MATERIAL EVIDENCE

How do archaeologists make effective use of physical traces and material culture as repositories of evidence?

Material Evidence: Learning from archaeological practice takes a resolutely case-based approach to this question, exploring instances of exemplary practice, key challenges, instructive failures, and innovative developments in the use of archaeological data as evidence. The goal is to bring to the surface the wisdom of practice, teasing out norms of archaeological reasoning from evidence. Archaeologists make compelling use of an enormously diverse range of material evidence, from garbage dumps to monuments, from finely crafted artifacts rich with cultural significance to the detritus of everyday life and the inadvertent trans- formation of landscapes over the long term. The contributors to Material Evidence identify particular types of evidence with which they grapple and consider, with reference to concrete examples, how archaeologists construct evidential claims, critically assess them, and bring them to bear on pivotal questions about the cultural past. Historians, cultural anthropologists, philosophers, and science studies scholars are increasingly interested in working with material things as objects of inquiry and as evidence – and they acknowledge on all sides just how challenging this is. One of the central messages of the book is that close analysis of archaeological best practice can yield constructive guidelines for practice that have much to offer archaeologists and those in related fields. Bob Chapman is Emeritus Professor of at the University of Reading, UK. His research focuses on , Mediterranean later prehistory, the development of human inequality, and the means by which this can be studied with archaeological data. He has pursued these interests in fieldwork projects in southeast Spain and the Balearic Islands, as well as in books such as The Archaeology of Death (1981), Emerging Complexity (1990) and Archaeologies of Complexity (2003). In recent years his research has turned increasingly to the use of historical materialism in archaeological interpretation, especially in relation to inequality and human exploitation. Running through this research activity has been a strong concern for the nature of archaeological interpretation, working with the complementary evidence of how people lived (e.g. what they produced, exchanged, and consumed, centred on settlement evidence) and how they were treated in death (e.g. their disposal, centred on burial evidence).

Alison Wylie is Professor of Philosophy and Anthropology at the University of Washington, and Professor of Philosophy at Durham University. She is a philosopher of the social and historical sciences who works on questions about objectivity, evidence, and research ethics raised by archaeological practice and by feminist research in the social sciences. Her longstanding interest in evidential reasoning is represented by her book Thinking from Things (2002) and by her contributions to Evidence, Inference and Enquiry (ed. Dawid, Twining and Vasilaki, 2011), How Well do ‘Facts’ Travel? (ed. Morgan 2010), and Agnatology (ed. Proctor and Schiebinger 2008). In recent work she focuses on the role of contextual values in science and on how research can be improved by internal diversity and by collaborations that extend beyond the research community. These interests are reflected in Value-free Science? (co-edited with Kincaid and Dupré 2007) and Epistemic Diversity and Dissent (edited for Episteme 2006), as well as in essays on stewardship and feminist standpoint theory. MATERIAL EVIDENCE

Learning from archaeological practice

Edited by Robert Chapman and Alison Wylie First published 2015 by Routledge 2 Park Square, Milton Park, Abingdon, Oxon OX14 4RN and by Routledge 711 Third Avenue, New York, NY 10017 Routledge is an imprint of the Taylor & Francis Group, an informa business © 2015 Robert Chapman and Alison Wylie for selection and editorial matter; individual contributions, the contributors. The right of Robert Chapman and Alison Wylie to be identified as the author of the editorial material, and of the authors for their individual chapters, has been asserted in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988. All rights reserved. No part of this book may be reprinted or reproduced or utilised in any form or by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying and recording, or in any information storage or retrieval system, without permission in writing from the publishers. Cover image: Rock and rock art – an example of material evidence in context, photographed by Andrew Cochrane and Aaron Watson (copyright Andrew Jones). Trademark notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identification and explanation without intent to infringe. British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library Library of Congress Cataloging in Publication Data Material evidence : learning from archaeological practice/edited by Robert Chapman and Alison Wylie. pages cm Includes index. 1. Archaeology—Philosophy. 2. Archaeology—Methodology. 3. Material culture. I. Chapman, Robert, 1949– editor. II. Wylie, Alison, editor. CC72.M39 2014 930.1028—dc23 2014029219

ISBN: 978-0-415-83745-3 (hbk) ISBN: 978-0-415-83746-0 (pbk) ISBN: 978-1-315-73927-4 (ebk)

Typeset in Bembo by Keystroke, Station Road, Codsall, Wolverhampton CONTENTS

List of figures ix List of maps xiii List of tables xv List of contributors xvii

1 Material evidence: learning from archaeological practice1 Alison Wylie and Robert Chapman

PART I Fieldwork and recording conventions 21

2 Repeating the unrepeatable experiment 23 Richard Bradley

3 Experimental archaeology at the crossroads: a contribution to interpretation or evidence of ‘xeroxing’? 42 Martin Bell

4 ‘Proportional representation’: multiple voices in archaeological interpretation at Çatalhöyük 59 Shahina Farid

5 Integrating database design and use into recording methodologies 79 Michael J. Rains vi Contents

6 The tyranny of typologies: evidential reasoning in Romano-Egyptian domestic archaeology 92 Anna Lucille Boozer

PART II Cross-field trade: archaeological applications of external expertise and technologies 111

7 The archaeological bazaar: scientific methods for sale? Or: ‘putting the “arch-” back into archaeometry’ 113 Mark Pollard and Peter Bray

8 Radiocarbon dating and archaeology: history,progress and present status 128 Sturt W.Manning

9 Using evidence from natural sciences in archaeology 159 David Killick

10 Working the digital: some thoughts from landscape archaeology 173 Marcos Llobera

11 Crafting knowledge with (digital) visual media in archaeology 189 Sara Perry

PART III Multiple working hypotheses, strategies of elimination, and triangulation 211

12 Uncertain on principle: combining lines of archaeological evidence to create chronologies 213 Alex Bayliss and Alasdair Whittle

13 Lessons from modelling Neolithic farming practice: methods of elimination 243 Amy Bogaard

14 Evidence, archaeology and law: an initial exploration 255 Roger M.Thomas Contents vii

15 Law and archaeology: Modified Wigmorean Analysis 271 Terence J.Anderson and William Twining

16 Traditional knowledge, archaeological evidence, and other ways of knowing 287 George Nicholas and Nola Markey

PART IV Broader perspectives: material culture as object and evidence 309

17 Evidence of what? On the possibilities of archaeological interpretation 311 Gavin Lucas

18 Meeting pasts halfway: a consideration of the ontology of material evidence in archaeology 324 Andrew Meirion Jones

19 Matter and facts: material culture and the history of science 339 Simon Werrett

Index 353 This page intentionally left blank FIGURES

2.1 Outline plan of South Lodge Camp 28 2.2 Outline plan of the later excavation of South Lodge Camp 30 2.3 A nineteenth-century photograph of the ‘Yak’s Quarter’ 32 2.4 Croftmoraig stone circle: (a) proposed structural sequence in 1971; (b) revised in 2012 34 2.5 (a) General view of Croftmoraig stone circle. (b) View towards the summit of Schiehallion 38 2.6 Midsummer sunset at Schiehallion viewed from the raised ground behind the stone circle 39 3.1 The experimental earthwork at Overton, Wiltshire, excavated 1992 48 3.2 The Moel-y-Gerddi roundhouse at Butser 49 3.3 Excavation of the Moel-y-Gaer roundhouse at St Fagans, 2009 50 3.4 Plan of artefact distributions in the Moel-y-Gaer roundhouse 51 3.5 (a) The wall line of the Moel-y-Gaer roundhouse. (b) The wattle-and-daub wall of the Moel-y-Gaer roundhouse with a decayed wood stake in the ground 54 4.1 Composite of single context plans of North Area, Çatalhöyük 62 4.2 Harris matrix of temporal succession of archaeological contexts at Çatalhöyük 62 4.3 Hodder’s 12-point plan towards reflexive methods 64 4.4 Priority tour of the foundation trench excavations, North Area, Çatalhöyük, 2007 68 4.5 Proforma recording forms in use in Building 77, North Area, Çatalhöyük, 2011 73 4.6 Team seminar in preparation for publication, Çatalhöyük, 2010 75 5.1 IADB – part of a stratigraphic matrix and a composite plan 81 5.2 A facsimile context record in the IADB 86 x List of figures

6.1 Karanis insula 100 6.2 House B2, Trimithis (Roman period Amheida) 103 6.3 House B2, reconstruction, Trimithis (Roman period Amheida) 105 7.1 Distribution of type 2 copper during the Early Bronze Age 120 7.2 Lead isotope ratio values for some Taunton Phase metalwork 125 8.1 Comparison of the original Libby 14C age expectation model versus the reality observed from known-age archives 132 8.2 Comparison of the Damon et al. (1974) 14C calibration curve versus the IntCal13 14C calibration curve of four decades later, and both versus the Libby model 133 8.3 Comparison of IntCal13 14C calibration curve (versus IntCal09, IntCal04, IntCal98 and the 1986 datasets) at (a) 0–50,000 years BP; (b) 35,000 to 25,000 BC; (c) 12,000 to 8500 BC; and (d) 1700 to 1200 BC 136 8.4 Comparison of the mid-point values of the previous 14C calibration datasets (IntCal09, IntCal04, IntCal98 and the 1986 dataset) versus IntCal13 137 8.5 (a) Comparison of the same radiocarbon date of 29,900 ± 35 14C years BP versus the two different radiocarbon calibration curves of IntCal09 and IntCal13 and (b) differing calendar age ranges determined for the same radiocarbon date of 29,900 ± 35 14C years BP, comparing IntCal09 with IntCal13 from (a) 138 8.6. Comparison of the IntCal13 (Northern Hemisphere) and SHCal13 (Southern Hemisphere) 14C calibration curves, 1000 BC to AD 1940 139 8.7 Early Cycladic sequence as published in Renfrew et al. (2012: Figure 9) 142 8.8 The calibrated age distributions for the Phase C to early Middle Cycladic Boundary 144 8.9 Adapted re-runs of the Renfrew et al. (2012: Figure 9) model 145 8.10 (a) Analysis of the 43 14C dates published on short-lived material from the Akrotiri volcanic destruction level. (b) Detail of the end Boundary ‘Thera volcanic destruction’ in (a). (c) End Boundary ‘Thera volcanic destruction’ from a re-run of the model in (a), excluding outliers 147 8.11 (a) Calibrated age distributions from a re-run and revised version of the Manning et al. (2006) sequence analysis of the Aegean early Late Bronze Age. (b) Some details of the modelled ranges. (c) Example of query: length of time (years) from the Akrotiri volcanic destruction level boundary = eruption date to the Late Minoan IB Late Destruction of Myrtos-Pyrgos 149 10.1 (a) Area of visibility for one mamoa calculated from single point. (b) Area of visibility for the same mamoa after using a set of points to represent it 175 List of figures xi

10.2 Density of path intersections for different scales of movement 179 10.3 Corridors of movement generated for medium and long-range movement 180 10.4 (a) Histogram showing the distribution of times that locations within the Barbanza peninsula are associated with certain horizon indexes. (b) Histogram showing the distribution of times when mamoas have been associated with certain distances from horizon locations 183 10.5 (a) Histogram showing the distribution of times mamoas were viewed within certain angles of elevation. (b) Histogram showing the distribution of mamoas as viewed within certain angles of incidence 185 11.1 The craftsmanship aesthetic 197 11.2 3D model merging the existing experimental house with the ‘hunting shrine’ at Çatalhöyük, south-east 202 11.3 3D model merging the existing experimental house with the ‘hunting shrine’ at Çatalhöyük, north 203 11.4 Plan of a proposed new experimental house at Çatalhöyük constructed from the visual outputs of multiple practitioners 204 12.1 Example of an up-to-date culture-historically based regional chronology 216 12.2 Probability distributions of dates from the Mesolithic use of the Howick hut 220 12.3 Schematic diagram showing the archaeological sequence in Figure 12.2 221 12.4 Probability distributions for the dates of occupation of Mesolithic huts analogous to Howick from the British Isles 222 12.5 Pottery traditions in Early Bronze Age Ireland 223 12.6 Probability distributions of dates associated with encrusted urns in Ireland 224 12.7 Schematic summary of the chronology of selected elements of the Early Bronze Age of Ireland 226 12.8 Research cycle employed in the Anglo-Saxon chronology project 227 12.9 Final correspondence analysis of male-grave assemblages and artefact types from the Anglo-Saxon chronology project 228 12.10 Probability distributions of dates from male-graves from the Anglo-Saxon chronology project 229 12.11 Schematic diagram showing date estimates for the appearance of Neolithic things and practices across selected areas of Britain and Ireland 232 12.12 Fussell’s Lodge long barrow 234 12.13 Probability distributions for key parameters for the dates of constructional events at Fussell’s Lodge long barrow 235 xii List of figures

14.1 Example of a Wigmorean chart 261 14.2 Simple example of an archaeological stratigraphic section and accompanying Harris matrix 263 14.3 Wigmorean chart for a fictitious archaeological example 266 15.1 Beginnings of Anderson’s macroscopic Wigmorean Analysis 280 16.1 Pre-contact clam garden, Waiatt Bay, Quadra Island, BC 293 16.2 Pictograph “dreamings” interpreted by Annie York, Stein River Valley, BC 296 16.3 Replicated pithouse, Secwepemc Heritage Park, Kamloops, BC 299 17.1 Abandoned house at the fishing village on Viđey, Iceland 312 18.1 Ormaig at night, illustrating the relationship between rock art and cracks and fissures 327 18.2 Excavating the clay platform on the eastern face of Tiger Rock, Torbhlaren 328 18.3 Linear fissure on the upper surface of Lion Rock, Torbhlaren 329 18.4 Selection of quartz hammerstones from Tiger Rock, Torbhlaren 332 MAPS

6.1 Map of Egypt (M. Matthews) 93 10.1 Location of the Barbanza peninsula in Galicia (NW Spain) 176 12.1 Map showing formally modelled date estimates (at 95 per cent probability, and at 68 per cent probability in brackets) for the spread of causewayed and related enclosures across southern Britain 231 This page intentionally left blank TABLES

7.1 Definition of copper types based on the presence or absence of arsenic, antimony, silver and nickel 119 7.2 Modelled depletion families for the 16 possible combinations of presence/absence of arsenic, antimony, silver and nickel 122 7.3 Ubiquity of copper types in Irish Early Bronze Age copper alloy assemblages 122 13.1 Relationship of the four major crop husbandry models to the three variables of permanence, intensity and seasonality 245 16.1 Secwepemc medicinal plants grouped according to medicinal use 294 This page intentionally left blank CONTRIBUTORS

Terence J. Anderson is Professor Emeritus, University of Miami School of Law, co-author (with William Twining) of Analysis of Evidence (1991, 1998; and with William Twining and David Schum, 2006) and author of ‘Wigmore Meets the Last Wedge’ (Twining and Hampshire-Monk, eds., Evidence and Inference in Law and History, 2003).

Alex Bayliss is Head of Scientific Dating at English Heritage and part-time Professor of Archaeological Science at the University of Stirling. She has pioneered the routine use of Bayesian statistics for chronological modelling, constructing hundreds of models from thousands of radiocarbon dates.

Martin Bell is Professor of Archaeological Science at the University of Reading, UK. His research interests are environmental archaeology, geoarchaeology, and experimental archaeology. He directed the Experimental Earthwork Project excavations in the 1990s and has excavated at several experimental sites.

Amy Bogaard (PhD Sheffield 2002) is Professor of Neolithic and Bronze Age Archaeology at the School of Archaeology, University of Oxford. Her research centres on the archaeology of farming practice.

Anna Lucille Boozer is an Assistant Professor of Roman Mediterranean Archaeology in the Department of History at Baruch College (City University of New York). Her research focuses on Roman Egypt, Meroitic Sudan, imperialism, and daily life. She directs the City University of New York excavations at Amheida (Egypt) and codirects the Meroe Archival Project (Sudan) with Intisar Elzein Soghayroun (University of Khartoum, Sudan). xviii List of contributors

Richard Bradley is Emeritus Professor of Archaeology at Reading University. His recent books include The Idea of Order (2013), Image and Audience (2011), Stages and Screens (2011), and The Prehistory of Britain and Ireland (2007).

Peter Bray is an AHRC Post-Doctoral Researcher at the Research Laboratory for Archaeology (RLAHA), University of Oxford. He has degrees from the Universities of Bradford and Oxford, and is a Junior Research Fellow at Linacre College.

Robert Chapman is Emeritus Professor of Archaeology at the University of Reading, with research interests including human inequalities and historical materialist approaches to archaeology. His publications include The Archaeology of Death (1981), Emerging Complexity (1990), and Archaeologies of Complexity (2003).

Shahina Farid is a field archaeologist who joined the excavations at Çatalhöyük in 1995 and was appointed Field Director 1997–12. Her main area of interest is in the good practice of excavation methods. She currently works for English Heritage.

Andrew Meirion Jones is a Reader in Archaeology at the University of Southampton (UK). His current research is on the decorated artefacts of Neolithic Britain. His recent books include Prehistoric Materialities (2012) and Archaeology after Interpretation (2013).

David Killick is Professor of Anthropology and Adjunct Professor of Materials Science and Engineering at the University of Arizona in Tucson.

Marcos Llobera (DPhil Oxon 1999) is an Associate Professor in Anthropology at the University of Washington. His main research interests are the development of computational models in archaeology (more broadly the development of archaeological information science), the design of new methods for landscape analysis, and the relation between archaeological practice and theory.

Gavin Lucas is a Professor of Archaeology at the University of Iceland. His main research interests lie in archaeological method and theory and the archaeology of the modern world. His most recent book was Understanding the Archaeological Record (2012).

Sturt W. Manning is currently Goldwin Smith Professor of Classical Archaeology, Director of the Cornell Institute of Archaeology and Material Studies, and Director of the Malcolm and Carolyn Wiener Laboratory for Aegean and Near Eastern Dendrochronology, Cornell University. For recent publications, see http://cornell. academia.edu/SturtWManning.

Nola Markey is a cultural anthropologist and archaeologist based in Kamloops, BC. Her experience includes project management of a variety of archaeological List of contributors xix assessments, environmental assessments, and community-based Aboriginal traditional use studies in British Columbia, the Yukon, and Ontario. She is also an archaeo- logical BC provincially certified Resource Inventory Standards Committee (RISC) course instructor. Nola is Saulteaux and is a member of the O-Chi-Chak-Ko-Sipi First Nation, Crane River, Manitoba.

George Nicholas is a Professor of Archaeology at Simon Fraser University, and Director of the Intellectual Property Issues in Cultural Heritage (IPinCH) project. He was the founding director of SFU’s Indigenous Archaeology Program in Kamloops, BC. His research focuses on Indigenous peoples and archaeology, intangible heritage, the archaeology and human ecology of wetlands, and archaeological theory.

Sara Perry is Director of Studies of Digital Heritage and Lecturer in Cultural Heritage Management in the Department of Archaeology at the University of York, UK.

Mark Pollard is Edward Hall Professor of Archaeological Science at the University of Oxford. His degrees are in physics, and on the way to Oxford he has held posts in Departments of Chemistry, Archaeology, and Archaeological Sciences.

Michael J. Rains has been responsible for the development of the Integrated Archaeological Database (IADB) at York Archaeological Trust. As a Visiting Research Fellow at the University of Reading, he has been involved with the use and development of the IADB within the Silchester Town Life project since its inception in 1996.

Roger M. Thomas studied Archaeology at the Universities of Southampton and Cambridge. He has worked for English Heritage since 1984 in a variety of roles. He also has a degree in Law, and is a non-practising barrister.

William Twining is Quain Professor of Jurisprudence Emeritus at University College London. His recent books include Analysis of Evidence (with Anderson and Schum, 2005), Rethinking Evidence (2006), General Jurisprudence (2009), and Evidence, Inference and Enquiry (ed. with Dawid and Vasilaki).

Simon Werrett is the author of Fireworks: Pyrotechnic Arts and Sciences in European History (2010). Since 2012, he has been a member of the Department of Science and Technology Studies at University College London.

