliable indicator of in utero and neonatal exposure (Gulson 1996). Most permanent dentition is formed in childhood from three to four months after birth to about 12 or more years of age. The measurement of lead uptake by an individual must take into account the burial environment. However, this study examined dental enamel, which is more resistant to post-mortem lead diagenesis (Barreiro et al. 1997; Budd et al. 1997a, 1998; Montgomery et al. 1999). 184

Sample Adecid Adecid Aperm Aperm Bdecid Bdecid Bperm Bperm Cperm Cperm Ddecid Ddecid Dperm Dperm

Tooth 1 C right 1 C right 1 M right 1 M right 1 c left 1 c left 1 P left 1 P left 2 P right 2 P right 1 c left 1 c left 1 C left 1 C left

Tissue En De En De En De En De En De En De En De

Sr (ppm) 75 230 71 207 103 248 57 184 55 162 72 282 54 250

87

Sr/86Sr 0.70955 0.70788 0.70878 0.70782 0.70844 0.70790 0.70928 0.70789 0.71007 0.70866 0.70849 0.70809 0.70897 0.70792

Pb (ppm) 0.26 0.44 0.33 0.35 0.29 0.79 0.15 0.31 0.23 0.37 0.68 0.41 0.25 0.20

207

Pb/206Pb 0.8378 0.8326 0.8321 0.8288 0.8324 0.8301 0.8318 0.8287 0.8498 0.8386 0.8356 0.8269 0.8279 0.8288

208

Pb/206Pb 2.0484 2.0523 2.0501 2.0440 2.0502 2.0492 2.0499 2.0448 2.0786 2.0636 2.0532 2.0418 2.0391 2.0465

Table 6.4 – Sr and Pb isotope ratios and compositions for the tooth samples. En - enamel, De - dentine. Strontium isotope ratios were undertaken by TIMS, external reproducibility of standards was 0.0045% (2ı, n=10). Pb isotope ratios were undertaken by PIMMS, errors are 0.014% for 207Pb/206Pb and 0.024% for the 208Pb/206Pb (2ı, n=27). Isotope dilution errors also account for sample heterogeneity and are therefore estimated as 14% for Pb and 10% for Sr (after Montgomery et al. 2000: table 1).

Figure 6.34 – Strontium isotope ratios for the deciduous and permanent tooth enamel and dentine of the three juveniles (A, B and D) and permanent tooth enamel and dentine for the adult female (C). Error bar represents 2į errors calculated from 10 replicates of the NBS 987 standard (after Montgomery et al. 2000: figure 1).

185

Figure 6.35 – Lead isotope ratios for the deciduous and permanent tooth enamel (large symbols) and dentine (small symbols) of the three juveniles (A, B and D) and permanent tooth enamel and dentine for the adult female (C). Errors on the Pb isotope ratios were determined from replicate analysis of the NBS 981 standard as 0.014% for 207Pb/206Pb and 0.024% for the 208Pb/206Pb (n = 27, 2į). These are smaller than the symbols at this scale (after Montgomery et al. 2000: figure 2).

Figure 6.36 – Location map showing Monkton-up-Wimbourne (Dorset) relative to the approximate extent of the Cretaceous chalk geology and the Mendips orefield to the north west. Approximate ranges for lead and strontium isotopes for each geology are indicated (after Montgomery et al. 2000: figure 4). 186

Sample MC MS1 MS2

87

Sr/86Sr 0.70750 0.70754 0.70753

Mean

0.70752

207

Pb/206Pb 0.8275 0.8256 0.8245 0.8259

208

Pb/206Pb 2.0449 2.0414 2.0391 2.0418

Table 6.5 – Strontium and lead isotope ratios and compositions for the Monkton chalk (MC) and soil (MS) leachates. Strontium isotope ratios were undertaken by TIMS, external reproducibility of standards was 0.0045% (2ı, n=10). Pb isotope ratios were undertaken by PIMMS, errors are 0.014% for 207Pb/206Pb and 0.024% for the 208 Pb/206Pb (2ı, n=27) (after Montgomery et al. 2000: table 2).

Figure 6.37 – Strontium isotope ratios for the juveniles’ tooth enamel relative to its time of formation in year before death. Error bars associated with the permanent teeth represent the maximum period of enamel mineralisation. The adult female, C, is represented by dashed lines indicating the values of her enamel and dentine (after Montgomery et al. 2000: figure 3). those of the chalk and of the Mendip orefield. These results are seen largely as a result of diagenesis, and may indicate that she moved to a different region after her childhood before her move to the chalkland.

Results and Conclusions Results and conclusions discussed in Montgomery et al. (2000:377-81) are summarised below. Their work concludes that the radical difference between the strontium and lead isotope ratios from individual C and those of the local chalk geology show that she cannot have spent her early childhood there. The results indicate that her diet between the ages of about 2 and 6 years, after which the crown of the permanent maxillary second premolar is fully mineralised, appears to have contained strontium with a 87Sr/86Sr ratio of ~0.71 and 207Pb/206Pb and 208Pb/206Pb ratios of ~0.85 and ~2.08 respectively (Table 6.4, Figures 6.34 and 6.35). While there are a number of localities in Britain with comparable ratios, the nearest to Monkton-up-Wimbourne (Down Farm Shaft) (Dorset) is the area of the Mendip orefield about 80 kilometres to the northwest (Figure 6.36). The next nearest locality is in the north Pennines some 350 kilometres to the north. The isotopic composition of the dentine from individual C was intermediate between

Enamel from both deciduous and permanent teeth of the three juveniles revealed significant movement amongst all. The deciduous enamel of the eldest two individuals, B and D, both have lead and strontium isotope ratios similar to the chalk geology, while the two children’s permanent enamel has far more radiogenic strontium isotope ratios. This suggests a move away from chalkland geology to a region with more radiogenic strontium. The youngest individual, A, shows the opposite pattern with deciduous enamel having more radiogenic strontium than that of her permanent enamel. This reflects a move towards chalkland geology between birth and early childhood (Table 6.5). The evidence suggests that the four individuals had considerable levels of mobility during their lives. The 187

kilometres from its source, and from Cranborne Chase, some 50 kilometres from the source, illustrate the movement of lithics across the landscape, though in the case of Oakhanger V only one blade of Portland chert was identified from the assemblage of 186000 artefacts (Care 1979; Mithen 1999b:51). Pitchstone from the Isle of Arran, off the west coast of Scotland, has been found both on the Isle of Jura and in the Tweed Valley of central Scotland (Affleck et al. 1988; T. Pollard 2000). In Ireland, both shale and mudstone were moved at least 60 kilometres from their source (Sheridan et al. 1992; Woodman 1978, 1987). This movement and circulation of artefacts on a small-scale has been seen as setting the foundations of the greater ‘trade networks’ which occur in the ‘Neolithic’ (see Bradley and Edmonds 1993; Care 1982).

female (C) would appear to have moved at least 80 kilometres between her early childhood and death. Further, based on the relative times of tissue formation between the various individuals, the three juveniles (A, B and D) had a similar pattern of movement, from chalkland, to a more radiogenic region and back again to the chalkland (Figure 6.37). The authors speculate that the four may have travelled together as part of a larger group, and while they recognised that this remains purely speculative, they conclude, “the study has provided clear evidence for medium-distance movement by a small group of Neolithic people” (2000:381).

Occupation and Mobility in the ‘Mesolithic’ Much has been written on ‘residential’ mobility in the ‘Mesolithic’ and earlier periods, generally linked to the procurement of resources, subsistence and/or ecological factors (Barton 1999; Mithen 1999b; Pollard 1999a; Smith 1992; Woodman 1985b; Young 2000b). For example, mobility during the period from the tenth to the middle of the seventh millennium BP may have been due to seasonal movement between different ecological environments for the purposes of obtaining gathered and hunted food resources (Figure 6.38) (Care 1982; Carter 1998, 2001; Jacobi 1979; Mellars and Rheinhardt 1978; Mellars and Wilkinson 1980; Mithen 2000a, 2000b, 2000c; Mithen and Finlayson 1991; Richards and Mellars 1998; Schulting and Richards 2002a; Spikins 1999, 2000a; Spratt and Simmons 1976; Woodman 1978). It has also been argued that the mobile nature of the population is reflected in the flexible and portable nature of their tool kits (Bradley 1987a; Bradley and Edmonds 1993:20; Ebert 1979; Edmonds 1987; Lynch et al. 2000; Pitts and Jacobi 1979; Reynier 1998; Saville 1981, also see Cowan 1999).

Circulation and Movement of Artefacts in the ‘Neolithic’ The ‘Axe Trade’ and Lithic Movement and Circulation Over the last thirty to forty years, much research has been done on the identification of the ‘axe trade’ across Britain and Ireland (Berridge 1994; Bradley 1990, 2000a; Bradley and Edmonds 1993; Bradley and Ford 1986; Bradley et al. 1992; Briggs 1988, 1989; Chappell 1986a, 1986b; Cooney 1998, 2000a; Cooney and Mandal 1995, 1998; Coope and Garrad 1988; Craddock et al. 1983; Cummins 1979, 1980; Curran et al. 2001; Darvill 1989; Edmonds 1993b, 1995; Gardiner 1990; Jones and Williams-Thorpe 2001; Mandal 1997; Mandal and Cooney 1996; Mandal et al. 1991-2, 1997; Pitts 1996; Saville 1982, 1999; Sheridan 1986, Sheridan et al. 1992). The tendency has been to interpret this phenomenon in terms of a modern economic rationale, hence the use of terms such as ‘axe trade’, ‘axe factories’, ‘specialists’, ‘traders’, ‘entrepreneurs’ and ‘middlemen’ (Cummins 1979; Houlder 1961, 1976, also see Bradley and Edmonds 1993:43). Rather than assuming a trading of axes per se, I will refer to movement and circulation of axes and other lithic material as I have done previously for animal and human skeletal material.