Alasdair Whittle is Distinguished Research Professor in Archaeology at Cardiff University. He has specialised in the study of the Neolithic period across Europe. With Alex Bayliss and Frances Healy, he has published Gathering Time, and is currently leading, with Alex Bayliss, a major ERC-funded project on chronological modelling for Neolithic Europe. xx List of contributors

Alison Wylie is Professor of Philosophy and Anthropology at the University of Washington, and Professor of Philosophy at Durham University. She works on philosophical issues raised by archaeology: objectivity, evidential reasoning, and research ethics. Publications include Thinking from Things (2002), essays in Value- Free Science? (2007), Agnatology (2008), and Ethics of Archaeology (2012). 1 MATERIAL EVIDENCE

Learning from archaeological practice

Alison Wylie and Robert Chapman

A passion for things

In recent decades, the arcana of archaeology have come sharply into focus as a subject and a resource that humanists and social scientists cannot afford to ignore, however resolutely text-based or wedded to face-to-face, “reactive” modes of inquiry they may be. The “in-depth study of things,” declare the editors of the Object Reader (Candlin and Guins 2009: 2), has taken shape in a sprawling diversity of research programs, ranging from metaphysical interrogation of materiality as such to probing analyses of the ways in which meaning and matter are entangled in specific objects and contexts of action. Object biographies have captured popular imagination, in the form of the wildly successful History of the World in 100 Objects (MacGregor 2010), and now figure as prominently in the history of science and technology as in art history and cultural studies (Daston 2008). This attention to objects – this appreciation of the dynamic, consequential social lives of things – has catalyzed the formation of a distinct interdisciplinary field of material culture studies (Myers 2001: 5),1 one that now has a history of its own in which the insights that set the field in motion are themselves subject to critical scrutiny. To insist that things be seen as a medium through which the social is articulated and meaning communicated is now decried as a “colonization of the object by the subject and the social” (Candlin and Guins 2009: 4), charged with trading in the very Cartesian oppositions between mind and matter that such an approach was meant to displace (Henare et al. 2007: 1–3).2 The turn to things – objects, bodies, artifacts, traces – is thus reinforced by renewed insistence that objects must be engaged in material as well as symbolic and social terms. In all these contexts object studies are compelling, not only because the stuff of lives lived is intrinsically interesting and is constitutive of those lives, but because it is invaluable as evidence. Thinking with (or through, or about) things has opened 2 Alison Wylie and Robert Chapman up otherwise inaccessible areas of inquiry and it has reconfigured our understanding of a great many longstanding topics of social scientific interest, from the dynamics of popular culture to the form and logic of political regimes (Auslander 1996), from the condensation of value and the nature of commodities to the ramifying construction of social difference and solidarity (Appadurai 1986; Myers 2001). The brief for assembling A History of the World in 100 Objects was to “tell a history of the world that [had] not been attempted before,” one that is “truer,” more compre- hensive and, crucially, “more equitable than one based solely on texts” (MacGregor 2010: xv, xxv, xix). Objects and traces have the potential to “give voice” to those who left no texts, to contest history as written by elites and victors, to bear witness to dimensions of life that no one thinks to tell, or that are actively suppressed. This commitment to explore the kinds of history that “only a thing can tell” (MacGregor 2010: xxii) is not new. “History from below” has been championed at least since the Marxist Historians’ Group took shape in the UK after World War II, and it has antecedents dating to the 1930s (Beard 1935; Becker 1932). Although E. P. Thompson’s Making of the English Working Class (1963) is the most widely cited example,3 Rodney Hilton is especially interesting because his research on medieval peasantry led him to initiate, with archaeologist Phillip Rahtz, the excavation of a deserted medieval village at Upton (Goucestershire); evidence of people’s houses and everyday possessions had the potential, he thought, to enlarge the scope of inquiry beyond dependence on texts.4 Expanding on the tradition of class-based analysis, feminist, critical race, and postcolonial historians, to name a few, have demonstrated just how different history looks when centered on the lives of those who have largely been written out of account. But to tell these counterhistories – of the everyday, of habit and localized practice, of the marginal – requires considerable ingenuity, reading canonical texts against the grain, expanding the archive to include what had been dismissed as ephemera, and, crucially, drawing on the non-textual evidence afforded by physical traces and material things. Historical archaeologists have been especially forthright in insisting that rigorous scrutiny of material evidence is not just a supplement to text-based histories but often the only resource we have for exposing and correcting “superficial and elitist . . . myths[s] for the contemporary power structure” (Glassie 1977: 29): the systematic distortions that arise from ignoring “the inarticulate” (Ascher 1974: 11), the “endless silent majority who did not leave us written projections of their minds” (Glassie 1977: 29). These themes are taken up by contributors to a recent discussion on “Historians and the Study of Material Culture” in the American Historical Review (Auslander et al. 2009), which begins with the observation that, “while some might still associate [the study of material culture] with objects found in museums or things from the remote past, it is in fact a field that takes an interest in all conceivable objects and every historical period”; it is especially relevant to any historical subject that takes as its subject a “concern for everyday life and the material circumstances of ordinary people” (AHR editor, Auschlander et al. 2009: 1355). Parallel arguments for attending to material evidence also figure in sociology, although for more strictly methodological reasons. In a classic of the 1960s, the Material evidence: learning from archaeological practice 3 proponents of Unobtrusive Measures (Webb et al. 1966) made the case that, given the inescapable limitations of “reactive” methods, it is folly to proceed “simply by asking,” whether this takes the form of participant observation or structured interviews, surveys or experimental interventions. They detail a range of interactive dynamics and interviewer or intervention effects that arise from the ways in which subjects manage their self-presentation in response to what they perceive as the expectations of a research setting, compromising the internal and external validity of standard research methods in the social sciences (Webb et al. 1966: Chapter 1). The only way forward is to engage the resources of multiple methods, including underdeveloped strategies for using inadvertently produced physical traces as evidence of patterns of action, preference, and intention that survey respondents or interviewees might not themselves be aware of, or might be disinclined to disclose (Webb et al. 1966: 3, 34). In a vigorous renewal of the case for “revaloriz- ing sources marginalized by dominant social science,” Lee emphasizes the value of “ephemeral traces” of movement and interaction in a social environment – physical erosion or accretion, the litter discarded, the “performative opportunities” afforded by objects – as “caches of data” that make possible strategies of triangulation (Webb et al. 1966: xx). The principle here is that evidence from very different sources, in this case archival and material, should be mobilized as an independent basis for assessing the results of reactive methods (Lee 2000: 1, 8, 14). The Tucson “Garbage Project,” initiated by archaeologist William Rathje in the early 1970s and later expanded internationally, embodies a similar rationale: the systematic analysis of what we throw away, recovered from curbside garbage collection and through the excavation of landfill sites, often reveals patterns of consumption that stand in stark contrast with the results of surveys that depend on self-reports (Rathje and Murphy 1992). In all these areas, then, an enthusiasm for the capacity of material things and physical traces to function as evidence reflects an appreciation of their stubborn concreteness, the “brute intransigence of matter,” as Daston puts it (2008 [2004]: 11), by virtue of which they are sometimes seen as impartial witnesses to the past, bearing marks of their makers, their various uses, and the shifting configurations of meaning and action in which they have been implicated in the course of their travels. However enigmatic they may be, this “bony materiality” sustains a certain epistemic optimism, even when the more naïve aspects of a “positivist histori- ography of facts” have been abandoned (Daston 2008 [2004]: 15–16). But for all this, a recurrent theme in the literature valorizing objects as subject and source is a paucity of attention to things themselves. This concern figures prominently in reflection on formative examples of object studies dating to the 1990s. Gell rejected sociological and iconographic alternatives to the “aesthetic preoccupations” of then-contemporary anthropology of art on grounds that they effectively ignore “the art object itself”; they “look . . . only at the power [of the object] to mark distinctions” or treat it as a “species of writing” and consider only its symbolic meaning (Gell 1992: 43).5 Similarly, Corn took aim at “object myths” in the history of technology. Reviewing publications through the 1990s he found that, despite a 4 Alison Wylie and Robert Chapman rhetoric privileging the object, a substantial majority of historians of technical artifacts relied exclusively on textual or oral sources; fewer than a third had undertaken object-based analysis and typically only for periods or contexts for which textual sources are unavailable (Corn 1996: 37). Although this has certainly changed, even historians who are committed to the study of material culture lament the fact that “we observe the march of images into the historian’s study, but physical objects are mostly kept at arm’s length”; material culture may be a subject of study, but material things and traces are still not “among the standard resources of academic research in the humanities and social sciences” (Sibum in Auslander et al. 2009: 1384). This is evident in the Object Reader and Things that Talk; text-based analysis of the social and symbolic significance of objects continues to dominate the study of material culture, both as a subject in its own right and as a resource for social and historical inquiry of other sorts. As Simon Werrett argues in his contribution to this volume, even though historians of science have largely abandoned intellectual history and cultivated an interest in the material culture of science, “they have only done so with a relatively limited disciplinary repertoire”; they tend to proceed by “reading about things rather than engaging with them directly” (see Chapter 19 in this volume, pp. 346). This disconnection between intent and practice should come as no surprise. The challenges of working with material evidence are legendary, and the prejudices of the literate continue to favor the seeming transparency of textual records and the direct testimony of subjects.6 These difficulties are reinforced by traditions of disciplinary specialization that set the human, social sciences apart from the biological and physical, reproducing a divide not unlike like that lamented by C. P. Snow in the late 1950s (1963 [1959]). To make effective use of physical traces and objects as evidence requires expertise in an enormous range of fields, most of them not ones in which humanists and social scientists have any training. Consider, for example, Marc Bloch’s reflection on the demands of research in historical ecology:

Now, if almost any important human problem . . . demands the handling of diverse types of evidence . . . the types of evidence necessarily mark off the several branches of technical scholarship. The apprenticeship for each is long, but full mastery demands still longer and almost constant practice. For example, very few scholars can boast that they are equally well equipped to read critically a medieval charter, to explain correctly the etymology of place- names, to date unerringly the ruins of dwellings of the prehistoric, Celtic, or Gallo-Roman periods, and to analyze the plant life proper to a pasture, a field, or a moor. Without all these, however, how could one pretend to describe the history of land use? (Bloch 1953: 68)7

Add to this catalogue of necessary expertise the resources of material science and physical geology, art historical interpretation and ethnohistory, the sciences of Material evidence: learning from archaeological practice 5 cognition and of biomedicine, all fields required to understand the cultural context and human implications of land use patterns, and the complexity of the enterprise is apparent. The gulf between these various fields is not always characterized by the stark incomprehension that concerned Snow, but the range of expertise required to ground material culture studies in rigorous analysis of the material – to meet MacGregor’s demand that objects be “interrogated and interpreted as deeply and rigorously as the written report (MacGregor 2010: xvi) – is daunting in its variety. As the editor moderating the American Historical Review discussion observes, there are a great many questions that “strike at the heart of methodological concerns we don’t often discuss” (Auslander et al. 2009: 1355); “to take cues from the objects themselves,” historians will have to “work hard to develop methods in order to make speak the silent representatives of the past” (Sibum in Auslander et al. 2009: 1359).

The wisdom of practice It is primarily archaeologists who have taken on these challenges, and yet they figure very little in the object studies literature. Since the early nineteenth century they have built up a repertoire of research strategies specifically designed to mobilize the evidence of human lives and events that survives in an enormous range of material evidence, from garbage dumps to monuments, from the physical traces of single events to the palimpsest of evidence that bears witness to long-term, large-scale cultural processes, from finely crafted artifacts rich with cultural significance to the inadvertent traces left by human activity on a continent-wide scale and in the minutiae of domestic consumption. In the process they have decisively enlarged, challenged, and reconfigured what we know, putting material evidence to work in the investigation of a great many different aspects of the cultural past. The tempo and process of development varies by region, but from the early nineteenth to the mid-twentieth centuries archaeologists built relative and absolute chronologies and made use of material evidence as a proxy for cultural affinity, focusing on artifacts in settlement and burial contexts, and on monumental architecture. By the mid- twentieth century in many areas these local histories of material culture had been integrated into regional and inter-regional typologies, and, with growing attention to botanical and faunal remains, they delineated distinct ecological and subsistence regimes and major cultural transitions. The ecological contexts of people’s everyday lives and subsistence practices became the subject of pioneering studies in Scandinavian and British archaeologies from the 1930s and were influential in the Americas in the 1950s, providing the basis for refined models of regional settlement and subsistence patterns of the kind developed by Gordon Willey in the Viru Valley, Peru (1953) and by Richard MacNeish in the Sierra de Tamaulipas, Mexico (1958). The processual archaeologists of the 1960s and 1970s built on such studies by analyzing the underlying adaptive dynamics of cultural systems and constructing large-scale, long-term systemic models of their social, political, and economic structures (see examples in Binford and Binford 1968). Through the 1980s, in 6 Alison Wylie and Robert Chapman reaction against this system-level eco-materialism, postprocessual archaeologists developed a program of material culture studies focused on symbolic meaning and agency (e.g. Hodder 1982). Archaeologists had long been interested in social identities – for example, in the form of enclaves like the Oaxaca barrio identified by Millon in the city of Teotihuacan (1973) – and, from the early 1990s, in gender, race, and ethnic identity (e.g. Gilchrist 1999; Battle-Baptiste 2011; Leach et al. 2009). In developments over the last two decades that parallel the object studies literature, there has been much focus on the materiality of objects marked by debate about what distinguishes materiality from materials (e.g. Ingold 2007; Knappett 2012), and about objects as “things” and their entanglements with the social, the ideational, and diverse forms of agency (e.g. Olsen 2010; Hodder 2012). The question that animates this collection of essays is, then: how do archae- ologists do this? How do they make effective use of physical traces and objects as repositories of evidence? There is a great deal of wisdom embodied in the practices by which archaeologists assemble, analyze, integrate, and adjudicate physical evidence, but all too often this wisdom has been obscured by sharply polarized internal debate about the limitations of material evidence. There have always been epistemic pessimists who insist that, insofar as claims about the past overreach the available evidence and the evidence itself is an interpretive construct, the conclusions that archaeologists draw can be nothing more than speculation. In an especially strong statement of this position in the mid-1950s, M. A. Smith insisted that there is no “necessary link” between the surviving record and “the human activities we should like to know about”; the inference from one to the other is, therefore, “a hopeless task” (Smith 1955: 3–4). Similar concerns have surfaced repeatedly in Anglo-American archaeology since the early twentieth century (see Wylie 2002: Part I), but, just as regularly, they have been rebutted by epistemic optimists who insist that the limitations lie not in the material record itself but in the intellectual and material resources archaeologists bring to bear in recovering and interpreting it. The advocates of in the 1960s and 1970s were particularly adamant that, if archaeologists adopted a properly scientific hypothesis-testing methodology they could meet the gold standard of deductive certainty they associated with rigorous scientific inquiry (Binford 1962, 1972; Watson et al. 1971). Postprocessual critics were quick to object that archaeological evidence is also a construct, therefore just as insecure a ground for testing as for building hypotheses about the cultural past. Some insisted, on these grounds, that relativist conclusions are inescapable: “there is literally nothing independent of theory or propositions to test against” (Shanks and Tilley 1987: 111); archaeologists should candidly acknowledge that they simply “create facts” (Hodder 1983: 6). The upshot, as Killick describes it, was that the “stories archaeologists tell are no more than politically charged fables imposed from the present onto the past” (Killick, Chapter 9 of this volume: p. 161). Few held such extreme positions for long; although intransigent in their rejection of the deductivist ideals of the New Archaeology, postprocessual critics nonetheless acknowledged that archaeological data have an impressive capacity to “resist theoretical appropriation” (Shanks and Material evidence: learning from archaeological practice 7

Tilley 1989: 44), and New Archaeologists made it a priority to build a robust, empirically grounded body of “middle range theory” to underpin the interpretation of surviving material traces as evidence. Although the “theory wars” have largely died down in recent years (Johnson 2010: 220-3), the epistemic anxieties and ambitions that gave them impetus continue to structure archaeological thinking. For example, in his contribution to this volume, Gavin Lucas traces the contrasting legacies of metaphors for the archaeological record that embody, on one hand, a “negative epistemology” – the archaeological record as “fragment,” incomplete and compromised – and, on the other, a more hopeful emphasis on survival and continuity into the present associated with the metaphor of “relic” (see Chapter 17, p. 314). For the most part debate about the epistemic status of archaeological evidence has been conducted at a level of abstraction that provides little useful guidance for practice. And although there are a great many handbooks on the practicalities of field work and techniques of analysis, archaeologists have published comparatively little that captures the underlying rationale for specific forms of practice, in the mid- space between “how to” guides and philosophical generalities. What is needed, we suggest, is resolutely case-based analysis of actual practice – key instances of exemplary practice, critical turning points, innovations, and instructive failures in the use of archaeological data as evidence – aimed at making explicit the norms of evidential reasoning that have taken shape in the context of evolving traditions of practical experience working with archaeological material. This is the goal we had in view when we took up the project of assembling this collection of essays; we draw inspiration from David Clarke’s brief for an internal philosophy of science that addresses the challenges faced by archaeologists, rather than imposing models developed to make sense of physics (Clarke 1973), and we see this work as an extension of Ian Hodder’s call for systematic reflection on the dynamics of “archaeological process” (Hodder 1999). To this end we asked our contributors to identify a particular aspect of evidential reasoning with which they grapple and to consider, with reference to concrete examples in their area of specialization, how they construct evidential claims, critically assess them, and bring them to bear on pivotal questions about the cultural past. We asked, more specifically, that they identify best practices and draw out lessons learned, both positive and negative, as a basis for articulating constructive guidelines for practice. The structure of the volume reflects key junctures at which questions about evidential reasoning arise in archaeological practice, as well as highlighting cross-cutting issues with which practitioners grapple in a wide range of fields, wherever they confront the challenge of “building scientific knowledge in the absence of infallible foundations,” as Hasok Chang puts it in a philosophical history of chemistry (Chang 2004: 234). We outline first the rationale for the overall structure of the volume, and then trace cross-cutting themes that knit it together. 8 Alison Wylie and Robert Chapman

What follows: overall structure We begin with a set of essays on “Fieldwork and recording conventions” (Part I). In the first of these, Richard Bradley makes explicit the insight that frames this section and much that follows: established norms of “good” excavation and survey practice, and conventions of rigorous recording and documentation, were designed to answer specific questions, which, in turn, presuppose a rich array of assumptions about the nature of the subject domain, about what is puzzling or interesting, and, crucially, what can feasibly be asked about it. They are purpose-built creatures of context that have become standardized and persist in training and practice long after the original questions are forgotten. Critical histories of archaeological practice recover these starting points and reveal how deeply they configure subsequent research: Bradley traces the legacy of Pitt-Rivers’ now canonical nineteenth-century excavations at Cranborne Chase (Chapter 2); Anna Boozer scrutinizes the shaky foundations of a typology of domestic architecture, lost to collective consciousness, that underpin contemporary descriptions and comparisons of house forms in Roman Egyptian archaeology (Chapter 6), a theme later taken up by Sturt Manning (in Chapter 8) when he reflects on the legacy of a chronology flawed by “a poorly based best guess” in Late Bronze Age Aegean archaeology (Chapter 8: p. 150); and Martin Bell warns against the risks of “xeroxing,” in which new research simply reproduces accepted results (Chapter 3, p. 50). The central question here is: to what extent do original purposes constrain the potential for addressing new questions and putting old evidence to work? Does the process of selection and of interpretation “at the trowel’s edge” (Hodder 1997) entail that archaeologists will recognize, collect, record, and categorize material in ways that answer to their expectations, as the most intransigent epistemic pessimists have suspected? In fact, while the cases considered in this section illustrate how deeply entrenched the conventions of past practice can be, the stubborn intransigence of material evidence is also clearly on view, as remarked upon by Shanks and Tilley (1989: 44); they draw attention to a range of ways in which practitioners can, and do, effectively come to terms with these conventions. Bell details experimental strategies for recognizing and counteracting “the influence of pre-understandings” (Chapter 3: p. 55); Shahina Farid assesses the successes and challenges of the reflexive, interactive pluralism instituted by Hodder at Çatalhöyük (Chapter 4); Bradley and Boozer both illustrate the value of critical histories designed to interrogate the limitations of past practice, the better to build on it (Chapters 2 and 6); and, turning to the impact of computing technologies on field recording, Michael Rains reflects on the variable impact of new computer technologies on field recording and data management practices, a current and ongoing process that throws into relief the need to tune new technologies to the evolving goals of inquiry as well as to existing working practices (Chapter 5). The essays that comprise the next section, “Cross-field trade” (Part II), address the complex web of interdisciplinary engagements by which archaeologists meet the challenges outlined by Bloch. Archaeology is resolutely a “trading zone”; throughout its history its horizons as a discipline have been expanded dramatically Material evidence: learning from archaeological practice 9 and continuously by importing expertise and technologies developed elsewhere for application to archaeological problems of data recovery and analysis. At the same time, this import trade has been bedeviled by miscommunication and unrealistic expectations about what these external resources can deliver. The result is a recurrent boom-and-bust lifecycle of enthusiastic early adoption followed by frustration and either premature abandonment or a long, difficult process of “settling in”: retooling external resources for archaeological purposes. The contri- butors to this section offer jointly historical and pragmatic assessments of how this process has unfolded in a number of key areas of engagement. Three consider the ongoing trade with high-profile physical sciences: Sturt Manning traces the fortunes of radiocarbon dating through three revolutions (Chapter 8); David Killick juxtaposes with this the challenges of realizing comparable precision and accuracy in other physical dating methods, for example refining stable isotope analysis as a tool for dietary reconstruction (Chapter 9); and Mark Pollard and Peter Bray reassess archaeometry at a point where technical refinement has destabilized the original questions about provenance that set the field in motion in the nineteenth century (Chapter 7). Marcos Llobera echoes many elements of these retrospective assessments of longstanding cross-field engagements, now with respect to newly emerging digital simulation techniques: it is crucial, he argues, to cultivate a sophisticated understanding of what these tools can (and cannot) do, to keep archaeological questions clearly in focus, and to purpose-build applications that address these questions rather than allowing tools designed for other purposes to dictate the archaeological agenda (Chapter 10). In a telling counterpoint to these examples of the impact of new technoscience resources, Sara Perry considers the long history of visualization, drawing attention to the pivotal but often discounted role played by professional illustrators in archaeology; their seemingly prosaic practice requires a high order of skill that embodies a misrecognized genre of “epistemic productivity” (Chapter 11: p. 198). Taken together, these essays bring into sharp focus a number of factors that determine the viability of cross-field trade in external expertise: institutional and funding structures, as well as internal hierarchies of expertise and labor, that constrain or enable effective transfers and collaborations; training that puts archaeologists in a position to adjudicate external expertise and successfully tune it to archaeological purposes; and a keen reflective appreciation of how the research agenda of the discipline has been both served and circumscribed by cross-field trade. It is by now a truism that archaeologists almost never depend on a single line of evidential reasoning; as the histories outlined in Part II illustrate, they make use of a wide range of distinct lines of evidence to assess and to calibrate even the most robustly grounded scientific evidential claims. Crucially, as the advocates of “unobtrusive measures” in sociology appreciate, the value of mobilizing multi- ple lines of evidence lies in their capacity to be mutually constraining as well as mutually reinforcing; triangulation raises the epistemic credibility of the claims they support insofar as the independence between distinct lines of evidence counteracts the risk of error in any one of them. The contributors to Part III, “Multiple 10 Alison Wylie and Robert Chapman working hypotheses,” illustrate various strategies by which archaeologists make an epistemic virtue out of what often seems a liability, bringing disparate lines of evidence into play in the construction and adjudication of interpretive claims of various kinds and scales. Amy Bogaard takes up themes articulated by Bradley and Bell, showing how an open-ended, iterative process of experimental testing that operates on a number of fronts at once – archaeological, ethnobotanical, ecological – makes possible a fine-grained comparison of alternative models of early Neolithic farming practice (Chapter 13). In a complement to Manning’s essay, Alex Bayliss and Alasdair Whittle illustrate how a Bayesian modeling approach enables them to systematically integrate multiple lines of evidence and rigorously assess margins of error in physical dating – the third radiocarbon revolution – in some cases providing temporal resolution that brings human-scale action and events into sharp focus (Chapter 12). In two essays on archaeology and the law, an archae- ologist, Roger Thomas (Chapter 14), and legal scholars Terence Anderson and William Twining (Chapter 15) make a case for applying to archaeological problems a system for schematically representing and assessing “mixed masses of evidence” which was developed in the context of early twentieth-century legal theory by John Henry Wigmore. Finally, expanding the frame of interpretation, George Nicholas and Nola Markey consider a diversity of ways in which lines of evidence originating in Indigenous oral traditions can be brought into productive engage- ment with archaeological research (Chapter 16). It is the friction between different types of expertise and epistemic standpoints, as well as between independent lines of evidence, that is productive in these cases. In the final section, “Broader perspectives: material culture as object and evidence” (Part IV), we close with three essays that take up directly the questions that animate the volume as a whole: about the presuppositions that configure archaeological practice, and what this practice has to offer the range of other fields in which there is growing interest in making use of material evidence. Gavin Lucas scrutinizes the ontological assumptions that underpin opposing, dominant con- ceptions of the archaeological record; he asks what this “record” is evidence of, and urges us to “re-imagine” material traces as surviving elements of enduring assemblages that integrate material with practice, and social worlds with agency of diverse kinds (Chapter 17, p. 321). Extending this reflection on agency, Andrew Jones articulates a parallel shift in the framework of assumptions that set the terms of archaeological engagement, one that closely tracks the arguments now dominant in material culture studies for taking material things seriously on their own terms (Chapter 18). Rather than assume that archaeological materials are “mute” and “inert” until brought to life by our intervention – animated or induced to “talk,” to use Daston’s metaphor (Daston 2008) – Jones explores the potential of an “intra- active” approach, inspired by Karen Barad’s agential realism (2007), in the context of a study of late prehistoric rock art; his practice is predicated on a recognition of and engagement with the distinctive “agentive potentiality of objects, things or materials” (Chapter 18, p. 334). In the final essay of the volume, Simon Werrett reverses the direction of inquiry and asks what historians of science might learn Material evidence: learning from archaeological practice 11 from archaeological practice (Chapter 19). His account of how material evidence has figured in historical inquiry sharpens the ambivalence Daston notes, and his response to this ambivalence sets against “the fragility of material evidence” (Chapter 19, p. 349) an appreciation of how archaeological analyses of, experi- mentation with, and modeling of material objects can enlarge the “disciplinary repertoire” of fields like the history of science in which there is, increasingly, an acute sense of their relevance, as subjects, protagonists, and crucial resources for inquiry.

Cross-cutting themes: getting things to talk in archaeology The sectional divisions we have outlined are, in their nature, somewhat arbitrary; there are all kinds of overlaps in the types of cases considered, the focal issues addressed, and the lessons drawn. Reading across the grain, all the contributors grapple, in one way or another, with what Boozer identifies as a paradox at the heart of archaeological practice (Chapter 6, p. 93), and in this they take up the threads of the internal debate about the limitations of material evidence. She describes material evidence as tangible in terms that resonate with Daston’s appraisal of material evidence. It can surprise and challenge settled assumptions in ways that reinforce the sense, reported by Daston, that it stands as an impartial witness to what happened in its past, but at the same time it is radically ambiguous; its significance as evidence is a function of sedimented layers of interpretation that can easily reproduce and amplify error. There is no escape from what Bell describes as “pre-understandings” (after Hodder 1999) and these configure not just the interpretive or explanatory claims that are the focus of inquiry, but what will count as evidence in building or adjudicating them. There is always the possibility that, as Thomas puts it, the assumptions that guide inquiry will “blind one to evidence that might lead in a different direction” (Chapter 14, p. 257), and they can do this in insidious ways. They may take the form of “value judgements masquerading as factual propositions” operating well below the threshold of conscious awareness (Anderson and Twining, Chapter 15 p. 275); they are embedded in conventions of representation, as Perry, Llobera, and Rains illustrate in quite different ways (recording forms, database structure, conventions of visualization); they are internalized through long training and embodied in skilled empirical practice, in archaeology (Bell) as in the sciences generally (Werrett); and, as Farid shows, taking up Hodder’s “trowel’s edge” argument, they are often built into the fabric of archaeological practice in the form of conventional divisions of labor and work discipline. The upshot, as Bradley encapsulates this line of thinking, is that the kind of “complete objectivity” – in the sense of freedom from interpretive presuppositions – that he finds associated with the technical virtuosity of General Pitt-Rivers and endorsed by Philip Barker (Chapter 2, p. 23) is, in principle, unrealizable. It stands in stark contrast to the messy realities of the bootstrapping and scaffolding practices by which archaeologists necessarily rely on shaky background assumptions to get 12 Alison Wylie and Robert Chapman inquiry off the ground, and then wrestle with the constraints that these tentative foundations impose on subsequent research; reliance on pre-understandings is an inescapable “condition of possibility” for archaeology (Jones, Chapter 18, p. 325). But rather than take these pre-understandings to be all of a kind and equally insecure, as Smith had done and as suggested (briefly) by some of the more intemperate postprocessual critiques, a great many different kinds of presupposition figure in the examples that follow and it is clear that they vary widely in the degree to which they are credible, in themselves and in archaeological application. As Anderson and Twining describe the background assumptions that underpin evidence-based inferences in legal arguments, they can include everything from “well-grounded science [and] evolving science [to] dubious, contested or junk science”; “practical expertise” and “everyday experience”; robust generalizations and one-off anecdotes; common sense, hearsay, “speculation . . . faith . . . prejudice” (Chapter 15, p. 275). There are archaeological counterparts to all of these identified in the case studies presented here: the robust background knowledge drawn from the physical sciences discussed by Pollard and Bray, Manning, Bayliss and Whittle, and Killick that, nonetheless, requires a great deal of scaffolding to be useful in archaeological contexts; the emerging information technologies appraised by Rains and by Llobera; the embodied expertise of professional excavators and illustrators discussed by Farid and by Perry; the rich but often discounted oral traditions of Indigenous-descendant communities highlighted by Nicholas and Markey; the contrasting culture-wide conventions of Western thought, like the “distinction between nature and culture” discussed by Bradley and by Werrett, and the empirical, archaeology-specific pre-understandings drawn from type sites and local sequences that are the anchors for regional typologies and chronologies of the kinds discussed by Boozer, Bell, and Manning. As the details of the case studies make clear, however, archaeologists do routinely call out presuppositions of these various kinds when they have become ingrained in practice and hold them accountable, conceptually and empirically. In fact, despite bringing the uncertainties of evidential reasoning into sharp focus, there are a number of striking narratives of success outlined here: chronologies calibrated, typologies reconfigured, focal questions reframed, specification of the range of application of “middle-range theory,” effective tuning of new technologies to archaeological problems, and nuanced adjudication of the claims for and against seemingly undecidable alternative working hypotheses. It is also clear that, to do this – to build, and continuously rebuild, credible background knowledge in the range of areas required to sustain archaeological reasoning from material evidence – requires a repertoire of critical and constructive methods that are as diverse as the presuppositions on which archaeologists rely. Here is a shortlist of strategies for accomplishing this that stand out for us as editors in the accounts that follow. First, and most obviously, there are all the strategies by which archaeologists generate new data: as Bell suggests, it ought not be so difficult to “escape from the tyranny of background knowledge and pre-understandings. The discipline is, after all, refreshed by a constant stream of new discoveries” (Chapter 3, p. 55). Of course, Material evidence: learning from archaeological practice 13 while some empirical discoveries arise unbidden and unanticipated, even these are consequential only given a prepared mind, and just as often striking new empirical discoveries depend on deliberate innovation – of questions, of search space, of technical tools that bring new empirical data on stream. The long process of reconciling discordant chronologies detailed by Manning, and the dramatic refram- ing of settled assumptions about cultural sequence and associations that results, as illustrated by Bayliss and Whittle as well as Manning, reinforces the point that transformative empirical discoveries depend heavily on technical scaffolding. So too the process of coming to terms with the compositional complexity of metal objects outlined by Pollard and Bray, where the weight of hard-won evidence unsettles the assumptions that inform a “narrowly defined ‘scientific’ approach”; they describe a growing recognition that repeated melting, mixing, and repurposing is not the exception but the norm and that the chemical components of these objects are themselves dynamic, which, in turn, prompts a fundamental rethinking of what it is that archaeometallurgists study: “not merely a set of material processes but also . . . the human decisions and structures that surround them” (Chapter 7, p. 125). The data force the point, but they come into view and are salient only given the evolving scaffolding of jointly physical and social scientific background knowledge that Pollard and Bray describe. This means that the action, where evidential reasoning is concerned, is largely off-stage: it is not so much a matter of discovering new data, as important as this can be, but of painstakingly building and continuously reworking the necessary background knowledge in an ongoing iterative process, much like the process that Chang describes for chemistry, where, he insists, there are no fixed, self-warranting foundations (Chang 2004: 228–231). The most direct ways of carrying out these ideas involve experimental and “practice-based” research (Bell, Chapter 3, p. 55) designed to test assumptions about practices and processes that produce, or that could have produced, specific types of archaeological trace. Long-term studies of the erosion of earthworks (Bell), “practical experimentation” with quartz assemblages (Jones, Chapter 18, p. 333), the chemical analysis of metal decay and transformation just described (Pollard and Bray), sharply focused testing of ethnographic analogies (Killick), and experimentation with crop yield and weed profiles under different farming scenarios (Bogaard) are all examples presented here of approaches that are now a staple of archaeological practice. Crucially, as Bell argues, this is not typically a process of establishing uniformitarian principles that can secure deductive certainty; it is, rather, a matter of specifying with as much precision as possible the conditions under which linking principles are likely to hold in specific contexts of application. The aim, Manning argues with reference to the third radiocarbon revolution, is to replace the “web of a priori assumptions” that underpins conventional chronologies, not with the certainty of an imagined silver bullet, but with systematic modeling of error (Chapter 8: p. 151), a practice that Bayliss and Whittle illustrate in a series of cases ranging from highly local human-scale occupations to large-scale landscapes and millennial timescapes. In addition, several contributors make the case for recognizing what might be 14 Alison Wylie and Robert Chapman described as indirect evidence generated by experimenting with models themselves: for example, the recognition of gestural knowledge as constitutive of experimental practice that arises from the reenactments by historians of science that Werrett describes (Chapter 19, p. 342); the sensitivity analyses Bayliss and Whittle use to assess alternative chronological models (Chapter 12, pp. 233–236); the distinctive interpretive insights that arise from visualizations described by Perry (Chapter 11) and by Farid (Chapter 4); and the methods of simulation that, Llobera argues, make it possible to explore “basic properties of the study area,” extending and con- textualizing field observation (Chapter 10, p. 186). When material evidence is put to work building, assessing, testing, and refining the models and hypotheses that are the focus of inquiry, as indicated earlier it is almost invariably in the context of a complex web of evidential reasoning. The contributors identify several different ways in which and purposes for which multiple lines of evidence are mobilized. For example: Bogaard makes the case for developing proxies for the farming practices she studies that are based on completely different datasets (Chapter 13, p. 250). The impact of bringing new archaeological comparanda to bear figures in Bradley’s and Lucas’ discussions of the circumstances that made it possible to recognize timber structures in the archaeological record, and is the motivation for Boozer’s insistence on the need to reassess architectural typologies in Roman Egyptian archaeology. Killick emphasizes the capacity for purpose-built lines of material science-based evidence to provide an independent test of interpretive assumptions that derive from sources that have dominated interpretation, like ethnography or common sense (e.g. about dependence on maize or the labor requirements of earthwork construction). And Bayliss and Whittle advocate a “pragmatic” Bayesian approach precisely because it is “a formal and explicit methodology for weaving together different strands of evidence” (Chapter 12, p. 217). As different as these cases are, they share a common rationale for expanding the range of evidence: to exploit the potential for indepen- dent lines of evidence to constrain one another, counteracting the risks of “xerox- ing” (Bell, Chapter 3) and identifying error in individual lines of evidence that might not otherwise be recognized, as well as to enrich interpretation, bringing previously inscrutable aspects of the past into view. A recurrent theme emphasized by a number of contributors is the need for rigorous transparency about the constituents of these webs of inference and the “analytical scaffolding” (Boozer, Chapter 6, p. 104) that underpins them. This is a particular virtue of Bayesian modeling – “it force[s] archaeologists to be explicit about their strands of reasoning” (Bayliss and Whittle, Chapter 12: p. 218) – and it is the reason why Thomas, and Anderson and Twining, advocate the application to archaeological problems of a modified Wigmore method for schematizing the empirical and inferential components of evidential reasoning. It is only possible to build on and refine tentative foundations in an ongoing, iterative process if the bases for earlier inferences are kept clearly in view. Citing Gero (2007), Boozer urges a commitment to “honor ambiguity” rather than smoothing, cleaning, or otherwise suppressing the messiness and uncertainty of the descriptive and inter- Material evidence: learning from archaeological practice 15 pretive claims that become the basis for subsequent reasoning with archaeological data (Chapter 6, p. 105). By extension, it is crucial to resist the kind of canalization of thinking that comes from investment in a seemingly “best” hypothesis. In this spirit, Bayliss and Whittle warn against the potential of provisional models to “fossilise” and “become received wisdom” (Chapter 12, p. 233), Nicholas and Markey note the risks of “intellectual complacency” that come with considering only the range of evidence and hypotheses that have become familiar staples of disciplinary inquiry (Chapter 16, p. 302), and Killick warns against a “‘good enough’ attitude towards scientific evidence” (Chapter 9, p. 168). They join Bogaard, among others, in stressing the importance of keeping multiple working hypotheses in play even while working to reduce the range of viable options by elimination testing. Model and hypothesis evaluation is typically comparative; the outcome can only be a (tentative) judgment about the relative strengths of the alternatives considered, which carries with it the risk that an “inference to the best explanation” will endorse the best of a set of limited set of possibilities. It is crucial, Bogaard argues, to guard against “propensity to infer that the ‘best’ explanation at any given moment is true”; it may only be “best” given a failure to conceive of relevant other possibilities (Chapter 13, pp. 252, 249–250). One way to realize transparency about underpinnings and alternatives is to put critical historiography to work as a way of taking distance from entrenched conventions and mobilizing the wisdom of hindsight. Bradley urges “a kind of historiography” in the context of reengaging legacy data that involves not just returning to the data and records generated by past excavations, but situating them in a broader motivational, social, and political context. This holds the potential, he argues, for realizing a degree of “repeatability”: rerunning the process of inter- pretation in light of new archaeological comparanda and with an awareness of how broad cultural sensibilities have shifted (Chapter 2, p. 25). It puts him in a position to recognize the cultural significance of geological features that had been dismissed as natural. In this spirit a number of other contributors develop genealogies of key concepts or modes of practice that configure contemporary archaeology: Lucas scrutinizes the legacy, in our understanding of “the archaeological record,” of metaphors of survival (relic) and fragmentation (ruin) rooted in nineteenth-century traditions of Romanticism and antiquarianism (Chapter 17, pp. 313–317); Perry considers the institutional and disciplinary history of archaeological illustration (Chapter 11); and those who discuss archaeological science trace the tangled histories of “cross-field trade.” Another way to promote transparency is to cultivate pluralism of a number of different kinds. The commitment to make use of the full “toolkit of available approaches” (Bell, Chapter 3, p. 42) is not just a matter of exploiting all available resources; extending this maxim, Farid highlights how the friction between specialisms, various forms of expertise, and standpoints can be enormously produc- tive in mobilizing critical and constructive insights that no one party brought to the table in the first instance (Chapter 4). She and Perry argue for recognizing diverse forms of “epistemic productivity” (Perry, Chapter 11, p. 191), including forms of 16 Alison Wylie and Robert Chapman expertise that are marginalized within archaeology, while Nicholas and Markey illustrate the value of taking seriously external resources that have selectively been read out of consideration: “even contradictory lines of evidence have value,” throwing into relief what has come to be “common sense” for archaeologists and opening up interpretive and explanatory horizons (Nicholas and Markey, Chapter 16, p. 301). A key advantage of Wigmore schemas as described by Anderson and Twining is that they require clarification of the standpoint from which an evidential argument has been developed, so that it can be evaluated from multiple perspectives. A pluralism of standpoints is a key resource for generating a repertoire of multiple working hypotheses that have the potential to counter the worries Bogaard raises about unconceived alternatives. This is reflexivity made concrete. It is not primarily a matter of introspection: we are often the last to be aware of the taken-for-granteds that inform our own practice. Rather, it is a form of (collective) self-appraisal that depends on putting second-order empirical methods of analysis to work – historical and genealogical, comparative and contextualizing – to bring into focus pre- understandings embodied in specific questions and contexts of inquiry now forgotten or internalized as default disciplinary or cultural assumptions. The reflexive appraisals of evidential reasoning presented here have a number of implications for orienting ideals and goals, and for the pragmatics of practice. They suggest, for example, the need to rethink the epistemic ideals that inform programmatic debate and practice. Rather than take “objectivity” to be self- evidently a matter of transcending context or insulating against the influence of pre-understandings, the argument emerging here suggests the need to unpack the specific epistemic virtues gathered together (uneasily) under the aegis of this term, and consider in concrete terms how they are best realized in particular contexts of practice, for particular purposes. We see a resolutely pragmatic, proceduralist conception of objectivity taking shape in these contributions, one that challenges the vernacular positivism that has set the terms of debate between processual and postprocessual archaeologists and that requires a more nuanced account of epistemic success. This reassessment extends as well to the practicalities of archaeological inquiry: how archaeologists are trained and how research is organized and funded. Killick and Pollard and Bray particularly emphasize the need for training designed to cultivate the kind of scientific literacy necessary for archaeologists to make effective, well-informed use of scientific analyses of archaeological material. The perils of treating scientific analysis of archaeological material as a “self-service buffet” (Pollard and Bray, Chapter 7, p. 114), appropriating the results of poorly under- stood specialist techniques without a clear sense of archaeological purpose, is reiterated with reference to digital technologies by Llobera and by Rains (Chapters 10 and 5). The central message here is that the silos of specialism that set disciplinary archaeology apart from the myriad other fields on which it depends constitute a structural constraint that archaeologists must continuously struggle to overcome. Perry and Farid make a similar point with respect to hierarchies of expertise within archaeology. The success of evidential reasoning depends on establishing institu- Material evidence: learning from archaeological practice 17 tional and funding structures, work processes and channels of communication that can sustain “productive and trustworthy forms of cooperation” (Killick, Chapter 9, p. 161) between archaeologists and those with whom they trade, internally and externally. This is to reaffirm the point made by Bloch in the 1950s, articulated here with attention to practicalities which make it clear that effective engagement with things will not be accomplished by tooling up new expertise within discrete fields. The successes working with material evidence that have been realized by archaeologists depend in large measure on creating archaeology as a dynamic trading zone (Galison 2010), building an expansive network of technical, empirical, theoretical exchange underpinned by the kind of “diplomatic tradition” Nicholas and Markey describe with reference to the Haudenosaunee. It is to be hoped that the passion for things now manifest in a growing range of neighboring fields may be an opening for extending this network.

Acknowledgments The planning for this book began between January and July 2010, when Alison held a Leverhulme Visiting Professorship in the Department of Archaeology at the University of Reading (UK). Bob had taken the initiative of sponsoring Alison for the professorship, and secured the Leverhulme funding that made this visit possible. We are grateful to the Leverhulme Trust and to the staff and postgraduate students of the Department of Archaeology for their support, as well as the School of Human and Environmental Sciences (as it then was) for hosting a workshop on “Material Culture as Evidence” in June 2010. The discussions generated by that workshop and the seminar series on evidential reasoning that we convened in the Department of Archaeology at the University of Reading were the catalyst for this project and several of the participants have contributed to this volume. We are also grateful to the University of Washington for providing the leave that made Alison’s extended visit at Reading possible, and to colleagues in the many departments of Archaeology and of Philosophy who hosted Leverhulme lectures and seminars while she was in the UK. Those conversations were also crucial to shaping our vision for this collection of essays. We thank all the contributors for their wholehearted commitment to the themes developed in this book, and for being willing to engage in productive, positive, and critical dialogue about their papers. Our special thanks are extended to Katie Banks, who acted as an enthusiastic and efficient editorial assistant. We could not have pulled everything together without her unerring eye for detail and much appreciate her unflagging optimism that this project would, indeed, be realized. Finally, thanks to Matt Gibbons at Routledge for his positive response to our enquiry about the potential publication of this volume, and to Amy Davis-Poynter, also at Routledge, for her patient and informative responses to our seemingly endless questions. 18 Alison Wylie and Robert Chapman

Notes 1 While Myers (2001) identifies material culture studies as taking shape since the 1990s, Miller describes entering a field already formed in the early 1980s (2005: 4). The impetus was, in any case, a critique of the presumption that objects are static media in which social meaning is reflected, “inert and mute, set in motion and animated, indeed knowable, only by persons and their words . . . divorced from the capacity of persons to act and the power of words to communicate” (Appadurai 1986: 4). 2 At an extreme of animating objects and attending to “things themselves,” the arguments for embracing distributed, hybrid forms of agency – the “vital materialism” of Bennett’s Vibrant Matter (2010), Haraway’s cyborgs (1991), and Latour’s hybrids (1993) – resonate with the case made by recent advocates of object ethnography for refusing Cartesian dichotomies altogether: mind and matter, meaning and material are inseparable, the locus of proliferating ontologies (Henare et al. 2007: 3). 3 In addition to E. P. Thompson, this group included Christopher Hill, Rodney Hilton, and Eric Hobsbawm, among others. 4 This collaboration was not a happy or successful experience. Although the excavations at Upton made a contribution to archaeological research on deserted medieval villages, Hilton found the fieldwork “rather tedious” (Rahtz 2001: 91). There was no integration of historical texts and archaeological evidence in the publication (Hilton and Rahtz 1966). 5 Jones extends this objection in his contribution to this volume, finding in Gell an example of the limiting conception of “material agency” that he means to push beyond. Although Gell advocates closer attention to material objects, on Jones’ account he nonetheless treats them as “a derived form of agency,” reproducing Cartesian dichotomies that reduce their “agentive potentialities”; even though art objects exist in a nexus of relations, in the end they “stand in for” human relations and intentions (see Chapter 18, pp. 333–334). 6 Sibum goes on to observe that there continues to be a systematic bias in favor of text- based scholarship; the study of literary sources is recognized as historical “scholarship” while close study of material objects is demeaned as a form of connoisseurship (Auslander et al. 2009: 1384). 7 We were alerted to this passage of Bloch’s The Historian’s Craft by Carole Crumley, who quotes it, in her own translation, in the Foreword to Advances in Historical Ecology (1998: ix).