Circulation and Movement of Artefacts in the ‘Mesolithic’ Possible evidence of this mobility may be documented in the movement and circulation in lithic material and other artefacts. In south Wales at Waun Fignen Femen (Powys), beach pebbles have been discovered which must have been brought from over 29 kilometres away, also Greensand chert, which was moved at least 80 kilometres from the nearest coastal sources. There are perforated mudstone beads, with the nearest source in Pembrokeshire some 100 kilometres away (Figure 6.39) (Barton et al. 1995). From southwest England, inland finds of beach flint, some as far away as Hampshire and Sussex, suggest that the lithics were being moved considerable distances. This view is supported by the similar microliths found in both coastal and inland areas (Care 1982:277; Jacobi 1979). Finds of white flint from Deepcar (South Yorkshire) originated some 80 kilometres away in the north Lincolnshire Wolds. The discovery of Portland chert at Oakhanger V (Hampshire), almost 80

Detailed examination of the movement and circulation of lithic material across Britain and Ireland, and further afield, has relied primarily on petrological analysis in order to identify petrologically distinct regions (Figures 6.40, 6.41 and 6.42, see Bradley and Edmonds 1993). As a result it has been shown that lithic materials travelled a considerable distance from their source. For example, axes found on the Isle of Man had come from both the Ulster and Cumbrian lithic sources (Coope and Garrad 1988). Cumbrian axes (Group VI) extend over much of Britain from the Highlands of Scotland to the southern coast of England and into Ireland. Cornwall axes (Group I) have been found throughout England and into Wales (Figure 6.43). Graig Lwyd axes from north Wales (Group VII) and axes from southwest Wales (Group VIII) have been found throughout Wales and into eastern England (Figure 6.44) (Bradley and Edmonds 1993). Axes also 188

Figure 6.38 – The inferred seasons of use of the Oronsay middens (Argyll and Bute), based on faunal evidence (after Richards and Mellars 1998: figure 2).

Figure 6.39 – Distribution of main raw material sources in relation to Waun Fignen Felen (Powys) (after Barton et al. 1995: figure 6).

189

Figure 6.40 – Principal sources of non-flint stone identified through petrological analysis. The hatched area represents the main outcrops of flint and the triangles mark the position of ‘Neolithic’ flint mines (after Edmonds 1995: illustration 5).

190

Figure 6.41 – Distribution map of Group VI axes in Ireland identified by thin section (open circles indicate county or area only provenance) (after Cooney and Mandal 1995: figure 4).

191

Figure 6.42 – Distribution of porcellanite axes in Ireland (after Cooney 2000a: figure 6.14).

192

Figure 6.43 – Distribution of stone axes from Cornish (petrological group I) and Cumbrian (petrological group VI) sources (after Edmonds 1993b: figure 7.2).

Figure 6.44 – Distribution of stone axes from sources in north (petrological group VII) and south (petrological group VIII) Wales (after Edmonds 1993b: figure 7.3). 193

Figure 6.45 – Location of porcellanite axes in Britain and tuff axes in Ireland (after Cooney 2000a: figure 6.16).

194

other suitable stone sources (Figure 6.45) (Edmonds 1995:50; Sheridan 1986:19; Sheridan et al. 1992:410). This work has also identified stone axes in Ireland from sources in southwest Wales and Cornwall (Cooney and Mandal 1995; Mandal 1997).

came from continental Europe, with a few Seledin axes from Brittany found in Britain (Le Roux 1979). Also, a thick-butted Scandinavian axe has been discovered at Julieberrie’s Grave (Kent) (Jessup 1939) and hundreds of jadeite and nephrite axes, probably from the Alpine region of continental Europe, have been found in Britain and possibly in Ireland (Bradley 2000a:120; Ricq-de Bouard 1993; Woolley et al. 1979). Examples in Britain come from Hambledon Hill Complex (Dorset), Sweet Track (Somerset), Cairnholy (Dumfries and Galloway) and High Peak (Devon) (Coles et al. 1974, Mercer 1980:23; Thorpe 1996:179). These jadeite axes are common in continental Europe, along with the production flakes. Such flakes are almost unknown in Britain and Ireland, suggesting that the jadeite axes arrived as finished objects (Edmonds 1995:50-1).

While many books and papers have been devoted to the circulation and movement of lithic material based on petrological analysis, the methods involved are not without their problems. Petrological groups have not always been thoroughly and rigorously defined. A number of methodological and interpretive problems call into question the reliability of some petrological identification (in particularly see Berridge 1994; Briggs 1976a, 1976b, 1977, 1979, 1989). Berridge (1994) notes that by the late 1980s it was suggested that Cornwall sourced 15 percent of all grouped implements and at least as many ungrouped items from the British Isles. However, the generalised rock type identified as coming from Cornwall is far from exclusive to this region, and is also found in Devon, southwest Wales, north Wales, northern England and Scotland. Secondary deposition via glacial erratics may have distributed certain rock types across Britain and Ireland by natural processes. This observation in Cornwall illustrates the complex factors at work and the problems of standard interpretation in which “the petrologists are sound and beyond dispute” (ibid.:47).

Portland chert, chalk flint nodules and pitchstone from the British Tertiary Volcanic Province, not necessarily as axes, have been found throughout Britain (Healy 1998b; Thomas 1984; Thorpe and Thorpe 1984). Chalk flint nodules have been found in Cornwall and the Walton Basin, some 100 kilometres and 125 kilometres respectively from their likely source (Healy 1998b:26). Pitchstone, a glassy volcanic rock found at several centres within the British Tertiary Volcanic Province (BTVP) in the west of Scotland and Ireland, has been located up to 300 kilometres from possible sources, indicating considerable movement.