References Appadurai, A. (ed.) (1986) The Social Life of Things: Commodities in Cultural Perspective, Cambridge: Cambridge University Press. Ascher, R. (1974) “Tin*Can Archaeology,” Historical Archaeology 8: 7–16. Auslander, L. (1996) Taste and Power: Furnishing Modern France, Berkeley: California University Press. Auslander, L., Bentley, A., Halevi, L., Sibum, H. O., and Witmore, C. (2009) “AHR Conversation: Historians and the Study of Material Culture,” The American Historical Review 114(5): 1355–1404. Barad, K. (2007) Meeting the Universe Halfway: Quantum Physics and the Entanglement of Matter and Meaning, Durham, NC: Duke University Press. Battle-Baptiste, W. (2011) Black Feminist Archaeology, Walnut Creek, CA: Left Coast Press. Beard, C. A. (1935) “Written History as an Act of Faith,” The American Historical Review 39(2): 219–229. Becker, C. L. (1932) “Everyman his own Historian,” The American Historical Review 37(2): 221–236. Bennett, J. (2010) Vibrant Matter: A Political Ecology of Things, Durham, NC: Duke University Press. Material evidence: learning from archaeological practice 19

Binford, L. R. (1962) “Archeology as Anthropology,” American Antiquity 28(2): 217–225. Binford, L. R. (1972) An Archaeological Perspective, New York: Seminar Press. Binford, S. R. and Binford, L. R. (eds) (1968) New Perspectives in Archeology, New York: Aldine. Bloch, M. (1953) The Historian’s Craft: Reflections on the Nature and Uses of History and the Techniques and Methods of Those Who Write It, translated by P. Putnam, New York: Vintage. Candlin, F. and Guins, R. (eds) (2009) The Object Reader, New York: Routledge. Chang, H. (2004) Inventing Temperature: Measurement and Scientific Progress, Oxford: Oxford University Press. Chippindale, C. (2002) “Capta and data,” American Antiquity 65(4): 605–612. Clarke, D. L. (1973) “Archaeology: The Loss of Innocence,” Antiquity 47: 6–18. Corn, J. J. (1996) “Object Lessons/Object Myths? What Historians of Technology Learn from Things,” in D. W. Kingery (ed.) Learning from Things: Method and Theory of Material Culture Studies, Washington, DC: Smithsonian Institution Press, 35–54. Crumley, C. L. (1998) “Foreword,” in W. Balée (ed.) Advances in Historical Ecology, New York: Columbia University Press, ix–xiv. Daston, L. (ed.) (2008 [2004]) Things That Talk: Object Lessons From Art and Science, New York: Zone Books. Galison, P. (2010) “Trading with the Enemy,” in M. E. Gorman (ed.) Trading Zones and Interactional Expertise, Cambridge, MA: MIT Press, 25–32. Gell, A. (1992) “The Technology of Enchantment and the Enchantment of Technology,” in J. Coote and A. Shelton (eds) Anthropology, Art and Aesthetics, Oxford: Clarendon, 40–63. Gero, J. M. (2007) “Honoring Ambiguity/Problematizing Certitude,” Journal of Archaeological Method and Theory 14: 311–327. Gilchrist, R. (1999) Gender and Archaeology: Contesting the Past, London: Routledge. Glassie (1977) “Archaeology and Folklore: Common Anxieties, Common Hopes,” in L. Ferguson (ed.) Historical Archaeology and the Importance of Material Things, Special Publication Series No. 2, Winnipeg, Manitoba: Society for Historical Archaeology, 23–35. Haraway, D. (1991) “A Cyborg Manifesto: Science, Technology, and Socialist-feminism in the Late Twentieth Century,” in D. Haraway, Simians, Cyborgs, and Women: The Reinvention of Nature, New York: Routledge, 148–181. Henare, A., Holbraad, M. and Wastell, S. (eds) (2007) Thinking Through Things: Theorising Artefacts Ethnographically, London: Routledge. Hilton, R. H. and Rahtz, P. A. (1966) “Upton, Gloucestershire, 1959–64,” in Transactions of the Bristol and Gloucestershire Archaeological Society 85: 70–146. Hodder, I. (ed.) (1982) Symbolic and Structural Archaeology, Cambridge: Cambridge University Press. Hodder, I. (1983) “Archaeology, Ideology and Contemporary Society,” Royal Anthropological Institute News 56: 6–7. Hodder, I. (1997) “Always Momentary, Fluid and Flexible: Towards a Reflexive Excavation Methodology,” Antiquity 71(273): 691–700. Hodder, I. (1999) The Archaeological Process: An Introduction, Oxford: Blackwell Publishers. Hodder, I. (2012) Entangled: An Archaeology of the Relationships Between Humans and Things, Oxford: Wiley Blackwell. Ingold, T. (2007) “Materials Against Materiality,” Archaeological Dialogues 14(1): 1–16. Johnson, M. (2010) Archaeological Theory: An Introduction, Oxford: Wiley Blackwell. Kingery, D. W. (ed.) (1996) Learning from Things: Method And Theory of Material Culture Studies, Washington, DC: Smithsonian Institute Press. 20 Alison Wylie and Robert Chapman

Knappett, C. (2012) “Materiality,” in I. Hodder (ed.) Archaeological Theory Today, 2nd edn, Cambridge: Polity, 188–207. Latour, B. (1993) “The Proliferation of Hybrids,” in B. Latour, We Have Never Been Modern, Hemel Hempstead: Harvester Wheatsheaf, 1–3. Leach, S., Lewis, M., Chenery, C., Müldner, G. and Eckardt, H. (2009) “Migration and Diversity in Roman Britain: A Multidisciplinary Approach to the Identification of Immigrants in Roman York, England,” American Journal of Physical Anthropology 140(3): 546–61. Lee, R. E. (2000) Unobtrusive Measures in Social Research, Buckingham: Open University Press. MacGregor, N. (ed.) (2010) A History of the World in 100 Objects, New York: Penguin Books. MacNeish, R. S. (1958) “Preliminary Archaeological Investigation in the Sierra de Tamaulipas,” Transactions of the American Philosophical Society 48(6): 1–210. Myers, F. R. (ed.) (2001) The Empire of Things: Regimes of Value and Material Culture, Santa Fe: SAR Press. Miller, D. (ed.) (2005) Materiality, Durham, NC: Duke University Press. Millon, R. (1973) The Teotihuacan Map, Part One: Text, Austin, TX: University of Texas Press. Olsen, B. (2010) In Defence of Things: Archaeology and the Ontology of Objects, Lanham: Altamira Press. Pettit, P. (1975) The Concept of Structuralism: A Critical Analysis, Dublin: Gill and Macmillan. Rahtz, P. (2001) Living Archaeology, Stroud: Tempus. Rathje, W. and Murphy, C. (1992) Rubbish! The Archaeology of Garbage, New York: Harper Collins. Schiffer, M. B. (1976) Behavioral Archeology, New York: Academic Press. Shanks, M. and Tilley, C. (1987) Re-Constructing Archaeology, Cambridge: Cambridge University Press. Shanks, M. and Tilley, C. (1989) “Archaeology Into the 1990s: Questions Rather Than Answers,” Norwegian Archaeological Review 22: 42–45. Smith, M. A. (1955) “The Limitations of Inference in Archaeology,” Archaeological Newsletter 6: 1–7. Snow, C. P. (1963 [1959]) The Two Cultures: A Second Look, Cambridge: Cambridge University Press. Taylor, W. W. (1948) A Study of Archeology, Carbondale, IL: Southern Illinois University Press. Thompson, E. P. (1963) The Making of the English Working Class, Toronto: Penguin Books. Tilley, C. (ed.) (1990) Reading Material Culture, Oxford: Basil Blackwell. Watson, P. J., LeBlanc, S. A. and Redman, C. L. (1971) Explanation in Archaeology: An Explicitly Scientific Approach, New York: Columbia University Press. Webb, E. J., Campbell, D. T., Schwartz, R. D. and Sechrest, L. (1966) Unobtrusive Measures: Nonreactive Research in the Social Sciences, Chicago, IL: Rand McNally. Willey, G. R. (1953) Prehistoric Settlement Patterns in the Viru Valley, Peru, Bureau of American Ethnology Bulletin 155, Washington, DC: Smithsonian Institution. Wylie, A. (2002) Thinking from Things: Essays in the Philosophy of Archaeology, Berkeley, CA: University of California Press. PART I Fieldwork and recording conventions This page intentionally left blank 2 REPEATING THE UNREPEATABLE EXPERIMENT

Richard Bradley

The first chapter of Philip Barker’s book Techniques of Archaeological Excavation is called ‘The unrepeatable experiment’. This is why:

Every archaeological site is itself a document. It can be read by a skilled excavator, but it is destroyed by the very process which enables us to read it. Unlike the study of an ancient document, the study of a site by excavation is an unrepeatable experiment . . . Since no two archaeological sites are the same, either in the whole or in detail, it is never possible to verify conclu- sively the results of one excavation by another. (Barker 1977: 12)

His approach set limits on the usefulness of excavated evidence, but at the same time it illustrated a conception of fieldwork that was not shared with all his colleagues. Martin Carver’s book Archaeological Investigation describes it in this way:

Barker, arch-exponent of empirical field archaeology, decided that asking what was there, or what had happened there, was the only valid question one may ask of a site . . . Why have a pet theory? . . . He advocated large areas (area excavation) and the detailed digging of everything (total excavation), even parts that were not thought significant to any research programme. (Carver 2009: 27)

Barker’s position is logical if the only purpose of excavation is to describe what is observed in the field. It underlies the idea that a well-conducted project amounts to ‘preservation by record’: the deposits that are removed are replaced by a drawn and written archive, supplemented by any artefacts recovered during the work. There are problems with this position. It assumes a stance of complete objectivity 24 Richard Bradley that not even Barker could achieve, and at times he acknowledged this. It also implies that field archaeologists are unconcerned with broader questions. By applying standardised techniques to a variety of different deposits, do they really meet their obligations to posterity? It is all too easy to forget that those field methods were devised as ways of answering questions, the details of which have often been forgotten. What began as an intellectual enquiry can easily become a technical exercise. To write human history using archaeological evidence it is necessary to make the assumption that few phenomena in the past were entirely unique. The process of synthesis must involve comparison. In the case of excavation those comparisons are between the results of different projects. How much variation can be identified across time and space, and to what extent have different researchers made similar observations to one another? General models cannot be based on single instances, but, taken to a logical conclusion, that is what Barker advocated. It is why his publications influence the practice of field archaeology but play a restricted role in interpretations of the past. This chapter focuses on the work of two British archaeologists with entirely different objectives, and on the information gained by renewing excavation at the places where they had worked. The first is General Pitt-Rivers (1827–1900), whose monographs published between 1887 and 1898 were such an inspiration for later fieldworkers that one of his site plans is reproduced as the end paper of Barker’s book. Pitt-Rivers is considered the founder of a school of meticulous recording, whose concern with accurate description and predilection for operating on a large scale anticipated the methods used in the twentieth century. He was the earliest proponent of the idea of preservation by record (although the term was invented later) and is regarded by many archaeologists as a disciplinary ancestor. This is particularly true in commercial archaeology, where many projects happen as part of the planning process (Bradley 2006). It is here that Barker’s influence has been greatest. He, on the other hand, devoted most of his career to research excavations, which were subject to fewer outside pressures. The second archaeologist is Stuart Piggott (1910–1996). He was primarily a culture historian with a special interest in the period between the origins of agriculture and the Roman age. His most famous book is Ancient Europe (Piggott 1965), but he also had a lifelong interest in the history of ideas and the early development of archaeology. He combined studies of prehistoric artefacts with wider interpretations of the past and conducted a series of excavations in England and Scotland. Almost twenty years before Barker’s book was published, Piggott had distanced himself from the concept of the unrepeatable experiment:

Excavation is often compared with a scientific experiment in a laboratory, but the analogy must not be pushed too far. To think that archaeological evidence is of the same order as that used in, for instance, chemistry is to misunderstand its nature . . . What the excavator extracts from his dig will Repeating the unrepeatable experiment 25

depend on his past experience, his comparative knowledge of the material he is unearthing, his quickness of apprehension and a hundred other qualities peculiar to the individual. (Piggott 1959: 14)

The contrast between the two men goes even farther. There is no doubt that Pitt-Rivers was a brilliant technician, but, far from advocating a stance of complete objectivity, he was known to his contemporaries as a theorist whose research was influenced by Darwin (Thompson 1977; Bradley 1983; Bowden 1991). Once that is understood, the character of his excavations seems very different. In Piggott’s case it is easier to identify the influences affecting his fieldwork. He made it clear which questions determined his choice of projects, and for that reason his methods are easier to understand than those employed by the General. Even so, they were coloured by certain undeclared assumptions, and to interpret his results these must first be identified. In both cases this calls for a kind of historiography. In the case of Pitt-Rivers it involves an investigation not just of the monographs documenting his excavations but also of his writings outside archaeology, the most important of which were collected after his death in a book entitled The Evolution of Culture and Other Essays (Pitt-Rivers 1906) It is equally important to understand his personal political beliefs and his views on public education. Piggott said much less about theoretical questions, but his stance was deceptive. His enormous personal library contained books on philosophy and anthropology and, like a number of his colleagues, his approach to archaeology was influenced by Robin Collingwood (1889–1943), whose best-known book is The Idea of History (Collingwood 1946). Like Collingwood, Piggott believed that archaeological research should be governed by a process of working through questions and answers, and he put that method into practice. This account discusses the re-excavation of two prehistoric monuments investigated by these influential figures. The first is South Lodge Camp in southern England, excavated by General Pitt-Rivers in 1893, and re-examined by John Barrett and this writer between 1977 and 1981 (Barrett et al. 1991: 44–83). The second is the Scottish stone circle of Croftmoraig (sometimes spelt Croft Moraig), investigated by Stuart Piggott and Derek Simpson in 1965 and revisited nearly fifty years afterwards (Piggott and Simpson 1971; Bradley and Sheridan 2005; Bradley in prep). In each case similar questions arise. If new research is conducted at places that have already been excavated, it requires a detailed understanding of how and why the original work was undertaken. It is unlikely to employ exactly the same techniques and what is observed in the ground may not be the same. To suppose otherwise is to be deceived by the idea of objectivity that Piggott questioned half a century ago. 26 Richard Bradley

South Lodge Camp: excavations in 1893 Why did Pitt-Rivers excavate South Lodge Camp, and why was it necessary to return there eighty-four years later? Pitt-Rivers is often claimed as an unrepentant empiricist. He was a country landowner and excavated the ancient monuments on his property on an unprecedented scale. For Collingwood, who combined his profession as an Oxford philosopher with research on Roman Britain, the General lacked ideas:

Pitt-Rivers was a very great archaeologist and a supreme master in the techniques of excavation; but as regards the problems to be solved by excavation he was for the most part (not quite consistently) in the pre- Baconian stage. He dug in order to see what he could find out. (Collingwood 1939: 125)

This assessment was believed for a long time, but it was based on a misunder- standing. For twentieth-century archaeologists Pitt-Rivers was a technician of genius, whose field methods anticipated their own practice. To some extent that is because he was chosen as a role model by another military man – Sir Mortimer Wheeler. Wheeler was impressed by the scale on which the General had operated and by the discipline he brought to his fieldwork and its publication. He praised Pitt-Rivers for providing a full record of his observations and aspired to do the same (Wheeler 1954: 11–14 and 182–184). The General’s reports could be re-read in the light of more recent discoveries. In fact, that is what happened when Christopher Hawkes and Stuart Piggott presented a new interpretation of a series of settlements and monuments excavated by the General over fifty years before (Hawkes and Piggott 1947). What all these writers ignored was Pitt-Rivers’s own account of why he practised archaeology. In earlier life, when Pitt-Rivers’s family name was Lane-Fox, he had devised a system for teaching musketry in the British army. At the same time he was persuaded that minor developments in the histories of weapons and other artefacts followed similar principles to those observed by Charles Darwin in the natural world. Lane-Fox considered that the principle of ‘utility’ played a comparable role to the survival of the fittest. He took this observation even further, linking the gradual development of different kinds of objects to similar processes – as he saw them – in the society of his day. By drawing an analogy between the evolution of these artefacts and the communities who used them, he convinced himself that radical change was unnatural and dangerous to human wellbeing (Bradley 1983; Bowden 1991). That essentially conservative message could be taught through museum displays. To this end he amassed an enormous collection of objects. They could be arranged in series to show how variations in material culture proceeded extremely slowly. This was the lesson he wanted to teach the public. He arranged his collections to illustrate the principle of gradual development, but there was a weakness in his approach, for these displays consisted of material Repeating the unrepeatable experiment 27 made and collected during the nineteenth century. The ‘sequences’ he sought to illustrate were largely hypothetical and were based on the idea that societies living in different parts of the world could be treated as representative of different stages of social evolution. What he thought of as ‘primitive’ or ‘savage’ artefacts were contrasted with their modern equivalents. His approach could be strengthened by using artefacts from the distant past. Inspired by the geological model of stratigraphy, Lane-Fox turned to excavation, for he considered that this method would be particularly suitable for demonstrating how artefacts changed over time. It would add substance to his ideal scheme and would also provide a medium for public education. From his initial studies, which were addressed to fellow officers, he sought to engage a wider audience, but the political message was the same. The study of material objects, in the past and in the present, demonstrated the unnatural character of sudden change. When Lane-Fox inherited a large estate and changed his name to Pitt-Rivers, his objectives remained the same but he conducted his fieldwork on a larger scale. It is true that he insisted on a comprehensive record of what was found, but his publications and museum displays show that his adherence to a kind of social Darwinism was never modified (Bradley 1983); all that happened is that he became more engrossed in the details of the archaeological record, engaging in practical experiments, analysing human remains, carbonised cereals and animal bones, and placing an increasing emphasis on settlements rather than cemeteries. His field methods are revealing. He recorded his finds by depth from the surface, employing a geological model. He catalogued them by the same method in his ‘relic tables’. He drew direct comparisons between these discoveries and ethno- graphic artefacts in his own collection. But he did not work according to either of the methods attributed to him by his successors. He did not distinguish between different stratigraphic deposits, preferring to dig in arbitrary spits, and he did not clear large open areas in the manner suggested by the plans of his excavations. Instead the work moved across the site in a narrow strip. Once the base of the trench had been exposed, it was refilled and the process began again. As a result, it is not surprising that he was unable to discern much patterning among the features that he found there. After twenty years of large-scale fieldwork his real concern was still with the evolution of material culture. That should have been evident from his account of South Lodge Camp published in 1898 (see Bowden 1991: 126–129). It was a small, roughly square ditched enclosure, located a short distance from a group of burial mounds. The 1893 excavation embraced the filling of the enclosure ditch and almost the entire area inside it (see Figure 2.1). His account of the work emphasised the paucity of subsoil features inside the earthwork and paid more attention to the large number of finds – ceramics, lithics, animal bones and metalwork – found at different depths in the deposits. In the absence of more detailed information, Pitt-Rivers concluded that the enclosure might have been a compound for animals. His report also discussed the evolution of chevron-decorated pottery and drew on the finds from this project and ethnographic examples. South Lodge Camp 1893

Field boundary

Trough

020m

Field boundary

Unexcavated Earthworks Ditch

FIGURE 2.1 Outline plan of South Lodge Camp, showing the area excavated in 1893, the two features found inside it by General Pitt-Rivers, and the earthworks of the field system located in 1925. The trough was reported by the first excavator as an ‘empty grave’. A small pit, rediscovered in 1979 can be identified as a posthole. (Drawing: Sarah Lucas; copyright: author.) Repeating the unrepeatable experiment 29

South Lodge Camp: re-excavation in 1977–1981 For many years, the General’s findings were taken literally, and it was only after eighty years that Stuart Piggott expressed doubts as to whether a nineteenth century fieldworker – however famous – would have been able to identify the traces left by ephemeral timber buildings (Piggott 1973: 396–397). It seemed possible that South Lodge Camp had been a settlement rather than a stock enclosure. The idea was tempting in view of the large quantity of pottery and worked flint found in 1893. Pitt-Rivers himself had not considered this possibility as his reading of ethnography suggested that prehistoric houses could not be found by excavation. Piggott’s question was important, as enclosures like South Lodge Camp were comparatively unusual. Most examples had been levelled by the plough, and those which had been examined on a large scale were also investigated by Pitt-Rivers. If there were doubts over his interpretation, they would extend to these earthworks as well. Another reason for returning to the site was to consider the results of a survey undertaken in 1925, when one of the General’s former assistants observed that the earthwork enclosure had been superimposed on the remains of an existing field system (Toms 1925). This made it less likely that South Lodge Camp was used by pastoralists, as later commentators had suggested. It was at this point that Pitt-Rivers’s field methods posed problems. Would his practice of exposing the bedrock in a narrow trench have made it harder to identify the remains of domestic buildings? And would his method of excavating in uniform spits have obscured the relationship between the enclosure and the field system? Between 1977 and 1981 the new project shed light on both these questions. As a result, it is possible to re-read the General’s excavation report and to interpret his observations in a different way. Three conclusions are of interest here, as none of them could have been derived from Pitt-Rivers’s own record, despite its seeming objectivity (Barrett et al. 1991: 144–183). The earthwork enclosed two houses, a deposit of burnt flint and a series of pits, few of which had been observed in 1893. The site was by no means empty and it is now known that it contained at least two wooden buildings of the kind known at many settlements of the mid-second millennium BC (see Figure 2.2). They would have been difficult to identify without stripping large areas of the bedrock, especially as South Lodge Camp was situated in an area of intractable geology. Second, it became obvious that the enclosure had been superimposed on a pre- existing field system, the full extent of which had not been apparent before. Although its surface remains are visible today, Pitt-Rivers did not see them. That is hardly surprising because this category of field monument was first recognised after his death. In any case the General’s method of excavating in spits would have obscured the relationship between these features. The same applies to the relations between any pits, postholes and ditches that impinged on one another. They were not considered in the General’s excavations. The phasing of the Iron Age and Roman settlements re-analysed by Hawkes and Piggott in 1947 was based on finds of diagnostic artefacts. South Lodge Camp 1977–81

Field boundary

Mound of burnt flint

Trough

020m

Field boundary

Excavation Earthworks Ditch Roundhouse

FIGURE 2.2 Outline plan of the later excavation of South Lodge Camp, showing the position of the trough found in 1893, the postholes of two roundhouses and a mound of burnt flint. (Drawing: Sarah Lucas; copyright: author.) Repeating the unrepeatable experiment 31

Last, the new project established that the ditch around South Lodge Camp had been refilled soon after it was first dug and that this process concealed a few artefacts that had been placed at pivotal points in its layout. They included several pieces of metalwork and an enormous urn, which remained intact. It is not clear why this happened, but it does mean that the earthwork did not silt up gradually in the way that Pitt-Rivers had supposed. For that reason, its filling would never have preserved a straightforward sequence of artefacts. His method of excavation was designed with an objective that it could never have achieved. The 1893 excavation of South Lodge Camp is important because its results were published in detail. Over the next eight decades they were reinterpreted, but entirely in terms of what the General had observed. That is consistent with the empirical approach favoured by Philip Barker, and at first sight it means that any limitations of that project are explained by errors made in the field. The difference between the results of these two excavations has another cause: each project was pursued with its own objectives, and the techniques which each employed were related to the questions being asked. In 1977–1981, the filling of Pitt-Rivers’s excavation was removed in its entirety to look for the traces of buildings that he would have been unable to recognise. In 1893, however, the main emphasis was on the vertical sequence of deposits, and the presence or absence of postholes was of little consequence. It is ironic that Pitt-Rivers constructed buildings of the kind that he had missed in his fieldwork as picnic bowers for visitors to his estate (see Figure 2.3; see also Pitt-Rivers n.d.). The General’s main objective was to establish the sequence of artefacts preserved in the enclosure ditch. This was entirely in keeping with his interest in evolutionary theory. Indeed, it is strange that the report on the re-excavation of South Lodge Camp has been claimed as an example of post-processual archaeology, while it is the previous excavation that had followed a theoretical agenda.

Croftmoraig stone circle: excavations in 1965 Croftmoraig is a prehistoric stone circle in the southern Highlands of Scotland. It was first excavated by Stuart Piggott and his former student Derek Simpson in 1965. The results were published six years later. The work was undertaken ‘with a view to eventual conservation as a possible site for guardianship protection’ (Piggott and Simpson 1971: 1). In the event, the monument could not be opened to the public because no space could be found for a car park. That might suggest that the project was confined to preparing the circle for display, but this assumption would be misleading. In fact, this was one of a series of excavations at well-preserved structures chosen because they epitomised different kinds of megalithic architecture. The objective of the work was to employ modern archaeological techniques to establish the date, character and development of particular monument types so that they could find their place in a wider account of insular prehistory. Piggott and Simpson were part of the team that had already excavated at Stonehenge, and Piggott himself had been responsible for investigating FIGURE 2.3 A nineteenth-century photograph of the ‘Yak’s Quarter’, one of the picnic bowers erected by Pitt-Rivers for visitors to his pleasure ground. He was unable to detect the remains of comparable buildings at South Lodge Camp. Repeating the unrepeatable experiment 33 the stone circle at Avebury. Following his appointment to Edinburgh University, he embarked on a series of excavations at representative Scottish monuments. The work at Croftmoraig formed part of that programme (Bradley 2004). Both the excavators wrote syntheses of British prehistory (Piggott 1949; Megaw and Simpson 1979) and knew that this task could only be achieved if it was based on well-documented excavations. For that reason, it is hardly surprising that the results of their work at Croftmoraig have played a significant part in more general accounts of prehistoric monuments – stone circles in particular (Burl 2000). Although Piggott and Simpson were unable to obtain radiocarbon dates for the site, their other objectives were met. The monument was almost completely excavated, and over a period of little more than three weeks approximately 90 per cent of the site was cleared to the natural subsoil. A small quantity of pottery was recovered, together with some broken quartz. The structures had several components, and the aim of the excavation was to characterise them and to place them in chronological order. The site occupied the crest of a low mound bounded by a bank of rubble. Within its area was a circle of tall monoliths, with two ‘portal stones’, which marked the position of an entrance. There was an oval setting of standing stones inside it. Excavation supplied further details. By each of the stones at the entrance was an oval pit, interpreted as the position of a grave (human bones would not have survived on such an acid subsoil). The innermost circuit of monoliths had been bedded in a shallow ditch, which was associated with a ring of postholes. This feature contained pottery dated to the fourth millennium BC. Piggott and Simpson suggested that the monument had developed in three stages (see Figure 2.4). The earliest was the circle of upright posts, and possibly the ditch that followed the same outline. It had a hearth at its centre, associated with a deposit of comminuted charcoal and a large stone. This structure was followed by the oval setting of monoliths and the rubble bank that enclosed the site. The outer circle, the portal stones and the possible graves were assigned to a final phase and dated to the late third or early second millennium BC (the Early Bronze Age). Both excavators were aware that other readings of the evidence might be possible, for there were only two fixed points in this sequence. The sockets for the inner setting of monoliths had cut through the filling of the ditch. They followed the same footprint as the earlier timber circle, and sherds of Neolithic pottery were associated with that earthwork. When Piggott wrote about excavation in 1959, he was well aware of the factors that might influence its results. It is worth repeating a statement quoted earlier in this paper. ‘What the excavator extracts from his dig will depend on his past experi- ence [and] his comparative knowledge of the material he is unearthing’ (Piggott 1959: 14). These factors affected the outcome of the project at Croftmoraig. Both excavators had considerable experience of the archaeology of southern England. Stuart Piggott had worked on some of the more significant Neolithic and Early Bronze Age monuments and had studied the finds from others. Derek Simpson was co-author of a book cataloguing an important collection of artefacts 34 Richard Bradley

N

(a)

Enclosed Wall structure

Stone oval 3500BC? 2000BC?

Schiehallion

Midsummer

(b) sunset Erratic d

n u

o

m l Glacia

Enclosed St structure 1410–1220BC on e

c i r

Erratic c

l e

2000BC? 1370–1120BC Midwinter sunset

Current phase 020m Previous phase

FIGURE 2.4 Croftmoraig stone circle. (a) The structural sequence postulated by Piggott and Simpson in 1971. (b) The revised structural sequence established in 2012. (Drawing: Sarah Lucas; copyright: author.) from Bronze Age burials in Wessex (Annable and Simpson 1964). They had direct experience of working at stone circles in lowland Britain, and Piggott had also investigated the Clava Cairns – megalithic monuments enclosed by rings of stones (Piggott 1956) – and another Scottish site at Cairnpapple where there was evidence of a demolished circle (Piggott 1948). Much of this work took place at a time when models developed in the south were extended to prehistoric structures in the north. That gave rise to certain expectations. One was suggested by work in Wessex, and Piggott and Simpson said as much in their report (1971: 11–12). There, timber circles, like the Sanctuary which Repeating the unrepeatable experiment 35 forms part of the Avebury complex, were replaced by rings of monoliths. That pattern was reinforced by subsequent excavations. At the same time, it seemed likely that post circles would date from the Neolithic period and might be associated with the distinctive pottery of that phase. This knowledge influenced the interpretation of Croftmoraig. Analogy with structures in the south suggested that the ring of wooden uprights would have been the oldest building on the site (although the ditch associated with it was difficult to explain). The pottery from their excavation included a few pieces of Early Neolithic date, and so the other sherds were assigned to the same period. The previous careers of both excavators suggested an interpretation that was consistent with what had been found in southern England. After their report was published, the oval stone setting at Croftmoraig was described as one of the oldest monuments of its kind in Britain (Burl 2000: 249–251). It was those same factors – comparisons with other monuments and the experience of working on particular projects – that first raised doubts about this reading of the evidence. That only happened because of research that took place after the results of excavation at Croftmoraig were published. Two developments were particularly significant. The first was the direct dating of cremated bone from Scottish monuments. This revealed that a number of stone circles had been reused long after they were first constructed (Bradley 2011: 169–173). It also allowed the coarse pottery with which these remains were associated to be dated accurately for the first time. They were used between about 1500 and 800 BC. Such vessels bore a disquieting resemblance to nearly all the pots from Croftmoraig. Since they were associated with the post circle on that site, it seemed possible that the post circle was not as early as Piggott and Simpson supposed. Moreover, these artefacts provided a terminus post quem for the oval stone setting, suggesting that, far from being the oldest monument of its kind in Britain, it was one of the latest (Bradley and Sheridan 2005). The discrepancy could be as much as two thousand years. That severed any connection with the sequence in the south. Piggott had emphasised the importance of comparison with the results of other excavations. More recent work in Scotland suggests a different analogy for the timber circle at Croftmoraig. This is the kind of structure described as a ‘ring ditch house’. It shares the same components: a circle of posts with a discontinuous gully on the inside, and a pair of timbers projecting towards the south or east that could mark the position of a porch. No examples had been excavated and dated in 1965, but subsequent research has identified buildings of this kind at a series of settlements in Scotland, where they were first constructed around 1500 BC (Cook and Dunbar 2008: 317–333). That would agree with the revised dating of most of the pottery from Croftmoraig. Stuart Piggott’s unpublished plan of the excavation held by the National Monuments Record of Scotland reinforces the comparison, for it provides more detail on the wooden structure than the published report. Taken together, these observations suggested that the timber building and the stone setting that replaced it at Croftmoraig may have been the latest structures on the site, rather than the earliest, but this revision depends on the same principles 36 Richard Bradley that Piggott advocated in 1959. The only difference is that the range of possible comparisons widened after the project took place. The same method supports the original dating of the outer stone circle. The excavators argued that the possible graves in its entrance were like those of the late third or early second millennium BC. Subsequent work, some of it based on the project archive, has drawn attention to the way in which the monoliths at Croftmoraig increase in height from north-east to south-west. The same pattern occurs at stone circles of this period that have been investigated more recently. This new scheme was published in 2005, but in the absence of radiocarbon dates it remained hypothetical (Welfare 2011: 259–261). That was why a small excava- tion was undertaken on the site in 2012. Although its aim was to collect samples that were too small to be dated fifty years ago, the results of this work were surpris- ing, for they shed unexpected light on some of the unstated assumptions made by Piggott and Simpson. These had not been apparent from the existing record.

Croftmoraig stone circle: re-excavation in 2012 In contrast to South Lodge Camp, Croftmoraig had been well recorded by the previous excavators. The trenches were precisely where the excavators had marked them on their plan of the monument, and the new work could add nothing to their observations of the subsoil features. Where the new project did break new ground was by sectioning the perimeter bank, which had not been investigated before, and by focusing on the baulks in between the 1965 trenches where charcoal had been observed in the previous project. The bank proved to be the foundation of a wall, composed of large stone blocks. According to Piggott and Simpson it was contemporary with the levelling of the site when the oval stone setting was built. The two structures are concentric with one another. The outer boundary of the monument was constructed on a footing of massive rubble bedded in a shallow trench. Six samples of charcoal from short- lived species were collected from the base of this feature, and, at two sigma, their dates overlap between 1410 and 1220 BC. Two more dates are for samples from the filling of the ring ditch. They overlap in the period between 1370 and 1125 BC. They are consistent with those from the perimeter wall and also with earlier estimates based on the pottery found in 1965. It is clear that the revised sequence is substantially correct and that the timber building, the oval stone setting and the enclosure wall were the latest structures on the site (see Figure 2.4). There matters might have rested, but other issues arose in the course of the new work. They are discussed here because they were not considered during the previous excavation and could never have been anticipated from the contents of the published report. They share one characteristic: all of them relate to elements that would have been overlooked by the prehistorians of the 1960s because of their adherence to the distinction between nature and culture that has had so much influence on Western thought. Since then, the distinction has been questioned (for an archaeological example, see Thomas 2004; for an anthropological example, see Repeating the unrepeatable experiment 37

Descola 2013). Until that discussion took place, certain components of Croftmoraig were unlikely to receive the attention they deserved. These were: the mound or platform on which the circles had been built; the character of the central ‘hearth- stone’; and the positions of the successive structures in relation to the sky and the mountains around them. Although a recent account treated the mound as an artificial platform (Welfare 2011: 259–261), it is clear that the platform is a geological feature that was formed by melting ice. It is one of many in the area, but has the unusual characteristic of having been deliberately sculpted by human action. Instead of presenting a rounded profile, its top has been shaped so that it rises in an even gradient from east to west. It was towards the west that its edge was cut away and the most substantial section of wall was built. The grading of the mound reflects the increasing heights of the monoliths in the outer circle. At the same time, there is a conspicuous banded stone, 1.2 metres long, on top of this mound. This can be recognised as a glacial erratic. The upper surface of this stone would have been visible when the monument was first built, but the 1965 excavation showed that people had previously dug around it to uncover more of its extent. It is bright orange when the surface is moist and illuminated by the sun. The rock is a kind of schist that has not been recognised anywhere else in the vicinity. It is exceptionally hard, yet several flakes were detached from its surface in antiquity. Beside it there was evidence of a fire, but no trace of charcoal remained when this area was re-excavated in 2012. Piggott and Simpson concluded that this stone marked the position of a hearth in the middle of the timber circle. They say little about it, but the photographs in the site archive tell a different story, for it obviously puzzled them. They pictured the stone from various angles before it became apparent that it was embedded in the natural clay and formed part of the geology of the site. At that point they lost interest and it hardly features in their report. The presence of such an unusual rock in the centre of the mound takes on a new significance in relation to the siting of the monument. Quite by chance the view from these features extends for 11 kilometres, as far as Schiehallion, one of the most conspicuous summits in the southern Highlands (this effect is enhanced because the gradient was modified when the monument was built – see Figures 2.5a and 2.5b). Seen from Croftmoraig, the sun sets into the top of the mountain at midsummer (see Figure 2.6). This can be observed from the monument, but is even more apparent when viewed from the rising ground a short distance further to the east. The long axis of the oval stone setting may have a similar significance, as it is directed towards the setting sun at midwinter. At the time of the original excavation such observations were treated with greater caution than they are today, and there is nothing to suggest that these phenomena were noticed by archaeologists until recently. But they may explain both the siting and layout of the monument at Croftmoraig. Perhaps the place was chosen precisely because someone noticed the striking relationship between the location of a natural mound, the distinctive rock in its surface and the positions of the sun at the turning points of the year. This may have happened because the FIGURE 2.5 (a) (a) General view of Croftmoraig stone circle. (b) View over the stone circle towards the summit of Schiehallion. (Photographs: Aaron Watson and Richard Bradley; copyright: author.)

(b) Repeating the unrepeatable experiment 39

FIGURE 2.6 The midsummer sunset at Schiehallion viewed from the raised ground behind the stone circle. (Photograph: author.) erratic was the same colour as the sky over Schiehallion at the midsummer solstice. At this stage there was no monument, and these chance observations could have taken place long before any structures were built. When structures were built, each successive ring was laid out with the stone at its centre. It may be no coincidence that the tallest monoliths in the western sector of the circle contained natural inclusions that would glitter in sunlight. The same is true of the material in the perimeter wall. Taken together, these features created a spectacle that would have been regarded with caution in the 1960s. This caution reflects the separation the excavators maintained between nature and culture.

Conclusion This chapter began with a passage from an influential book on field archaeology, which contended that excavation represents an unrepeatable experiment. Both the cases considered here suggest a different view. Of course some ancient monuments are entirely destroyed, especially when they are threatened by commercial develop- ment, but it is an illusion to suppose that a single excavation, however accom- plished, can encompass as much as Philip Barker expected. Nor can any field project – even the best – be entirely free of the ideas current at its time. It is ironic that Pitt-Rivers’s excavation at South Lodge Camp has been celebrated as a model of objective recording when the methods that he used there 40 Richard Bradley were the product of his individual understanding of evolutionary theory (Bowden 1991). His supposedly comprehensive account of this important monument did less than one might have expected to inform a new interpretation of the site. By contrast, the original excavation of Croftmoraig stone circle followed a more traditional agenda. It was conducted to a very high standard, and it is only possible to suggest a new structural sequence because of the quality of the work done there in 1965. The new interpretation follows the principles that Piggott had enunciated some years before, and the revised chronology depends on advances in pottery studies and radiocarbon dating that have happened since. These developments could have been predicted, for in their excavation report Piggott and Simpson conceded that other interpretations of the evidence might be valid. What they could not have foreseen was a feature of the monument that prehistorians in the 1960s would have found difficult to acknowledge; that is, the interplay between the natural and the cultural – in this case, between geology and architecture – that has become an important issue in more recent years (Descola 2013). It provides the vital clue to the interpretation of this remarkable site, and yet it was overlooked. What better demonstration could there be that every excavation is a product of its time and that Barker’s ‘unrepeatable experiment’ must sometimes be repeated?

References

Annable, K. and Simpson, D. (1964) Guide Catalogue of the Neolithic and Bronze Age Collection in Devizes Museum, Devizes: Wiltshire Archaeological and Natural History Society. Barker, P. (1977) Techniques of Archaeological Excavation, London: Batsford. Barrett, J., Bradley, R. and Green, M. (1991) Landscape, Monuments and Society: The Prehistory of Cranborne Chase, Cambridge: Cambridge University Press. Bowden, M. (1991) Pitt-Rivers, Cambridge: Cambridge University Press. Bradley, R. (1983) ‘Archaeology, evolution and the public good: the intellectual development of General Pitt-Rivers’, Archaeological Journal, 140: 1–9. Bradley, R. (2004) ‘Stuart Piggott, Ancient Europe and the prehistory of Northern Britain’, in G. Barclay and I. Shepherd (eds) Scotland in Ancient Europe, Edinburgh: Society of Antiquaries of Scotland, 3–11. Bradley, R. (2006) ‘Bridging the two cultures. Commercial archaeology and the study of prehistoric Britain’, Antiquaries Journal, 86: 1–13. Bradley, R. (2011) Stages and Screens: An Investigation of Four Henge Monuments in Northern and North-Eastern Scotland, Edinburgh: Society of Antiquaries of Scotland. Bradley, R. (in prep) The Use and Reuse of Stone Circles: Fieldwork at Five Scottish Monuments, Oxford: Oxbow. Bradley, R. and Sheridan, A. (2005) ‘Croft Moraig and the chronology of stone circles’, Proceedings of the Prehistoric Society, 71: 269–281. Burl, A. (2000) The Stone Circles of Britain, Ireland and Brittany, New Haven, CT: Yale University Press. Carver, M. (2009) Archaeological Investigation, London: Routledge. Collingwood, R. (1939) An Autobiography, Oxford: Oxford University Press. Collingwood, R. (1946) The Idea of History, Oxford: Clarendon Press. Repeating the unrepeatable experiment 41

Cook, M. and Dunbar, L. (2008) Rituals, Roundhouses and Romans: Excavations at Kintore, Aberdeenshire, 2000–2006, Edinburgh: Scottish Trust for Archaeological Research. Descola, P. (2013) Beyond Nature and Culture, Chicago: Chicago University Press. Hawkes, C. and Piggott, S. (1947) ‘Britons, Romans and Saxons round Salisbury and in Cranborne Chase’, Archaeological Journal, 104: 27–81. Megaw, J. V. S. and Simpson, D. (1979) Introduction to British Prehistory, Leicester: Leicester University Press. Piggott, S. (1948) ‘The excavations at Cairnpapple Hill, West Lothian, 1947–1948’, Proceedings of the Society of Antiquaries of Scotland, 82: 68–123. Piggott, S. (1949) British Prehistory. Oxford: Oxford University Press. Piggott, S. (1956) ‘Excavations on passage graves and ring cairns of the Clava group, 1952–1953’, Proceedings of the Society of Antiquaries of Scotland, 88: 173–207. Piggott, S. (1959) Approach to Archaeology, Harmondsworth: Penguin. Piggott, S. (1965) Ancient Europe, Edinburgh: Edinburgh University Press. Piggott, S. (1973) ‘The final phase of bronze technology’, in E. Crittall (ed.) A History of Wiltshire, Vol. 1, part 1, Victoria County Histories, Oxford: Institute for Historical Research, 376–407. Piggott, S. and Simpson, D. (1971) ‘The excavation of a stone circle at Croft Moraig, Perthshire, Scotland’, Proceedings of the Prehistoric Society, 37(1): 1–15. Pitt-Rivers, A. (1887–1898) Excavations in Cranborne Chase, 4 vols, privately published. Pitt-Rivers, A. (1906) The Evolution of Culture and Other Essays, Oxford: Clarendon Press. Pitt-Rivers, A. (n.d.) A Short Guide to the Larmer Grounds, Rushmore, King John’s House and the Museum at Farnham, privately printed. Thomas, J. (2004) Archaeology and Modernity, London: Routledge. Thompson, M. (1977) General Pitt-Rivers: Evolution and Archaeology in the Nineteenth Century, Bradford-on-Avon: Moonraker Press. Toms, H. (1925) ‘Bronze Age or Earlier Lynchets’, Proceedings of the Dorset Natural History and Archaeological Society, 46: 88–100. Welfare, A. (2011) Great Crowns of Stone: The Recumbent Stone Circles of Scotland, Edinburgh: Royal Commission on the Ancient and Historical Monuments of Scotland. Wheeler, M. (1954) Archaeology from the Earth, Oxford: Clarendon Press. 3 EXPERIMENTAL ARCHAEOLOGY AT THE CROSSROADS

A contribution to interpretation or evidence of ‘xeroxing’?