Further problems with identifying the geographical source of lithic material have recently been identified in south Wales by Jones and Williams-Thorpe (2001). The chemical composition of 11 British prehistoric implements previously identified as from likely sources in south Wales (Group VIII from Carnalw, Group XIII from Carnmenyn) were re-analysed. Atypicality indices, together with mineralogy, indicate that only one out of five identified as Group VIII are likely to be Group VIII and only two out of six supposed Group XIII artefacts are likely to be Group XIII. While illustrating the usefulness of atypicality and mineralogy indices, this indicates the variability that can exist within petrologically defined groups and further questions the petrological techniques and standard interpretations, the basis for identifying the ‘axe trade’ and the movement and circulation of other lithic material. This work illustrates that our current knowledge and understanding of the extent and nature of the ‘axe trade’ may be invalid.

Healey and Robertson-Mackey’s (1983) survey of lithic material from ‘Neolithic’ contexts identified a number that have lithic raw material from over 32 kilometres away, including Staines (Surrey), Hurst Fen (Suffolk), Orsett (Essex), Bishopstone, Windmill Hill (Wiltshire), Pamphill (Dorset), High Peak, Hembury (Devon) and Carn Brea (Cornwall). Over 80 Old Red Sandstone querns and rubber fragments found at Hambledon Hill Complex would have come from the Mendips, some 40 kilometres to the northwest (Healy 1998b:25). Saville (1982) has estimated that 1175 kilograms, approximately 1000 large nodules or cores of chalk flint, were taken into the Cotswolds every year during the ‘Neolithic’, some 2350 tonnes over 2000 years. Over the last ten to fifteen years extensive work has been carried out to identify the source of lithic material in Ireland, utilising both petrological and geochemical techniques (Briggs 1988; Cooney 1998, 2000a; Cooney and Mandal 1995, 1998; Curran et al. 2001; Mandal 1997; Mandal and Cooney 1996; Mandal et al. 1991-2, 1997; Sheridan 1986; Sheridan et al. 1992). A variety of rock types was utilised in Ireland, such as mudstone, gabbro and ‘porphyry’, but the most common was porcellanite, accounting for about 70 percent of the total of 20000 known axes (Cooney and Mandal 1998). The porcellanite sources found at Brockley and Tievebulliagh (Co. Antrim) (Group IX) (Figure 6.42) were predominant, with well over half of all axes across Ireland being manufactured from these two sources (Cooney 2000a:200). A total of 180 porcellanite stone axes have been identified across Britain from northern Scotland and southern England, sometimes quite close to

One of the most widely discussed and contentious cases for the movement of lithics in the ‘Neolithic’ is the ‘transportation’ of the large Bluestones for the construction of Stonehenge (Wiltshire) (Cleal et al. 1995; Green 1997; Howard 1982; Scourse 1997; Thomas 1923; Thorpe et al. 1991; Williams-Thorpe et al. 1997). ‘Bluestone’ covers a variety of geologically similar igneous rocks that have sources in the Preseli Mountains (Pembrokeshire), south Wales. This stone has been found on and near the Salisbury Plain in a variety of sizes, from large slabs to small fragments. It is also used for constructions throughout south Wales, up to 40 kilometres from the source (Thorpe et al. 1991). Generally, the argument has been debated by those who 195

assemblage at Cherhill (Wiltshire) has identified a closer source of fossil shell. The fossil shell fragments and ooliths inclusions found within the pottery assemblage from Cherhill could have been derived from the Corallian limestone, about 9.5 kilometres west and northwest from Cherhill (Figure 6.46). This source is 12 kilometres west and northwest from Windmill Hill, providing a closer source than the Bath/Frome area suggested by Smith in the 1960s (Zienkiewicz 1996:93-4; Zienkiewicz and Hamilton 1999:269).

argue that ice moved the bluestones to the Salisbury Plain prior to the ‘Neolithic’ (see Thorpe et al. 1991; WilliamsThorpe et al. 1997) and those who argue that humans were the agents that moved the bluestones in the ‘Neolithic’ (see Green 1997; Scourse 1997; WilliamsThorpe et al. 1997). Current evidence relating to the extent and movement of Quaternary glaciers suggests that the glacial hypothesis cannot explain the bluestone’s occurrence on the Salisbury Plan. Even at their greater extent in Britain, the glaciers of the Quaternary Ice Age never reached the Salisbury Plain (William-Thorpe et al. 1997:318). As Scourse (1997:308) concludes: The hypothesis that the Stonehenge bluestones were transported by ice to the Wiltshire Downs in not consistent with current glaciological or glacial geological theory or observational data. Thus, the current evidence suggests that movement of the bluestone from the Preseli Mountains to the Salisbury Plain was, at least in part, due to human action.

Healy (1998b:25) has recently discussed the sources of clays used to produce pottery excavated from Maiden Castle and Hambledon Hill Complex (Dorset). At Maiden Castle some pottery contained clays from Jurassic sources, which probably come from 14 kilometres to the west. From the Hambledon Hill Complex more than 24000 pottery sherds have been recovered. Analysis by Darvill indicated that none of them was produced from clays on the actual hill, rather most came from the surrounding 2-3 kilometres. However, significant quantities came from further afield. About 16 percent of the assemblage (over 4000 sherds) came from a number of Jurassic sources lying in a belt to the west, northwest and southwest, at least 20 kilometres away. One percent of the pottery assemblage contained fresh shell temper, which probably came from the coast, some 30 kilometres to the south.