Martin Bell

Experimental archaeology has featured very little in recent writings about archaeo- logical theory (Johnson 2010), even those concerned with the role of science in archaeology (Jones 2002). Experimental approaches enjoyed a higher profile in the 1960s and 1970s, when archaeologists were increasingly turning to scientific methods. There followed a period in the 1980s and 1990s when post-processual approaches were in the ascendancy and experimental approaches were marginalised from the academic mainstream (Millson 2011), a position from which experimental archaeology is only now struggling to recover. The attenuated timescale of recovery is striking, given that, from the late 1990s, there has been growing recognition that a diversity of approaches, both scientific and social, can and must contribute to archaeological interpretation (Hodder 1999). In a stimulating essay, Koerner (2011) has recently considered the role of theory in experimental archaeology, although without practical examples to signpost the way forward. In the same volume, Millson (2011) argued that experimental archaeology can play a bridging role between data and theory, science and arts. Informative as a range of social theories and philosophical perspectives derived from the humanities have proved to be, we can be no more confident that they provide a reliable guide to interpretation than analogical reasoning (interpretation by comparison), offered, for instance, by ethnographic and experimental evidence. The argument here is that we need to utilise the full toolkit of available approaches, social and scientific, including experiments, and to devise effective combinations that enable an interpretation derived from one source to be critically examined against evidence from other sources. Experimental approaches are most straightforward and are most adequately reflected in the literature, where they concern how artefacts were made and used (Coles 1973, 1979; Ferguson 2010). The work of van Gijn (1990) on the experimental investigation of wear traces on flint is particularly exemplary in this Experimental archaeology at the crossroads: interpretation or ‘xeroxing’? 43 regard. Other examples are given in recent edited volumes (Mathieu 2002; Cunningham et al. 2008; Millson 2011). Such artefact experiments are often rela- tively short term and may be carried out as part of wider schemes of investigation, often employing a range of scientific techniques. The emphasis of this chapter is rather different; it concerns particularly longer-term experiments. Many of these take place in open-air museums, where the primary role is education: sites, as Paardekooper (2013) says, that are ‘located somewhere between archaeological science and the public’.

Experiments: some key concepts and issues As a first step, some key concepts are noted and critiqued; these are italicised for emphasis because some feature rather less in the somewhat atheoretical experi- mental literature than it is argued they should. Traditionally, experiments involve defined steps. They begin with observation, highlighting a strong empirical basis; then a hypothesis is formulated to account for the observations; next, an experi- mental methodology is identified to test the hypothesis; and, finally, the resulting observations, measurements and tests produce data that provides a basis to refute, or confirm, the original hypothesis. The result may be further tested by replicated experiments. Experimental archaeology rests on analogy between an experimental and a past context. Analogy is unavoidable in archaeology (Johnson 2010) and may derive from various sources: experiments, environmental processes, anthropology, etc. No analogy will be precise and the extent to which it is appropriate demands critical appraisal (Wylie 2002: Chapter 9). A distinction can be made between formal analogy, based on comparisons, and relational analogy based on causal relationships (Hodder 1999), the latter being perhaps less problematic and more amenable to experimental investigation. Formal analogy, involving cross-cultural and cross- temporal comparisons, will generally be more problematic. Similarly central is the concept of uniformitarianism – the present as the key to the past – originally derived from geology. At one level, uniformitarianism is a fact of life: how else can we know the past, except through our lived experience of the present? We have no other starting point. The concept is inherent to any attempt to create experimental conditions today that can be used as an analogy for past. It should be emphasised that the highly literal uniformitarianism of the nineteenth century, with an emphasis on gradualism, has been tempered by the development of methodological uniformitarianism, with a more subtle appreciation of timescales of change, the role of contingency, perturbations and extreme events (Bintliff 1999; Bell and Walker 2005: 21). The setting up of conditions for hypothesis testing and a clear methodology is basic to experimentation. It is equally fundamental to most applications for archaeological funding, whether or not they are experimentally based. The philosophical conun- drum is that the gathering of data is not independent of the theory to be tested. Part of the problem concerns proof, which is of a different order in the context of 44 Martin Bell sciences such as physics and chemistry as compared to historical sciences such as archaeology, where we are attempting to work out what happened in the past and are seldom in a position to prove anything conclusively. However, an experiment or analysis may establish that we can accept a particular hypothesis not as proven but accepted pending further critical enquiry, as Adams (1988) cogently posits. Thus in the historical sciences experiment can be seen not as establishing irrefutable facts but as part of a critical and questioning dialogue in the relentless search for a better interpretation. Binford (1977) advocated hypothesis testing using middle-range theory, grounded in the modern world (e.g. through ethnography and experiments), to construct bridging arguments translating the largely static archaeological record into the dynamics (time-related changes) required of accurate statements about the past. Binford, from the North American tradition, drew his middle-range analogues largely from ethnographic observation. Middle-range theories, he argued, needed to be ‘devised and tested in intellectual contexts divorced from the theories about past behaviour which we seek to evaluate’ (Binford 1983: 213). In practice it is often difficult to identify analogues that are fully independent of, for instance, uniformitarian assumptions about human behaviour. Bogaard (2004 and Chapter 13 of this volume) provides an exemplary case in which theories of Neolithic shifting cultivation were tested by evidence provided by the functional attributes of plant communities recorded from agricultural field experiments. Formation processes require particular attention. Archaeology is neither philosophy nor linguistic theory, nor any of the other myriad disciplines that contribute stimulating ideas to help us interpret the past. It is the study of the past using material remains. The latter here is broadly defined to include the material interactions between people and the world of which they form a part: plants, animals, rocks, etc. Formation processes should be as central to the study of archaeology as source criticism is to the classicist or historian. We need to give focused attention to the distinctive characteristics of our main source material, how material culture enters the archaeological record and the effects of time and recovery on that record. Schiffer’s work on formation processes (1987) and behavioural archaeology (2010), grounded in the North American tradition, is mainly ethnographically based. His identification of cultural (c) and natural (n) transforms has enhanced understanding of formation processes and is particularly amenable to experimental investigation. The role of faunal agents is further discussed below. Experiments will be more robust, and more amenable to effective critique by others, when the research design and methodology are clear and the results made available by publication, or accessible archiving. These steps are obvious, but are not always followed on sites that are presented as experimental. A basic tenet of an experiment is that it should be capable of replication (Reynolds 1999). In practice, however, contingent (chance) factors mean that true replication is rare in archaeology (Hodder 1999). Hurcombe (2008) has argued that exploration of the variable ways in which things can be done is of greater Experimental archaeology at the crossroads: interpretation or ‘xeroxing’? 45 value than repetition. Examples are outlined below relating to roundhouses and earthworks, where it has proved valuable to have closely related experiments that are not strict replicates. Experimentation, as well as much of field archaeology, is essentially an empirical process, that is, it is based on observation and recording. This assumes that a degree of objectivity can be achieved, and that the observations are not totally obscured by underlying assumptions based on pre-understandings. Experience shows that some individuals and projects are much better at making observations independent of their particular theories than others. If that spectrum between subjectivity and objectivity can be recognised, then it should be possible to aspire to the more objective end of the spectrum. The philosophically problematic nature of objec- tivity has arguably led archaeologists to put insufficient energy into honing their empirical skills, which seem to come more naturally to the geologist or natural historian. Experiments are an especially effective way of honing these skills because they represent a physical engagement with how things were made and done and also how the archaeological record forms. Doubts about objectivity and the lack of emphasis on empirical skills may have exacerbated a tendency in archaeology to base interpretation less on the detailed observations made at a specific site but rather on comparisons with what others have found at similar sites, a tendency caricatured here as xeroxing. Examples from experimental archaeology are noted below.

Integration of science and social perspectives; multivocality, reflexivity and triangulation The scientific approach based on observation, analogy, experiment and the testing of theories has to confront some significant challenges. These have been debated, and arguably resolved to some degree, by recent publications, which, it is con- tended, offer multiple slightly different versions of the same solution. Put simply, these involve the rigorous critical examination of interpretations derived from one source of evidence against other sources of evidence of varying independence. Much of the debate is in part about the meaning, or possibility, of objectivity. Hodder (1999) emphasises that data is not independent of theory but related to it; pre-understandings will influence but not determine interpretation. Influence implies variability from case to case, and this may be accessible to evaluation by the researcher, or in debate with others. As Johnson (2010) contends, we can never prove that underlying uniformitarian (or other) analogy is valid. However, that statement sets the bar unrealistically high because, as noted above, in the historical sciences we are seldom in a position to prove anything. Increasingly there is recognition that the solution to this conundrum is to test the extent to which the analogy is valid against evidence from a range of other sources. The challenge then becomes establishing to what extent the evidence gathered can be independent of the ideas being tested, or has been influenced by pre-understandings derived from those ideas. Various related approaches to independent testing have been developed. Binford’s concept of middle-range 46 Martin Bell theory has already been noted. Another is the multi-proxy approach of the environmental archaeologist, whereby environmental interpretation is based on comparison of as wide a range of sources as possible (Bell and Walker 2005), an especially successful strategy in wetland contexts, where many sources may have been preserved. The evaluation of multiple working hypotheses is another related approach that ensures that the widest possible range of interpretations is considered, often alongside a multi-proxy investigation. In Jones’s (2002) view, the fragmentation of evidence among a range of specialists contributing to an archaeological project is problematic and leads to decontextualisa- tion. A converse proposition is posited here: greater independence of pre- understandings is achieved where a range of specialists from different disciplines and backgrounds contribute to interpretative debate. Related approaches in anthro- pology are interpretations based on cables, comprising multiple strands of evidence, and tacking, a dialectical process between the contrasting perspectives of the anthropologist and the subject (Wylie 2002: Chapter 11). This is analogous to the process that Wylie (2002) has called ‘triangulation’, whereby an interpretation that may not be entirely convincingly demonstrated from any one source becomes more precisely grounded by triangulation with other sources of varying independence. Hodder (1999) introduced two concepts that strengthen this approach by challenging tendencies to scientific dogmatism and the arrogance of the ‘expert view’, which can make scientific approaches problematic. The first of these is multivocality, which values the contributions of diverse voices and approaches, both scientific and social. The second is self-reflexivity, a critical awareness of the effects of scientific and archaeological assumptions on the data obtained and the effects of these on the communities involved. Hodder has demonstrated the value and practical achievements of these concepts through the project at Çatalhöyük, Turkey, which includes a small experimental area serving both heritage outreach and scientific experimental objectives. Koerner (2011) contends that experimental archaeology can have a role in overcoming dichotomies such as objective– subjective and culture–nature. She argues for a reflexive experimental archaeology, ‘to sustain and enhance plurality of lived cultural heritage’ (Koerner 2011: 85). The approach is exemplified by the work of Croes on many aspects of waterlogged sites on the American north-west coast, and especially in his work on basketry, which involves complementary analysis of the organic archaeological remains, experi- mental replication and the traditional knowledge of First Nations communities, all contributing to an appreciation of cultural identity, achievement and the value of diversity (Croes 1988; Croes et al. 2009). In many ways, the approaches advocated reflect the ways in which archaeologists actually work in practical terms because, at least in the more complex projects, research is generally a collaborative activity, drawing on a range of specialist knowledge and contrasting perspectives in order to arrive at an interpretation. This is not, however, an argument for business as usual. It must be acknowledged that science does not always live up to its ideals of being critical and questioning, rather than dogmatic and uncritically accepting of the expert view. Greater reflection on, Experimental archaeology at the crossroads: interpretation or ‘xeroxing’? 47 and acknowledgement of, the influence of pre-understandings and the process of interpretation is required. These points are especially pertinent to experimental archaeology, which has tended to focus more on presenting ideas about the past than on active debate about interpretation.

Experiments with earthworks Particularly notable is a series of experiments on round barrows at Lejre, Denmark, between 1995 and 2004. The experiments addressed the problem of how exceptional preservation conditions developed in Bronze Age round barrows, where, on sandy, well-drained soils, oak coffins and organic artefacts including clothing were preserved, most famously at Egtved. The hypothesis was that preservation was related to iron pan-like deposits observed within the barrows. The approach adopted was twofold, involving: first, the excavation of a barrow and the investigation of the iron deposits; second, the creation of 1:5 scale models of barrows in the experimental centre at Lejre. In these, dead piglets were substituted for human bodies in wood coffins, which also contained textiles and other artefacts. Conditions in the barrows were monitored and excavations took place three or four years after construction. It was shown that reducing conditions rapidly developed within the mound after burial, and a redox pan formed an envelope round the core of the barrow within which anaerobic conditions obtained (Breuning-Madsen et al. 2001; Breuning-Madsen and Holst 1998). There are hints that Bronze Age communities may have deliberately induced appropriate condi- tions, and it is striking that dendrochronology shows that the burials occur in a tight chronological band in the fourteenth and beginning of the thirteenth century BC, making chance causation unlikely (Holst et al. 2001). The experiment is exemplary in its focus on significant cultural archaeological questions and its demonstration that an experiment lasting three to four years can contribute to the understanding of preservation over millennia. The Experimental Earthwork Project has a much longer timescale. Started in 1960, it was based on two sites with contrasting geology: Overton, Wiltshire, on chalk (Figure 3.1), and Wareham, Dorset, on sand. The experiment was concerned with change over the highly ambitious timescale of 128 years. There was a clear research design to investigate the ways in which the earthwork weathered, became vegetated and the decay and preservation of buried materials within replicate sections excavated at intervals (Jewell 1963). Results from the first thirty-two years demon- strate the contribution of short and medium-term experiments to the understanding of long-term formation processes. They show the value of replicates in time series as well as the role of contingent factors and the value of a multi-proxy approach involving a team from diverse disciplinary backgrounds (Bell et al. 1996). As the experiment developed it proved possible to incorporate new research objectives and techniques not foreseen in the original research design. The problems of maintaining momentum in a long-term experiment must also be acknowledged. From 1985, a further octagonal earthwork experiment was established by Peter Reynolds: this had 48 Martin Bell

FIGURE 3.1 The experimental earthwork at Overton, Wiltshire, excavated 1992. (Photograph: A. Sacre; copyright: Experimental Earthwork Project.) a decadal timescale and more restricted objectives concerning weathering and vegetation colonisation through time (Reynolds and Hedge 1999). Four earthworks were made, one on brickearth at Fishbourne and three on chalk: one at Wroughton and two on successive Butser sites (Reynolds 1996). Few results from the octagonal earthworks had been published at the time of Reynolds’s death in 2001, but there has been some subsequent work on the buried soils beneath them that was not planned in Reynolds’s original research design (Bell 2009).

Experiments with buildings A long-term Lejre experiment concerns the effects of fire on the destruction of buildings, such as the Iron Age byre-houses, of which there are many burnt examples in Denmark. A rectangular byre-house was built when Lejre was established in 1965, and was deliberately burnt in 1967 followed by keyhole excavation. Twenty-five years later, as originally planned, a full excavation of half the building took place, followed by a publication exemplary in that it discusses the original experimental construct, its later excavation and a reflective analysis concerning interpretation of the Iron Age burnt structures (Rasmussen 2007). A related example at West Stow, Suffolk concerns Anglo-Saxon sunken feature buildings, the question of whether they had suspended floors and the interpretation of microtraces of activity within them. A building constructed between 1992 and 1998 was burnt down by arson in 2005 and excavated. Post-excavation analysis here was notably multi-proxy, including a range of scientific analyses (micro- morphology, magnetic properties, botany, insects and fire forensics, etc.). The results are presented in the context of wider discussion of many archaeological examples of the building type (Tipper 2012). Experimental archaeology at the crossroads: interpretation or ‘xeroxing’? 49

Most prolific of the experimental contexts in the UK are roundhouses (Figure 3.2), largely stimulated by P. J. Reynolds’s work at Butser and elsewhere. In the early years at Butser from 1972 onwards Reynolds experimented with a range of methods, which led to some failures, and the frank acknowledgement of these are testimony to his learning curve. The Balkesbury house failed after a year in 1975, the Maiden Castle house after ten years and a Danebury house proved incapable of being roofed (Reynolds 1979, 2006). The Butser constructs also provided possible explanations for several previously puzzling features encountered in the original excavations (Harding 2009; Harding et al. 1993). By 1977, with the con- struction of the Pimperne house, Reynolds had established key structural principles. This is the most exemplary of his constructs, with a publication that covers the original excavation by Harding and Blake in 1960–1963, the construct at Butser in 1976–1990 and the results of limited investigation at the time of demolition when the farm moved, in 1990 (Harding et al. 1993). There is a clear progression in Reynolds’s published roundhouse work, from concern with structural principles to understanding the archaeological record, as exemplified by his paper ‘Life and Death of a Posthole’ (Reynolds 1995), and, near the end of his life, enlistment of scientists to examine microtraces of activities. This last is demonstrated by investigation of the Moel-y-Gaer house, in which cattle and sheep had been overwintered. It was burnt down and subsequently subject to limited excavation and micromorphological research when Butser moved site (Goldberg and Macphail 2006; Macphail et al. 2004). Work on microtraces has continued in a limited way since Reynolds’s death (Bell 2009; Banerjea et al. 2013). This has included the excavation and micro-analytical investigation of the Moel-y-Gaer house that Reynolds constructed at St Fagans (Figure 3.3). Artefact distributions on the floor show marked concentrations around the wall line and hearth, and contrasts between artefacts derived from the dressed building and

FIGURE 3.2 The Moel-y-Gerddi roundhouse at Butser. (Photograph: author.) 50 Martin Bell

FIGURE 3.3 Excavation of the Moel-y-Gaer roundhouse at St Fagans, 2009. (Photograph: author.) re-enactment activities and those left by visitors mainly behind benches either side of the door (Figure 3.4). Such artefact patterns might contribute to the identifi- cation of furniture, subdivisions or activity areas within buildings, especially when accompanied by a range of micro-analytical techniques (Banerjea et al. 2013). Artefact patterns likewise proved illuminating in the previously noted excavations of the constructs at West Stow (Tipper 2012) and Lejre (Rasmussen 2007). Reynolds’s roundhouse constructs have been highly influential – in a way, rather too much so, because many examples made by others fall into the ‘xeroxing’ category, whereby one experiment reproduces rather than tests the results of another. In Britain, sixty-one experimental roundhouses have been constructed on thirty-two different sites, the vast majority based on the principles developed by Reynolds at Butser. This is hinted at if we look at the distribution of the constructs in terms of the archaeological period they are designed to represent: Mesolithic 1 (2 per cent); Bronze Age 5 (8 per cent); Iron Age 50 (82 per cent); Romano-British 1 (2 per cent); no defined period 4 (7 per cent). There are very large numbers of excavated Bronze Age roundhouses and a body of literature on their construction and the social use of space within them; however, archaeologists have not thought it worthwhile to investigate these structures experimentally on any scale. The notable exceptions are at Flag Fen (two houses), Millfield (Northumberland) and Saverock (Cornwall), but none of these has been the subject of any published experimental work. The period distribution of roundhouses can be largely attri- Experimental archaeology at the crossroads: interpretation or ‘xeroxing’? 51

N

Visitor artefacts Re-enactment artefacts

FIGURE 3.4 Plan of artefact distributions in the Moel-y-Gaer roundhouse, St Fagans, distinguishing those relating to re-enactment activities and visitors. (Graphic: J. Foster; copyright: author.) buted to the iconic significance of the Iron Age structures Reynolds created and the fact that others have reproduced rather than innovated. Some thirty of the sixty- one roundhouses were based on defined archaeological structures and are represented by some sort of publication, but in many cases that is very brief. Of those thirty, Reynolds himself was responsible for seventeen roundhouses (57 per cent), again highlighting the extent of his influence. As Harding (2009) notes, very few of the roundhouses are based on adequate excavated evidence. They have been neither described nor formally published, and their life is not documented, with the result that we cannot take advantage of the opportunity they should provide for testing the Butser results. Townend (2007), in discussing the rebuilding of the chieftain’s house at Castell Henllys, Wales in 1991–1992, critiqued Reynolds’s work, particularly what he saw as inherent notions of optimization and efficiency. He argued that there should be greater engagement with the act of construction and the human element, which Reynolds (1994) sought to exclude. However, 52 Martin Bell

Townend did not make a convincing case for the achievements, or insights, provided by the approach he advocates. One question is how long roundhouses lasted. Reynolds (1994) argued that with maintenance they could have a life of two centuries or more, with obvious implications for the inference of site density, settlement continuity and population levels. However, such longevity has not been borne out by subsequent results: Moel-y-Gaer house at St Fagans, 1992–2008, sixteen years; Moel-y-Gerddi house at St Fagans, 1992–2012, twenty years; Longbridge Deverill house at Butser, 1992–2007, fifteen years; Glastonbury House xiii, Somerset Levels and Moors, 1992–2008, sixteen years. One must acknowledge, of course, that, as always, the analogies are not perfect. Some buildings, such as Longbridge Deverill, were not built of the optimal wood type, and there are likely to have been fewer regular fires and less day-to-day maintenance than in occupied buildings. A number of writers have observed that excavated Iron Age roundhouses generally face south and east, and some exhibit distinct patterns of artefact distri- bution on their floors. This has led to the theory, partly based on structuralist analogies with ethnographic examples, that the layout follows cosmological principles related to the passage of the sun, the way it ordered daily life (Parker Pearson 1996) and perhaps sun worship. The model became widely accepted and applied to work on Iron Age Britain. Pope (2007) analysed the evidence in detail with a much larger sample of roundhouses, which did not suggest a pattern anything like as regular and convincing as previously claimed. Pope moved the emphasis away from cosmology and towards the optimisation of access to light, drawing on some limited observations from experimental structures. What is striking is that the many reconstructed roundhouses have played so limited a role in delimiting the effects of pre-understanding and testing the cosmological model. One problem is that many of the constructs are not orientated as the originals were. When a construct is made, the research questions to which it may become relevant cannot necessarily be foreseen; all the more reason, then, for sticking to one of Reynolds’s (1994) basic maxims – that experiments should be based closely on specific archaeological examples.

Agents: faunal and human Our tendency to assume that what we dig up is the way it was when left by past communities fails to take account of the interactive relationships between human and animal agents in structuring the archaeological record, as experiments can demonstrate. Darwin (1881) investigated how earthworms bury archaeological sites, and the long-term experiments that he set up on his lawn at Down House, Kent were a pivotal influence on the Experimental Earthworks Project ( Jewell 1963). That influence came full circle with the demonstration of the role of earthworms in reworking the buried soil below the Overton Down Experimental Earthwork (Bell et al. 1996). Earthworms are a particularly significant factor on calcareous soils. The effects of other faunal agents are now being documented Experimental archaeology at the crossroads: interpretation or ‘xeroxing’? 53 on different geologies. On the acid heathland soils of the Wareham Experimental Earthwork, biological activity within the buried soil is minimal (Bell et al. 1996). After thirty-two years, the crest of the bank had not been fully colonised by vegetation, leaving a sandy strip along its crest. The original hypothesis was that this was a result of erosion by rain. Careful observation showed otherwise: a key factor in maintaining bare sand was burrowing activity by sand lizards (Lacerta agilis) as they laid their eggs, many of which were encountered during the excavation; since this is a protected species these eggs had to be taken away for incubation. Another faunal factor was represented by the footprints of deer, which had exploited the earthwork as a place for observation and perhaps display, since it commands an extensive vista across heathland to the south. The experimental earthwork has thus created a suitable context for two very specific patterns of animal behaviour, which have themselves contributed to site formation processes. It may not be unreasonable to speculate that concentrations of breeding lizards or displaying deer contributed to the ways in which prominent barrows on heathland were perceived, and to the particular significance attached to them in prehistory. A further example of the role of faunal agents emerged during this writer’s 2009 excavation of the Moel-y-Gaer house, originally constructed by Peter Reynolds at the National History Museum, St Fagans, Wales in 1992 (see Figures 3.3 and 3.4). That site is on glacial till and it is unusual among the experimental structures in being built in woodland and in a somewhat damp situation. During the excavation it became clear that there was a colony of wood ants round the wall line, occupying the interstices at the base of the wattle-and-daub wall and the holes where most of the wall posts had rotted below ground level. The interstices of the wall base and the post pipes left by rotted posts were marked by a concentration of cereal husk fragments fallen from the roof and carried there by ants (Figure 3.5). Small mammals probably also contributed. When the Pimperne house at Butser was demolished, Reynolds (1994, 1995) similarly found that the post bases had rotted away and there was a gully below the wall line created by the activities of rodents (mice, rats and voles), which had removed the evidence for the original wall stakes in places. The channel formed might have been mistaken for a drip gully round the building. The post voids were contemporary with the life of the building and contained artefacts associated with its use, rather than, as normally assumed, with post-building sedimentation (Reynolds 1995). Artefacts associated with building use were similarly found in post voids and the vegetable frass in the St Fagans roundhouse. The role of animals in affecting the archaeological record has been recognised since Darwin (1881), and noted in subsequent work by Schiffer (1987), Wood and Johnson (1978) and Johnson (2002). Not previously highlighted is the potential for constructs made for mainly educational purposes to contribute to understanding of both faunal agency and the reciprocal relationships between human and faunal agents, as illustrated by the Wareham example. (a) FIGURE 3.5 (a) The wall line of the Moel-y-Gaer roundhouse, showing plant material collected by ants and small mammals. (b) The wattle-and-daub wall of the Moel-y-Gaer roundhouse with a wood stake decayed in the ground and replaced by humus. (Photographs: author.)

(b) Experimental archaeology at the crossroads: interpretation or ‘xeroxing’? 55

Conclusions Archaeologists ought not to find it difficult to escape from the tyranny of back- ground knowledge and pre-understandings. The discipline is, after all, refreshed by a constant stream of new discoveries. Yet there is still a tendency to interpret what we have found in terms of what others have found, rather than on the basis of detailed, empirically based work. Analogy drawn from ethnography and experi- ment is one of the ways in which our frame of reference and pool of ideas can be expanded. Experiments enable us to test interpretations and evaluate or limit the influence of pre-understandings. In the North American tradition, ethnography has been more influential than experiment – for instance, in Binford’s (1977) middle-range theory and Schiffer’s (1987, 2010) work on formation processes and behavioural archaeology. However, it should be noted that European experimental archaeology owes much to the Scandinavian ethnographic tradition, characterised by the desire to preserve artefacts and the structural evidence of regional diversity. This has led to the foundation of iconic open-air museums, including: Skansen, Sweden in 1891; the Norsk Folkemuseum, Norway in 1894; St Fagans, Wales in 1948; and Lejre, Denmark in 1965, the last two with experimental archaeological activity. Traditional agricultural methods in Finland also inspired the slash-and- burn agricultural experiments at Draved, Denmark (Steensberg 1979), based on what were inferred to be aboriginal European agricultural practices. Furthermore, several of the experiments foregrounded in this chapter are in some ways ethnographic, being sites designed and used mainly for educational purposes, e.g. the Lejre byre-house, the West Stow farmer’s house and the roundhouse experi- ments. These were not precisely defined, fully controlled experiments with clear aims and objectives set down from the outset – that would require a level of control seldom feasible in medium to long-term experiments. Other examples, such as the Danish Lejre experimental barrows or the British experimental earthworks, are more like traditional experiments, although even here the role of chance or contingency becomes apparent. An attempt has been made to demonstrate that more tightly controlled hypothesis-driven experiments, as well as empirically based work on sites with an educational focus, both have contributions to make to understanding how the archaeological record has been formed, testing interpretations and providing new ideas derived from careful observation. The open-air museums, which are more educationally focused, with a diverse visitor base, have the potential to play a wider, more active role, not just presenting ideas about the past but challenging interpretations by providing a physical arena in which debates about interpretation can occur. An objective should be to bring experimenter and visitor together in a shared learning experience, in which both have insights to provide. Hansen (2007), in the context of bronze casting, highlights the role of both academic and practical knowledge, the value of practice-based research and the concept of thinking in action. The particular insights gained or reinforced in the bronze-casting case concerned the furnace as a social catalyst and the way in which casting became a 56 Martin Bell performance. An inherent strength of archaeology is that it draws on expertise from a wide range of sources, so that people from diverse backgrounds have a valuable perspective to contribute. Indeed, with anthropology, a pivotal role of the subject is exploring the diversity of human existence. Such an approach is also central to the multivocal and self-reflexive approach advocated by Hodder (1999) and the enhanced opportunities that diversity of backgrounds, approaches and pre-under- standings create to triangulate and test particular interpretations of the past using a range of methods, including experiments.

Acknowledgements My debt to the pioneering work of the late Dr Peter Reynolds will, I hope, be obvious. Subsequent work at Butser has been facilitated by Christine Shaw, Simon Jay and Maureen Page. Postgraduate students have contributed to scientific aspects of experiments there, especially Dr Rowena Banerjea, Dr Chris Speed and Dr Amy Poole. I am likewise grateful for collaboration with all those I have worked with on the Experimental Earthwork Project. The National Museum of Wales is thanked for making possible work at the National History Museum St Fagans. The University of Reading has contributed to this work through the project ‘Developing Experimental Approaches in Archaeology’.

References Adams, R. M. (1988) ‘Introductory remarks: spatial and temporal contexts’, in J. Bintliff, D. A. Davidson and E. G. Grant (eds) Conceptual Issues in Environmental Archaeology, Edinburgh: Edinburgh University Press, 1–15. Banerjea, R., Bell, M., Matthews, W. and Brown, A. (2013) ‘Applications of micro- morphology to understanding activity areas and site formation processes in experimental hut floors’, Archaeological and Anthropological Science. Available at: SpringerLink 1866-9565 [6 December 2013]. Bell, M. (2009) ‘Experimental archaeology: changing science agendas and perceptual perspectives’, in M. J. Allen, N. Sharples and T. O’Connor (eds) Landscape and People, Oxford: Oxbow/Prehistoric Society, 31–45. Bell, M. and Walker, M. J. C. (2005) Late Quaternary Environmental Change: Physical and Human Perspectives, 2nd edn, Harlow: Longmans. Bell, M., Fowler, P. J. and Hillson, S. W. (eds) (1996) The Experimental Earthwork Project 1960–92, Research Report 100, York: Council for British Archaeology. Binford, L. (1977) For Theory Building in Archaeology, New York: Academic Press. Binford, L. (1983) In Pursuit of the Past, London: Thames and Hudson. Bintliff, J. (1999) Structure and Contingency, London: Leicester University Press. Bogaard, A. (2004) Neolithic Farming in Central Europe, London: Routledge. Breuning-Madsen, H. and Holst, M. K. (1998) ‘Recent studies on the formation of iron pans around the oak coffins of the Bronze Age barrows of Denmark’, Journal of Archaeological Science, 25: 1103–1110. Breuning-Madsen, H., Holst, M. K. and Rasmussen, M. (2001) ‘The chemical environment in a barrow shortly after construction – an archaeological pedagogical experiment’, Journal of Archaeological Science, 28: 691–697. Experimental archaeology at the crossroads: interpretation or ‘xeroxing’? 57

Coles, J. (1973) Archaeology by Experiment, London: Hutchinson. Coles, J. (1979) Experimental Archaeology, London: Academic Press. Croes, D. R. (1988) ‘The significance of the 3000 BP Hoko River waterlogged fishing camp in our understanding of southern northwest coast cultural evolution’, in B. Purdy (ed.) Wet Site Archaeology, Caldwell: Telford Press, 131–152. Croes, D. R., Fagan, J. L. and Zehendner, M. N. (2009) ‘Sunken village, Sauvie Island, Oregon, USA: a report on the 2006–2007 investigations of National Historic Landmark Wet Site 35MU4’, Journal of Wetland Archaeology, 9: 1–216. Cunningham, P., Heeb, J. and Paardekooper, R. (eds) (2008) Experiencing Archaeology by Experiment, Oxford: Oxbow. Darwin, C. (1881) The Formation of Vegetable Mould through the Action of Worms, with Observations on Their Habits, London: Faber and Faber. Ferguson, J. R. (2010) Designing Experimental Research in Archaeology, Boulder, CO: University of Colorado Press. Goldberg, P. and Macphail, R. I. (2006) Practical and Theoretical Geoarchaeology, Oxford: Blackwell. Hansen, C. (2007) ‘Experiment and experience: practice in a collaborative environment’, in P. Cunningham, J. Heeb and R. Paardekooper (eds) Experiencing Archaeology by Experiment, Oxford: Oxbow, 69–80. Harding, D. W. (2009) The Iron Age Round-House, Oxford: Oxford University Press. Harding, D. W., Blake, I. M. and Reynolds, P. J. (1993) An Iron Age Settlement in Dorset: Excavation and Reconstruction, Department of Archaeology Monograph 1, Edinburgh: Edinburgh University Press. Hodder, I. (1999) The Archaeological Process, Oxford: Blackwell. Holst, M. K., Breuning-Madsen, H. and Rasmussen, M. (2001) ‘The South Scandinavian barrows with well-preserved oak-log coffins’, Antiquity, 75: 126–136. Hurcombe, L. (2008) ‘Organics from inorganics: using experimental archaeology as a research tool for studying perishable material culture’, World Archaeology, 40: 83–115. Jewell, P. A. (ed.) (1963) The Experimental Earthwork at Overton Down, Wiltshire, 1960, London: British Association for the Advancement of Science. Johnson, D. L. (2002) ‘Darwin would be proud: bioturbation, dynamic denudation and the power of theory in science’, Geoarchaeology, 17(1): 7–40. Johnson, M. (2010) Archaeological Theory: An Introduction, 2nd edn, Oxford: Wiley Blackwell. Jones, A. (2002) Archaeological Theory and Scientific Practice, Cambridge: Cambridge University Press. Koerner, S. (2011) ‘Experimental archaeology after simplicity – implications for reflexity of insights that a “Common World” is not “given”’, in D. Millsom (ed.) Experimentation and Interpretation, Oxford: Oxbow, 61–95. Macphail, R. I., Cruise, G. M., Allen, M. J., Linderholm, J. and Reynolds, P. J. (2004) ‘Archaeological soil and pollen analysis of experimental floor deposits with special reference to Butser Ancient Farm, Hampshire UK’, Journal of Archaeological Science, 31(2): 175–192. Mathieu, J. R. (2002) Experimental Archaeology, International Series 1035, Oxford: British Archaeological Reports. Millson, D. C. E. (2011) ‘Introduction’, in D. C. E. Millson (ed.) Experimentation and Interpretation, Oxford: Oxbow, 1–6. Paardekooper, R. (2013) The Value of an Open-Air Museum is in its Use, Leiden: Sidestone Press. Parker Pearson, M. (1996) ‘Food, fertility and front doors in the first millennium BC’, in T. C. Champion and J. R. Collis (eds) The Iron Age in Britain and Ireland: Recent Trends, Sheffield: Collis Publications, 117–132. 58 Martin Bell

Pope, R. (2007) ‘Ritual and round houses: a critique of recent ideas on the use of domestic space in later British prehistory’, in C. Haselgrove and R. Pope (eds) The Earlier Iron Age in Britain and the Near Continent, Oxford: Oxbow, 204–228. Rasmussen, M. (ed.) (2007) Iron Age Houses in Flame, Lejre: Historical-Archaeological Experimental Centre. Reynolds, P. J. (1979) Iron Age Farm: The Butser Experiment, London: British Museum. Reynolds, P. J. (1994) ‘Experimental archaeology: the future perspective,’ Reuvens Lecture 5, Stichting voor de Nederlande Archeologie, Holland. Reynolds, P. J. (1995) ‘The life and death of a post-hole’, in E. Shepherd (ed.) Interpreting Stratigraphy 5: 1994 Norwich, Norwich: Interpreting Stratigraphy Group, 21–25. Reynolds, P. J. (1996) ‘Experimental domestic octagonal earthworks’, in M. Bell, P. J. Fowler and S. W. Hillson (eds) The Experimental Earthwork Project 1960–92, Research Report 100, York: Council for British Archaeology, 225–227. Reynolds, P. J. (1999) ‘Butser ancient farm, Hampshire, UK’, in P. G. Stone and P. G. Planel (eds) The Constructed Past: Experimental Archaeology, Education and the Public, London: Routledge, 124–135. Reynolds, P. J. (2006) ‘The scientific basis for the reconstruction of prehistoric and protohistoric houses, euroREA, 3: 58–68. Reynolds, P. J. and Hedge, R. (1999) The Butser Ancient Farm Earthworks Research Project, Monograph 4, Petersfield: Butser Ancient Farm. Schiffer, M. B. (1987) Formation Processes of the Archaeological Record, Salt Lake City, UT: University of Utah Press. Schiffer, M. B. (2010) Behavioural Archaeology: Principles and Practice, London: Equinox. Steensberg, A. (1979) Draved: An Experiment in Stone Age Agriculture: Burning, Sowing and Harvesting, Copenhagen: National Museum of Denmark. Tipper, J. (2012) ‘Experimental archaeology and fire: the investigation of a burnt reconstruction at West Stow Anglo-Saxon Village’, East Anglian Archaeology, 146: 114–20. Townend, S. (2007) ‘What have reconstructed roundhouses ever done for us . . .?’, Proceedings of the Prehistoric Society, 73: 97–111. van Gijn, A. L. (1990) The Wear and Tear of Flint. Principles of Functional Analysis Applied to Dutch Neolithic Assemblages, Leiden: Analecta Praehistorica Leidensia 22. Wood, W. R. and Johnson, D. L. (1978) ‘A survey of disturbance processes in archaeological site formation’, in M. B. Schiffer (ed.) Advances in Archaeological Method and Theory 1, New York: Springer, 315–383. Wylie, A. (2002) Thinking from Things, Berkeley, CA: University of California Press. 4 ‘PROPORTIONAL REPRESENTATION’

Multiple voices in archaeological interpretation at C¸atalhöyük

Shahina Farid

Reopening C¸atalhöyük In 1993, after nearly three decades of neglect, excavations were reopened at the Neolithic site of Çatalhöyük in central Anatolia, by Ian Hodder. He committed to a 25-year excavation and research program, with the aim of placing the world- famous site – with its layered symbolism relating to its practices of early settled life, farming and art – within an environmental, economic and social context in Southwest Asia; he planned to focus on the origins of the site and its early develop- ment. To achieve this, Hodder has employed all the latest archaeological methods available for excavation and specialist analysis; an impressive array of scientific techniques, digital technologies and analytical tools have been applied within a robust excavation and sampling framework, all based on the British single context excavation and recording system and collated in a fully integrated, relational GIS database. What was really innovative about this research program, however, was Hodder’s resolve to put these cutting-edge methods to work in the context of a distinctive theoretical framework characterized by a commitment to reflexivity and multi- vocality. He argued that the archaeological community has a duty to diverse stakeholders – local communities, the public, tourist industries and national and international policy makers – and that all these voices should be represented in the research agenda and the interpretation of the site or, at the very least, that they should be provided with a platform to express their ideas or concerns. One goal of the Çatalhöyük project was to show how this duty could be translated into practice. For those who participate actively in the work – the digging, the research, the analysis – this means cultivating a self-critical awareness that, as archaeologists, our interpretations reflect our personal histories, experiences and backgrounds. Hodder established an international team structure not only to secure a broad resource base for this large-scale, long-term project but also to see whether archaeologists from 60 Shahina Farid different cultural backgrounds and with different archaeological skills, all working at the same site, would generate divergent interpretations or arrive at comparable narratives of the past. My aim in this chapter is to reflect upon some of the successes and failures of this experiment and to assess problems associated with the distinctive methodologies employed at Çatalhöyük, focusing particularly on the fieldwork strategies implemented between 1995 and 2011.

The setting When James Mellaart excavated Çatalhöyük in the 1960s, little was known about the Neolithic outside the Fertile Crescent where food-producing societies first emerged, herding livestock, planting crops and building permanent houses. In fact, it was generally accepted that there would be no evidence of the Neolithic in Anatolia, given the cold climatic conditions (Hodder 2006). Mellaart’s excavations therefore shifted the boundaries of the Neolithic profile westwards, with the identification of crop cultivation and domesticated sheep, goat and cattle at Çatalhöyük. The site rapidly became famous for its large size and dense occupation of closely packed mudbrick houses interspersed with midden areas, and also for its art. Houses were dismantled and rebuilt in more than 13 construction phases, resulting in the 20-meter-high mound of today, which covers an area of over 32 acres and represents over 1400 years of continuous Neolithic occupation dating to 7400–6000 BCE. The houses were entered by ladders from roof openings, and inside a wide array of the remains of everyday life had been preserved. These remains included evidence of routine daily activities and household production that had taken place in the context of a highly organized configuration of house ‘furniture,’ as well as burials; the dead of Çatalhöyük were buried underneath house floors. Some houses were elaborated with wall paintings: abstract compositions of geometric patterns in red and black on the white background of the plastered walls, or narrative scenes including hunting scenes with small human figures in opposition with oversized wild animals. Others incorporated relief sculpture and elaborate installations embellished with animal parts.

Excavation methodology Preliminary work was conducted at Çatalhöyük in 1993–1994, the initial two years of the 25-year program (Hodder 1996), by a team comprised of the field director, laboratory analysts and a group of students. For the first excavation season in 1995, Hodder added a volunteer ‘professional archaeologist,’ the role that I fulfilled from 1995 through to 2011. It was at this time that an excavation and recording system was devised for the long-term project, designed to suit the archaeology of the site, based on the 1993–1994 season’s preliminary work (Matthews and Farid 1996), and with two other considerations in mind: the need for a system that would ensure continuity at a site where long-term research was envisioned; and the theoretically informed commitment to a reflexive methodology. ‘Proportional representation’: multiple voices at Çatalhöyük 61

Combining continuity and flexibility Where this first was concerned, it was inevitable that new or revised questions, approaches and analysis would be introduced in the course of the excavation program, so, from the outset, provisions were made for incorporating new data into a pre-existing system. In some cases these new directions required the adoption of new modes of data retrieval and recording methods for the duration of a particular research angle. Changes in team personnel could also alter these practices so it was crucial that a log be maintained recording when alterations were adopted, why and for how long. This would enable future researchers to reconstruct what earlier research teams had done and to make well-informed use of the data they generated. At the same time, however, it was necessary to ensure that a basic set of data was routinely recorded to ensure consistency throughout the duration of the long-term excavation program. To achieve this, purpose-designed recording sheets were developed for logging mundane but crucial information related to excavated units: their location, context type, dimensions, volume; the specifics of samples and finds recovered from them; their stratigraphic associations and relationships to a hierarchy of associated features, buildings and areas; and a record of photo and plan logs. Prompts on proforma sheets such as these served as a reminder to the excavator to record all relevant data, without which meaningful information may have been irretrievably lost. The proforma recording sheets were modeled on the British single context system, which is based on the excavation and recording of individual events, depositional or intrusive (cuts, removal, collapse). The excavator identifies these episodes and isolates them within a stratigraphic sequence, capturing the order in which each event occurred in time. Once defined, the single context is recorded in relation to the site grid: a single coordinate system established for the whole of the site, based on a primary baseline that is tied into a global coordinate system. It is this system that allows the location of the site to be viewed within its environmental setting and in relation to other sites. All excavation trenches or areas of a site, and everything excavated and recorded within a trench or area, are thus spatially related in the system (see Figure 4.1). The single context system is especially suited to the complexities of deep urban archaeology of the kind represented by the 20-meter-deep complex stratigraphy at Çatalhöyük. Every episode identified by the excavator is given a unique number, the context number, which at Çatalhöyük is referred to as the ‘unit.’ As the deposit is excavated, this unique number serves as the ‘primary key’ to the unit, appearing in the prefixes and suffixes for all finds and samples associated with that unit. This makes it possible to coordinate all references to the unit in field and laboratory notes and in reports, and all measured plans and drawings, photography and video records of the unit, across all datasets from the time the unit is excavated through to publi- cation. Anchored as it is in the identification and excavation of single deposition episodes in stratigraphic sequence, this system makes it possible to assign a context number to each unit, which represents its relation to all other units illustrated within the Harris matrix as above (later), below (earlier) or equals (same as) (see Figure 4.2). FIGURE 4.1 Composite of single context plans within their correct spatial location on the site grid. North Area. (Screen grab; copyright: Çatalhöyük Research Project.)

FIGURE 4.2 The Harris matrix depicts the temporal succession of archaeological contexts and thus the sequence of deposition. At Çatalhöyük, it is grouped into features, which appear in spaces, which in turn form buildings. (Adapted from Farid (in press a): Figure 8.2.) ‘Proportional representation’: multiple voices at Çatalhöyük 63

This establishes a basis for grouping related events, which, at Çatalhöyük, are drawn together as features (walls, fixtures, events) defined within spaces (rooms, areas). Spaces in turn constitute buildings, which are ordered in the sequence in which they were built over earlier buildings or spaces. In this way the history of the site is gradually built up in a diagram of numbered units, features, buildings and spaces representing the stratigraphic relationships. Further refinement of the matrix involves the interpretation of phases – related periods of activity (Farid in press b) – which, depending on the type of site, can represent major or micro-scale time periods. When provenance is secure it is possible to compare and analyze the resulting soil components and inclusions, both natural and cultural artifacts, within the framework of their stratigraphic relationships. All the meaningful patterns unpicked and elucidated by the analysts, and all interpretations of these patterns, are based upon these spatial and temporal relationships. The value of any object, find or sample depends fundamentally on the security of its provenience; rarely does the importance of an object, find or sample transcend its provenience. Put simply, a poorly excavated site will result in an inaccurate understanding of temporal sequence, which, in turn, will result in analysts’ identifying spurious patterns, and this cannot but result in misleading interpretive narratives.