Similar practices have been noted in the construction of other structures. Frank Mitchell (1992) has identified the movement of small cobble stones for incorporation into the chambered structures at Newgrange and Knowth (Co. Meath), some of which were moved some 60 kilometres. The fine dry-stone walling between the stone uprights at West Kennet (Wiltshire) are composed of oolitic limestone, which comes from 30 kilometres to the southwest. Four or five of the large stones used in the construction of the stone alignment at Devil’s Arrow (North Yorkshire) seem to have been moved 15 kilometres (Meighan and Simpson 2001:347).

Interpreting the movement and circulation of artefacts and raw materials Identifying that shell moved 30 kilometres and axes several hundred kilometres, may only have implications for the mobility patterns for a certain section of the population. The interpretation of what this actually represents is complex. Are we seeing an extensive trading network, hand-to-hand exchange, mobile residential occupation, direct or indirect acquisition or exchange, or a combination of these (Darvill 1989:41; Edmonds 1995:53; Whittle 1999a:65)? All of these are feasible explanations. What I want to concentrate on is what this movement and circulation reveals about the movement of people and material across the landscape.

Pottery, Clay and Shell Since the 1960s research has been carried out which suggests that some pottery and/or the raw material for production have been moved and circulated over considerable distances (Bradley 1982; Cornwall and Hodges 1964; Ellison 1981; Hodges 1962; Howard 1981; Parker Pearson 1990; Peacock 1969). For example, pottery with Gabbroic inclusion from Cornwall has been excavated throughout southwest England and into Dorset and Wiltshire (Edmonds 1995:55; Thomas 1999a:102; Thorpe 1996:176). Much discussion has centred on the inclusion of shell in Grooved Ware pottery, the predominant temper used in such pottery (Thomas 1999a:121). Throughout Wessex and beyond, the inclusion of shell occurs even when the local region lacks marine shell, shell-bearing clays or fossil shell (Cleal et al. 1994:445). Analysis of Grooved Ware from the Chalk Plaque Pit, Woodlands and Ratfyn (Wiltshire), identified that shell inclusion came from marine oysters and not fossilised shell. This demonstrates that the shell must have come from the coast, the closest area being some 50 kilometres away (ibid.:446-7).

The evidence outlined above illustrates that artefacts were moving and being circulated, by whatever human means, across the landscape. As demonstrated some of these artefacts were coming from considerable distances away. I would argue that the patterning of human and animal skeletal material identified in Chapter 5 of certain skeletal parts being found across the landscape could be viewed in a similar way. While the population were moving and circulating artefacts they were also carrying skeletal material and incorporating this material into deposited assemblages. Thus, I would argue that the assemblages of artefacts and skeletal material which have been documented throughout this book and the Online Database are the ‘materiality’ – the pattern on the ground – produced by a mobile population moving across the landscape.

The Avebury region was one of the first areas where detailed pottery studies were carried out. Smith suggested that the pottery assemblage from Windmill Hill (Wiltshire) included fossil shell from the Bath/Frome area, some 30 kilometres to the northwest (1965a:546). However, recent work by Darvill (1983:98) on the pottery 196

Figure 6.46 – Simplified geological map of the area around Cherhill (Wiltshire), showing clay outcrops and the approximate locations of sample points. A: ST 994 791; B: SU 012 771; C: SU 029 748; D: SU 024 744; E: SU 018 684 (after Darvill 1983: figure 27). the ‘Neolithic’. Thus the residential activity may have encompassed mobility, short-term and perhaps long-term sedentism. In Chapter 2, evidence for broad-based subsistence strategies was identified which showed considerable continuity with the preceding ‘Mesolithic’, which would have required similar movement across the landscape. In relation to ‘residential’ mobility much weight has been given to evidence from artefact scatters and broad-based subsistence strategies, particularly in relation to a populations’ seasonal movements across the landscape (see Chapter 2). Currently, artefact scatters are viewed as active and significant components of a dispersed and mobile system of land use that is increasingly being argued for the ‘Neolithic’ in Britain. Thus, arguments that are currently offered for these scatters, view them in a similar light as those in ‘Mesolithic’ contexts – an archaeological signature of a mobile ‘Neolithic’ population utilising various locations at different times during the annual economic

Conclusions: Identifying ‘Residential’ Mobility One of the problems in the treatment of the British evidence has been the implicit assumption that the pattern of mobility seen in the interpretation of the Neolithic evidence in Wessex…can be extended across the whole of the island and indeed across the sea to Ireland (Cooney 2000a:38) There has been a long tradition of identifying ‘Neolithic’ populations as synonymous with a mixed farming economy, the establishment of permanent settlements and a sedentary lifestyle, but as we have seen this view has changed (see Chapter 2). This chapter has identified evidence for the movement and circulation of artefacts, artefact scatters and lead and strontium isotope analysis, which may indicate ‘residential’ mobility in 197

al. 1986) and Ireland (Cooney 1999a, 2000a). However, the presence of these structures does not rule out that the society, to whatever degree, was mobile. Other evidence identified across different parts of Britain and Ireland suggests that mobility was part of the residential pattern of the population.

and residential cycle. Montgomery et al. (2000) place their analyses within the current discussion on the nature of occupation and ‘residential’ mobility in ‘Neolithic’ Britain. From the general absence of remains of permanent settlement throughout most of Britain, combined with the presence of artefact scatters and broad-based subsistence strategies, scholars have suggested that during the ‘Neolithic’ there was strong continuity with the preceding ‘Mesolithic’ in terms of ‘residential’ mobility. While the sample used by Montgomery et al. (2000) was small, it offers us an insight into the movements of a small group of people, of special interest due to the group comprising three juveniles, and has the potential through a wider application to make a fundamental contribution of the lifetime movements of populations in the ‘Neolithic’ and ‘Mesolithic’.

Finally, the evidence presented in Chapter 5 and this chapter can be viewed in the same light. The movement and circulation of stone, clay and shell across the landscape can be seen as analogous to the movement of human and animal skeletal material by the living population. This evidence illustrates that artefacts were moving and being circulated, by whatever human means, across the landscape. I would argue that skeletal material also could have moved in similar ways, that individual items were being moved and circulated across the landscape by humans. The current discourse has sought complex explanations for the movement of human skeletal material across the landscape. The phenomenon, in the broader sense, is not complex but the patterning of deposition which would be produced by a population that is mobile using a variety of residential options. This provides a parsimonious argument for both the ‘Mesolithic’ and ‘Neolithic’ with one tradition of ‘action’, whatever specific verbalised meanings may have been involved.

The individual lines of evidence for mobility presented in this research on their own would not be significant enough to identify ‘residential’ mobility in the ‘Neolithic’. However, conclusions can be drawn by combining the evidence for a dispersed mode of land use in the ‘Neolithic’ as identified by artefacts and broadbased subsistence strategies, with evidence of long-range artefact movement and circulation with the new work on lead and strontium isotopes. The minimal and most parsimonious hypothesis is that individuals were moving across the landscape of Britain during the second half of the seventh to the late fifth millennium BP. There is some logical risk in attempting one holistic, all-encompassing model for the entirety of Britain and Ireland. It is better to think in terms of degrees of mobility and sedentism. Substantial structures, which have been identified as locations of occupation, have been found in northern and western Scotland (Armit 1996; Clarke and Sharples 1985; Richards 1992b; Ritchie 1985a; Whittle 1985a; Whittle et

The next chapter will case study the Avebury region in Wiltshire drawing on both primary and secondary sources relating to archaeological and environmental data gathered over the last few centuries to attempt to offer a new interpretation. The mobility argument will examine the empirical evidence without resorting to the use of standard social theory; continuity in human social behaviour across the ‘Mesolithic’/‘Neolithic’ divide will be identified from the early tenth to the late fifth millennium BP.

198

PART III CHAPTER SEVEN: THE AVEBURY REGION The archaeological discourse in Britain and Ireland is dominated by attempts to identify human social behaviour on a familiar social and individual action level. Attempts are made to identify explicit human thoughts and classes of verbal meaning, especially related to mortuary ritual. But archaeological assemblages in themselves are silent and alone cannot reveal such details about verbal meaning and significances. In order to give a voice to archaeological assemblages, supplementary propositions are employed which give verbal meaning to the material.

A minor fact is worth more than a great fiction. (Malmer 1997:7) Chapters 5 and 6 identified the movement and the dispersed patterning of artefacts, animal and human skeletal material across the landscape. While artefact scatters and the subsistence base of the population have been viewed as possible evidence of a dispersed and mobile pattern of land use, the location of the skeletal material across the landscape has not been considered in the light of this proposed mobile residential pattern. There is a need for a simple interpretation of the dispersed landscape of Britain and Ireland in the ‘Neolithic’ to be formulated – one that combines current arguments relating to mobility and land use and the dispersed and widely distributed patterning of artefacts and skeletal material across the landscape.

The current pitfalls of the interpretive framework of the ‘Neolithic’ discourse using standard social theory were outlined in Chapters 1 and 3. With this in mind this case study will attempt to look at the evidence for human social behaviour in a new light. This case study is not presented as a model of practices across the whole of Britain and Ireland, but is used to illustrate an approach, where unparalleled aspects of human social behaviour are allowed to be expressed and continuity over the millennia is identified.

The empirical evidence relating to the dispersal of human skeletal material across the landscape can be understood at a gross level as the ‘fall-out’ from a dispersed, mobile occupation system. Specific, complex verbal meaning systems of elaborate mortuary practices and rituals will, of course, have been associated with these remains. However, with deposition occurring over several hundred, perhaps several thousand years, we cannot assume that the deposition events are the result of the same particular social practice in each region and at each time. While other approaches may have their own specific value, the mobility argument will stress that it is important to recognise that a pattern with considerable social behavioural significance can be identified at larger hierarchical scales of explanation (King 2001). This book has viewed the empirical evidence using a larger spatial and temporal perspective which allows the recognition of continuity in human social behaviour that cuts across the convention of the ‘Mesolithic’/‘Neolithic’ divide. The core problem is that at the general level of explanation taken in ‘Neolithic’ studies, elaborate mortuary practices and rituals are identified, which have exaggerated the ‘Mesolithic’/‘Neolithic’ divide. Hence, the movement and circulation of axes has generally been interpreted in economic terms as an ‘axe-trade’, while dispersed human skeletal material has generally been viewed as evidence of specific mortuary ritual. However, I would argue that there could have been a similar reason of ‘economy’ for the dispersal of human skeletal material across the landscape (see Thomas 2000a:660).

The Avebury Region: A case study The chalk downland of north Wiltshire contains the famous monuments of Windmill Hill, Avebury, West Kennet and Silbury Hill. The Avebury region occupies the Upper Kennet Valley, and is defined as an area roughly 7-12 kilometres across around the Avebury henge, approximately 100 square kilometres (Figures 7.1 and 7.2). The Avebury region has been a major area of antiquarian, archaeological and palaeoenvironmental research since John Aubrey’s work in the seventeenth century, William Stukeley in the eighteenth century and Sir Richard Colt Hoare and John Thurnam in the nineteenth century. Many earthen, stone and timber constructions dominate the region, along with artefact scatters, which will be discussed in this case study. The constructions and the deposition of artefacts in, around and below them will be discussed in relation to the changing environment and woodland cover from the early tenth to the late fifth millennium BP (Table 7.1, Figures 7.3, 7.4, 7.5, 7.6, 7.7 and 7.8, this table and these figures will be the basis for the case study). All the radiocarbon dates (and the samples dated) from the case study region have been represented graphically in Appendix 4.

199

200

Figure 7.1 – General location map of the Avebury region (Wiltshire) (after Whittle et al. 1999a: figure 7).

Figure 7.2 – ‘Neolithic’ features of the Avebury region (Wiltshire), referred to in figures 7.3, 7.4 and 7.5 (after R.W. Smith 1984: figure 1). AV - Avebury: Henge; BR - Beckhampton Road: Earthen Long Mound; G.55 Avebury G.55: Round Mound; G.6a - West Overton G.6a: Round Mound; HA - Hackpen: Occupation Debris; HK - Hemp Knoll: Round Mound; HP - Horslip: Earthen Long Mound; SA - Sanctuary: Timber Circle; SB - Silbury: Mound; SS - South Street: Earthen Long Mound; WA - Waden Hill: Occupation Debris; WK - West Kennet: Chambered Mound; WKA - West Kennet Avenue: Occupation Debris beneath stone avenue; WH - Windmill Hill: Causewayed Enclosure.

201

202

3950-3650

F

4514/4409-4083/3909

4866/4732-4514/4409

5311/5059-4866/4732

5646/5492-5311/5059

5949/5776-5646/5492

6304/6195-5949/5776

Calibrated BP ranges at 1 ı

Trend to open country, though timber still available. Cultivation. Pigs

Renewed clearances

Trend to more scrub or woodland again

More clearances, in mosaic pattern. Some plough cultivation and cereals. Animal husbandry; some herding beyond area ?

Scattered small clearances in woodland. Animal husbandry

Woodland. First clearances ?

Environment and subsistence

Development of Avebury. The Sanctuary. Silbury Hill at phase E/F border Palisade enclosures. Beaker burials

Small contexts and pit groups. Larger lithic scatters

More permanent or marked occupation areas ?

? Simple circles ? Start of West Kennet Avenue

More elaborate and larger mounds. Enclosures towards the end of phase C

Dispersed, with ? local nucleation. Small contexts, pit groups, lithic scatters

Uncertain. Density as in phase A and B ?

Simple mounds. Constructions ?

?

Constructions

Dispersed. Small pit groups

?

Occupation

Table 7.1 – Outline summary proposed by Whittle of phases in the ‘Neolithic’ of the Avebury region. Information from this will be referred to throughout the chapter (after Whittle 1994a: table 7).

4250-3950

4850-4550

C

E

5150-4850

B

4550-4250

5450-5150

A

D

BP uncalibrated range

Phase

203

Figure 7.3 – Schematic reconstruction of the early ‘Neolithic’ landscape of the Avebury region (Wiltshire) (after R.W. Smith 1984: figure 8).

204

Figure 7.4 – Schematic reconstruction of the middle ‘Neolithic’ landscape of the Avebury region (Wiltshire) (after R.W. Smith 1984: figure 9).

205

Figure 7.5 – Schematic reconstruction of the late ‘Neolithic’ landscape of the Avebury region (Wiltshire) (after R.W. Smith 1984: figure 10).

206

Figure 7.6 – Avebury region (Wiltshire): Earlier ‘Neolithic’ (after Thomas 1999a: figure 9.1).

207

Figure 7.7 – Avebury region (Wiltshire): Middle/later ‘Neolithic’ (after Thomas 1999a: figure 9.4).

208

Figure 7.8 – Avebury region (Wiltshire): Later ‘Neolithic’/Beaker (after Thomas 1999a: figure 9.7).

totality with which secondary woodland invaded the infilling ditches indicates for the first time and unequivocally the presence of woodland on the high chalklands in the middle Holocene. A similar pattern has been identified at Horslip, based solely on evidence from molluscan analysis of buried soils (Whittle 1978:37). The reliability of such evidence has recently been questioned by Carter (1990) and in earlier work by Edwards (1979). The implications of their work is that due to problems with temporal and ecological resolution, patterns identified from molluscan assemblages can be exaggerated and misleading.

Environment It is simply impossible to know exactly where the trees and bushes were in relation to sites and monuments…Even the greatest refinements in techniques such as pollen and molluscan analysis can provide only a relatively coarse assessment of these matters…they tell us all too little of what the forests and open areas actually looked like. (Tilley 1994:73) The environmental history of the Avebury region has been intensively studied over many years (Ashbee et al. 1979; Cartwright 1999; Davies 1999; Davies and Wolski 2001; Evans 1966, 1968, 1972, 1990, 1991, 1993; Evans and Smith 1983; Evans et al. 1978, 1985, 1988a, 1993; Fairbairn 1999; Fishpool 1992, 1999; Limbrey 1992; Mount 1991; Powell et al. 1996; Robertson-Mackey 1980; Rouse and Rowland 1999; Smith 1965a; Walker 1999; Whittle 1993, 1994a, 1997b; Whittle et al. 1993). Evidence for the past environment has come from molluscan assemblages, buried soils, faunal assemblages, ditch sediments, charcoal and carbonised plant remains, generally from ‘site’ contexts but also from a few non‘site’ contexts.

Sixth millennium BP deforestation in the Avebury region was small-scale and occurred both on the valley floor and also on the slopes and the plateau. The trajectory of change in the woodland component of the landscape was not constant or continual, with evidence suggesting that this component was in constant flux. Whether the pattern was due to human deforestation or a natural phenomenon, or a mixture of these, is not the concern here (see Brown 1997). My concern is the spatial visibility across the landscape and accessibility within the landscape, both of which were affected by the extent of the woodland vegetation cover. This issue, along with inter-visibility, has been dealt with by a number of publications (Devereux 1991; Evans 1985:84; Fraser 1983, 1988; Gibson 2002a:9-10; Phillips 2000, 2002:58-9; Thomas 1993c; Tilley 1993, 1994; Vyner 2001; Wheatley 1995), but more work is needed to view and interpret the significance of the constructions in this landscape over millennia (King 2001).76

Some time between 14000 and 10000 BP, non-humic marl was laid down across the Avebury region, with molluscan evidence suggesting open country at this time. In response to climate warming at the Devensian lateglacial/Holocene boundary, a stable land surface developed at West Overton. Between c.10300 and 8500/8300 BP the environment was probably open marsh, then the open country developed into woodland. This woodland, as shown by molluscan faunas, was terrestrial or marshy but not swampy, and is dated by mollusca to after c.9000-8500 BP. The Holocene woodland is attested to in the river valley floors and from slopes and dry valleys at South Street, Avebury, Cherhill, Beckhampton Road and Easton Down (Evans 1972; Evans et al. 1985a, 1993:185; Whittle et al. 1993). The second half of the seventh millennium BP witnessed the disappearance of primary woodland at West Overton, Cherhill, Easton Down, Avebury, Windmill Hill, South Street and Beckhampton Road, and some woodland decline may date to the early seventh millennium BP (R.W. Smith 1984). The disappearance of the woodland should not be viewed as the result of one episode, but rather the result of probably small-scale (local) and sporadic incidents of woodland manipulation (Evans et al. 1993:187; Whittle et al. 1993).

The recent examination of assemblages from Windmill Hill has produced evidence for the environmental setting of the enclosure’s construction and use (Fishpool 1992, 1999). Radiocarbon dating specifies that the enclosure was used throughout the sixth millennium BP, especially the mid-sixth millennium BP (Ambers and Housley 1999). Land mollusca analysis indicates the natural vegetation was probably woodland (Fishpool 1999:131). This pattern is seen also at Avebury and Easton Down, illustrating that woodland existed in the seventh millennium BP (Evans et al. 1985; Whittle et al. 1993). Prior to the construction of the bank in the early sixth millennium BP, the environment consisted of grassland 76. D. Wheatley’s (1995) cumulative viewshed analysis (CVA) employed a GIS method for investigating intervisibility between related locations within a landscape in relation to the topology. The method has been employed in relation to the Avebury and Salisbury Plain areas. Wheatley highlights the problem of such an analysis within a wooded landscape.

Following this, there was a trend towards increased scrub or woodland in the late sixth millennium BP at Knap Hill, Cherhill, South Street, Dean Bottom, Roughridge Hill, West Kennet, Millbarrow, Windmill Hill and Easton Down (Connah 1965; Evans and Smith 1983:109; Whittle 1994a, 1997b:140; Whittle 1993:232). In relation to the Easton Down long mound, Whittle et al. (1993:228) state: The combined evidence of woodland from the pre-barrow subsoil hollow and the speed and 209

Easton Down Millbarrow South Street Beckhampton Road

Construction date (calibrated BP)

Minimum time to wood regeneration (years)

Maximum time to wood regeneration (years)

5311-5051 5322-4887 5650-5050 5294-4879

0 11 339 0

579 590 918 562

Table 7.2 – Dates of long mound construction and theoretical minimum and maximum time until woodland regeneration. All dates in calibrated BP and at the one standard deviation level from Whittle et al. 1993. Relevant laboratory codes are: Easton Down (OxA-3762), Millbarrow (BM-2730 and BM-2729), South Street (BM-356 and BM-358a) and Beckhampton Road (BM-506b) (after Davies and Wolski 2001: table 1). Easton Down A. pura P. pygmaeum C. tridentatum A. aculeate

max 2779 1853 926 < 463

min – – – –

Millbarrow max 2832 1889 944 < 472

min 35 18 9