Reflexive methodology However much excavation and recording is standardized, it is inescapable that archaeological interpretation reflects the individual life experiences, cultural background and education of practitioners. It is this recognition that motivated the second set of considerations raised above: at Çatalhöyük the goal was to put these influences to work productively by implementing a reflexive methodology. Rather than assembling the results of compartmentalized analysis and interpretation after the fact, Hodder envisioned an ongoing collaborative process designed to combine different types of archaeological practice, skills and perspectives in ways that, it was hoped, would enhance our understanding of the relationship between objects and contexts and peoples of the past, and ultimately our understanding of the com- plexities of the site and of this period in history. This process is outlined in Hodder’s 12-point plan for incorporating reflexive methods into excavation practice (Hodder 2000). He had been developing this plan since the beginning of the project in 1993 (Hodder 1999), and it is informed by his critique of the assumption, often reinforced by the conventions of archaeo- logical publication, that the research process is linear, moving from excavation and recording to analysis and interpretation. Hodder argues that this is misleading, rejecting the generally accepted premise that excavation produces an empirical foundation of objective data – archaeological material, objects and contexts – that are independent of theory. Rather than assuming that the interpretation of these data takes place in a separate stage of inquiry after they are recovered and recorded, he insists that there is an ongoing interaction between data and theory; the 64 Shahina Farid

FIGURE 4.3 Hodder’s 1 Priority tours – team engagement 12-point plan towards reflexive methods. 2 Sampling strategies – diversity (Adapted from Hodder 3 Contextualized artifacts – quick feedback on artifacts 2000.) 4 Integrated and fluid database – immediate access 5 Diary entries – self-reflection 6 Video diaries – visual 7 Resident anthropologists – process of interpretation 8 Web deposition – shared access 9 Hypertext – relational database 10 Multivisual data – use of multimedia 11 Virtual reality – use of diverse illustration techniques 12 Different windows – use of diverse excavation methods relationship between these stages and aspects of an archaeological project is multi- directional and dynamic. The rationale for implementing a reflexive methodology at Çatalhöyük is that it explicitly recognizes this process and engages it ‘at the trowel’s edge’ (Hodder 1997); it has been extensively written about and discussed, and variously adopted by different excavation projects (Chadwick 1998; Farid et al. 2000; Hamilton 2000; Framework Archaeology, T5 Heathrow; Andrews et al. 2000; Berggren 2009; Davidovic 2010; Nilsson and Berggren 2010). Although many will argue that cross-team engagement is common in archae- ology, these interactions are generally informal, and even when formal team dis- cussions are conducted at the report-writing and publication stage they do not typically involve all the original team members. The result is that the daily reflections of those who worked ‘at the coalface’ are rarely heard, either in the interpretation process or in the final report. Hodder’s 12-point plan (see Figure 4.3) is intended to formalize cross-team engagement and cement it into the daily routine of excavation and recording so that interpretation arises from continuous, inclusive collaboration, and excavation itself is informed by multiple perspectives. In addition, this plan puts a priority on explicitly documenting decision-making processes that typically remain implicit. At Çatalhöyük these goals were to be achieved by the introduction of:

• priority tours and structured team discussion; • purpose-designed sampling strategies; • conventional and video diaries that document practice; • shared-access data for all team members; • engagement with the wider context of the project, including local and regional as well as national and global interests.

Hodder emphasizes that these interactive methods are intended to empower the excavator to make better decisions and interpretations. Everyone is better informed ‘Proportional representation’: multiple voices at Çatalhöyük 65 when specialists engage with excavators and vice-versa; this feeds into sampling decisions and can lead to innovative ‘trowel’s edge’ interpretations. In what follows I discuss what can be learned from the experience of implementing the first four of these key elements of a reflexive methodology in the excavation process at Çatalhöyük. The last of the key elements relates to community and diverse group engagement, site presentation and the overall relationship to and impact of the project in the local and national arena, but this would require a chapter of its own and has been discussed elsewhere (Hodder 2000; Bartu 2000: Bartu Candan 2005; Farid forthcoming).

Engaging with reflexive methods

Conditions for success For inter-team dialogue to be effective and meaningful several factors must be considered. The aims of the project and its methodologies need to be made explicit to all team members at the outset, and everyone involved must be prepared to make a commitment to engage them. In addition, all teams need a good balance of experienced members and trainees. Generally great importance is placed on recruit- ing highly qualified and reputable material analysts, with less emphasis on ensuring that the excavation team is similarly qualified. All too often excavation is treated as an exercise that anyone can turn a hand to: volunteers, students or academics, with or without training and experience. And yet, what is required of a good excavator is a set of skills garnered through years of experience: the ability to identify individual depositional events, to unravel stratigraphic sequence, to adapt excavation and sampling methods so they are efficient, to establish a robust database and to serve evolving research goals. To address these needs at Çatalhöyük, the UK-led team employed professional contract excavators, but this was not true of all teams. The result was that, although from the outset the aim had been to mix all teams, to share experiences and methods, this proved logistically difficult given different levels of expertise and ways of practicing archaeology, not to mention the challenges of language barriers and different timetables for on-site work. This last raises another important issue: that, for consistent flow of information, all teams and all team members need to be on site during the whole of the excavation season. On many projects the excavation and the materials analysis are conducted at different times or, even if they occur at the same time, with a minimum of communication between laboratory and field team members. The physical separation of field and laboratory work means that analysts rarely see the excavations and therefore lack an appreciation of the context of deposition; ‘finds’ are separated, bagged at the trench edge and delivered to different analysts who have little access to other datasets. Drawing the different strands of data together is deferred to the final stages of writing up and publication. Conversely, the excavation team rarely gets feedback from the analysts about the results of their work. Çatalhöyük afforded an opportunity to plan for excavation and materials 66 Shahina Farid analysis that would run in tandem and to build regular inter-team discourse into the daily routine to lead to collaborative interpretation.

Challenges Despite the best of intentions, this ideal was not fully realized at Çatalhöyük, which contributed to inter-team tensions and frustrated expectations, impeding the implementation of Hodder’s reflexive methodology. In practice there was a marked imbalance between experienced excavators and students, which was aggravated by the regular rotation of students who joined the project for short periods of time and required constant training. The excavation teams were under enormous pressure to achieve the seasons’ excavation goals and manage the daily workload, while at the same time being expected to engage with the team tours and to partici- pate in collaborative discussions of interpretative issues and outreach activities. Similar imbalances were evident within the analysts’ teams: the experienced laboratory heads were rarely present for the duration of the season so they relied on a rotation of assistants and students whose input was therefore as limited as their experience. In addition, the information flow between field and laboratory was uneven because not all material finds were analyzed during the field season. Although on-site laboratories covered key materials while the excavations were ongoing – for example, human and animal bone, pottery, botanical remains – other types of material (e.g. worked stone, lithics, clay balls) and smaller assemblage groups (shell, basketry or small mammal bone) were only studied every few years when a corpus of material had been collected. The impact on the multidisciplinary makeup at Çatalhöyük was visible; there was a constant state of flux within and between laboratory and field teams, making it difficult to establish effective, ongoing communication. There were also tensions between the excavation teams – in particular, between the British-led paid teams and other teams that were led by academics and relied on volunteers and students. The major point of contention was the use of the single context excavation and recording system with which many team members, including excavators and analysts, were unfamiliar. This came to represent a British tyranny of imposed excavation and recording methods, especially for those who did not see the single context system as especially well suited to deep complex stratigraphy. As new teams joined the project the single context system would be scrutinized, questioned and criticized and duly defended by its practitioners. In response to their concerns, additional recording methods were introduced as a supplement to the single context method, such as the use of daily sketch plans, which, for some, served as the primary way of tracking what was excavated by the end of each day. These were most useful as a daily visual diary, but they record physical, not stratigraphic relationships, so they were no substitute for systematic use of the single context recording sheets, and for those familiar with the single context system they represented an additional layer of superfluous work. Overall, this process of questioning the recording system served to ‘make explicit the ‘Proportional representation’: multiple voices at Çatalhöyük 67 documentation process,’ which, in turn, elucidated the decision making by which excavators contributed to the process of knowledge production and interpretation. It can be productive to challenge predetermined approaches, but, with excavation teams continuously criticizing one another’s methodologies and distrusting their interpretations, it was also unsettling and came at a cost to overall team cohesion.

Reflexive methods in practice

Priority tours and structured team discussion Priority tours evolved over the first few seasons of the project, as a strategy for fostering communication between teams and team members. Hodder’s initial idea had been to have analysts work at the side of the trenches so they would be on hand to provide immediate in situ interpretation; for example, a mixed assemblage would indicate a midden or pit fill, or a sterile deposit might identify infill, the premise being that excavators were unable to make such interpretations. In practice, however, trench-side lab-work stations were deemed impractical from the outset and so, instead, daily laboratory tours to the excavation areas were timetabled for cross-information exchange. These proved to be time-consuming, adding to the existing overload of work in the field and laboratories, and it became apparent that not all components of a context were feeding into the interpretation process, skewing context-by-context interpretation. While smaller assemblages could be studied and recorded in full, lab work on the large groups, such as the faunal and lithics assemblages, lagged behind. So, for instance, a deposit might be interpreted as hearth-related given a report that it was rich in seeds and charcoal with some obsidian flakes, but this would be misleading if, in fact, it also contained faunal and stone assemblages characteristic of a midden deposit that had not yet been studied. In practice it was the excavators who had a comprehensive under- standing of what was coming out of the ground, but in these early years their interpretive skills were not given the same recognition and weight as those of the lab-based specialists. The ‘lab tours’ thus evolved into ‘priority tours,’ with the objective of ‘prioritizing’ a subset of contexts under excavation for complete material analysis. On alternate days a representative from each working laboratory joined a tour of the site led by excavators who explained the current stage of work on deposits or features, the type of materials being recovered, their stratigraphic location and their working interpretations (Figure 4.4). One of the main aims of the tour was for all present to negotiate the selection of contexts whose material components would all be ‘fast-tracked’ – through the process of flotation; finds washing, sorting and logging; distribution to different laboratories for analysis – so that the results of analysis and working interpretations could be reported back by the following priority tour. This ensured that a proportion of contexts excavated over the course of a field season would be fully studied and collaboratively interpreted. There were no set criteria for the choice of a priority deposit; much depended on what was 68 Shahina Farid

FIGURE 4.4 Priority tour of the foundation trench excavations, North Area, 2007. (Photograph: Jason Quinlan; copyright: Çatalhöyük Research Project.) under excavation at the time, and on any anomalies that had been noted during excavation. For example, the selection of an abnormally rich deposit of large bone for priority analysis made possible its interpretation, at the feedback stage, as ‘feasting’ discard; in conjunction with its location, analysis showed that it was a primary deposition of large, meaty elements of cattle and that it was buried relatively quickly, given the absence of weathering and fragmentation (Russell et al. 2014). In other cases deposits were targeted for fast-track analysis by the archaeobotanist because they were rich in organic remains, sometimes with unexpected results. On feedback, several of these proved to be rich in micro-bone elements and obsidian chips as well as organic material, initiating a collaborative process of interpretation that took into account the context of deposition as well as its composition, but it was the location of the deposit in the oven zone of the building where household sweepings had accumulated amidst the burnt dung fuel that was crucial to its interpretation. This kind of exchange does show what can be accomplished by implementing Hodder’s plan: the process of team collaboration results in a better understanding of the formation of deposits and of the objects found within these contexts, and this feeds back into interpretation of the contexts and can challenge pre-existing assumptions about them. One of the site-wide interpretations taken for granted today that was not considered in the 1960s was resolved by the priority tour system. A pattern emerged across the site early on in the excavations, which suggested that, before the construction of a new house over the footprint of its predecessor, the shell of the old house was infilled with a fine-grained homogeneous deposit with a sterile ‘Proportional representation’: multiple voices at Çatalhöyük 69 material signature. The source and function of this large quantity of infill was not understood and became the subject of nearly a season’s worth of priority tours, with intensive sampling and assemblage analysis conducted across a number of similar deposits under excavation. The homogeneity and rapid deposition of the infill material were established through excavation and its sterile signature confirmed by material analysis. An initial collaboratively developed interpretation was that this soil came from an off-site source and was sifted prior to deposition to provide a solid, compact foundation for the new building. This interpretation had to be revisited, however, when the infill of another building broke this pattern: it was comprised of mixed differentially deposited layers, which, given the results of fast- track analysis, indicated a typical midden assemblage signature. After much prioritizing and sampling and analysis of deposition episodes, a new interpretation was arrived at that identified the fill as redeposited midden. This ultimately led to a revision of the interpretation of all house infill deposits. The house that broke the pattern (Building 2, see Farid 2007) had not been immediately rebuilt after abandonment; after some time the vacated house plot had to be filled in to create a use horizon for neighboring houses as they were rebuilt, which was done by introducing soil from elsewhere. This led to a further re-evaluation of sterile infill which established that it was made up of the crushed walls of the upper portion of the vacated house; systematic comparison showed that the composition of in situ mudbrick, mortar and plaster typically matches that of the infill, although in some buildings infill is topped up from other sources as well. This called into question the assumption that infill material had been laboriously transported from sources outside the settlement and it also introduced the concept of intra-site landscaping. The new data generated through the process of inter-team collaboration, with back and forth engagement between analysts and excavators, made it possible to reassess – and ultimately fundamentally change – long-established interpretations.

Purpose-designed sampling strategies Different types of sampling strategies, designed to address specific questions raised by contexts under excavation, were also introduced during priority tours (Farid et al. 2000). From the outset a 100 percent sampling methodology for all excavated contexts was instituted at Çatalhöyük. Not all excavations have the resources to do this – it is tedious and time consuming, and raises huge storage issues – but it has had a measure of success at Çatalhöyük. Routinely a (small) predetermined measure of material from each deposit is stored as an ‘archive’ sample, to provide for any analysis that future researchers might want to undertake. The remaining deposit is put through flotation, and if the context is large a predetermined amount is wet- sieved and the remainder dry-sieved. In addition, all samples, including the archive, flotation and special-purpose samples, and especially the dry-sieved component, are quantified so that cross comparisons of material densities can be made. Even though the constraints of fieldwork mean that this quantification is quite crude – it relies on counting (same-size) buckets and half-buckets per wheelbarrow – it 70 Shahina Farid has proven to be invaluable for many of the research programs undertaken at Çatalhöyük. For example, the interpretative narratives that distinguish daily activities from ‘ritual’ or other unusual activities all depend upon the identification of density patterns in the material components of single deposits. In addition to these routine samples, as new research questions and projects are introduced other types of samples are added and the sampling program as a whole is regularly updated (Farid et al. 2000: Appendix 2.1). The list is long and ever growing. For example, a program of cattle DNA analysis required the collection of a complete set of cattle teeth or metapodia. Another program, targeted for a PhD thesis, involved identifying coprolite types in middens, and other studies required soil analysis of the stomach region of human skeletons. Others included pigment analysis and residue analysis on pottery, clay balls and ground stone. Although productive in ways I will describe shortly, each analysis type had to be collected in the field in specific and time-consuming ways. This was an arduous task, especially for team members who were not accustomed to using the single context system. In cases where teams excavated by feature (multiple contexts) or by ‘spit’ (arbitrary layers), it became impossible to compare material types and analyses across deposition episodes. As a growing number of sampling strategies are introduced to address different research questions, it becomes ever more important to maintain the single context method with its tight stratigraphic control and identification of single deposition episodes. In practice, many research questions could only be addressed using a rigorous sampling system undertaken in the framework of the single context method of excavation and recording. For instance, in some trenches the single context system made it possible to differentiate between brick, mortar and plaster when excavating a wall. This brought to light one of the common features of the buildings at Çatalhöyük: that the same mudbrick and mortar was used throughout the construction of each house, so that generally superimposed walls can be visually identified by a change in composition of mudbrick and mortar. A PhD program (Love 2010) located preferred sources of raw material, and patterns of use and attrition of resources in the environment over the life of the site, based on analysis of the soil composition of individual brick, mortar and plaster components and its comparison with soil formation in the surrounding landscape. Similarly, individual components of construction and maintenance of household furniture, such as the inner core and multiple clay and plaster renders of ovens, platforms, benches, bins and basins, were excavated and sampled. This resulted in a PhD thesis that focused on the contribution made by shared resources, including knowledge of house building and maintenance techniques, to the stability of the Neolithic settlement through its long history (Tung 2008). Conversely, there are examples whereby intensive excavation and sampling of individual deposits were readdressed because of the results of detailed analysis. It had been assumed that internal floors would show evidence of daily house-based activities, which were believed to have been organized into different zones demarcated by kerbs and the edges of platforms and benches. The outcome of ‘Proportional representation’: multiple voices at Çatalhöyük 71 lens-by-lens excavation of the floors, and sampling for chemical analysis, phosphate, phytoliths and micro-debitage on a 1-meter grid, was the recognition that these repeatedly plastered floors were clean, and sometimes scoured, at the closure phase of the house. Although the composition of the floor plaster and bedding clays could be identified and analysis did provide evidence that reed matting and skins were sometimes used across the floors, little evidence of household activities survived. Based on these results, floors are no longer intensively sampled. Similarly intensive sampling and analysis of the deposits found inside bins and basins produced sterile signatures with no evidence of primary use or function – they, too, had been scoured clean prior to the house being filled in at the closure phase so, again, this type of intensive sampling was stopped. These deposit signatures are now well established but it remains for the excavators to continually reassess and adapt excavation methods as divergences of current patterns are recognized, or new questions, and therefore sampling strategies, are introduced. One of the areas where single context excavation and sampling is impractical is middens. These are external areas where daily refuse accumulated and was frequently trampled; as a result they consist of finely lensed, inter-digitated deposits. It proved impractical to identify individual lenses in these contexts, being diffuse and discontinuous, so, after some years of attempting to excavate lens by lens, middens are now excavated in discrete deposition horizons separated by intervening events: a fire, or the excavation of a pit, or weathered surfaces. The important factor here is the excavators’ adaptability to requirement – knowing when and when not to employ single context excavation. While the reflexive refinement of purpose-designed sampling strategies was productive in many ways, in practice the outcome of collective deliberation was not always positive. At times excavators felt oppressed by what was expected of them, while laboratory teams felt that the excavations compromised the samples and artifacts on which they based their interpretations. One such example is micromorphology analysis. Despite concerns voiced by different team members at the start of the excavations in the mid-1990s, it was decided that the entire sequence of all buildings would be excavated in alternate 1-meter squares, in a checkerboard fashion. The resulting sections were expected to reveal the history of the building from which micromorphology blocks could be taken for analysis of micro- deposition events and composition. For the excavators this was a truly debilitating methodology; they were left with no stratigraphic control and, given differences in how individual squares were dug, it was sometimes not possible to link the layers identified in one square with those identified in adjoining squares (for Building 2, see Farid 2007; for Building 1, see Cessford 2007). This system highlighted a major rift between team members with respect to the use of sections (profiles) (Hamilton 2000). Some archaeologists rely on sections as a guide and check and establish as many as possible, while others use an open-plan excavation method and make use of sections only where they occur at trench boundaries. Because of the restrictions that grid excavation for micromorphology imposed, it was eventually abandoned in favor of excavation in plan; instead micromorphological blocks for analysis were 72 Shahina Farid taken from discrete upstanding blocks established in areas of interest, or from the sides of cuts and sections. A better approach, one that would have made it possible to assess grounded arguments for and against both systems, would have been to adopt a strategy by which a sample of houses was excavated using the grid method, from which the insights gained from micromorphology could be compared with the results from excavation in plan.

Documentation practice Hodder’s reflexive methodology calls for continuous documentation aimed at making explicit the collaborative processes by which ‘interpretation at the trowel’s edge’ takes shape. This documentation is undertaken in different ways and for different reasons: to capture decision making; to support the process of collaborative interpretation; to revisit interpretations made ‘at the trowel’s edge’; and to provide a record for the future of what is lost through excavation. To compensate for the perceived rigidity of the system of proforma recording sheets (Figure 4.5), these were supplemented by digital diary entries; the hope was that this would provide an opportunity to record individual thought processes in the style of the old ‘note- book’ system. Team members are encouraged to engage in free-thinking reflection on the processes of excavation, recording and interpretation, offering a more personal and self-critical account of their work than they could include in the standardized site forms. Unfortunately these practices were taken up unevenly. The excavation teams were expected to engage actively with diary keeping, while the same was not expected of the laboratory teams. The result was that the excavators came to feel that they were under scrutiny while the laboratory teams were disgruntled because they were comparatively voiceless; there was not the same interest in how they arrived at the identification of a particular find or assemblage. Also, the value of diary entries is only as good as the quality of the information entered. It requires experience to identify streams of consciousness that are relevant to the construction of an archaeological interpretation, as opposed to just entering observations like: ‘bad day, ipod died.’ Many found it a forced exercise: it was difficult to switch on self-critical analysis at the end of a long day when paperwork and data entry had also to be tied up. Lack of time and motivation led to increasingly limited engagement with diary entries (Boddington 2013). In addition to personal text diaries, video diaries were introduced as a means of capturing visual information, words and sounds that would supplement the excavation text records, photos, plans and other drawings. A completely different perspective can be gained from watching and listening to team members explaining their work; if video capture is employed on a regular basis, it can provide insight into the process of interpretive, reflexive thinking as it develops over the course of a season. These reflexive processes are also sometimes captured by the project illustrators. While reconstruction drawings, such as the iconic visualizations produced by Alan Sorrell, are familiar in archaeology, more typically archaeological ‘Proportional representation’: multiple voices at Çatalhöyük 73

FIGURE 4.5 Proforma recording forms in use in Building 77, North Area, 2011. (Photograph: Jason Quinlan; copyright: Çatalhöyük Research Project.) illustrators patiently produce scale drawings of pot sherds, lumps of corroded metal, flint, stone and clay, and occasionally something quite stunning like a Saxon brooch. As digital technologies – digital photography, 3D scanners and assorted software – have become available, expediting the production of traditional ‘finds’ illustrations, archaeological illustrators are using their skills to depict events of the 74 Shahina Farid past, extrapolating beyond the data represented in conventional drawings of objects and creating picture-form narratives. This can be an invaluable tool: it is often the visualization of specific scenes or activities that can clarify whether an archaeological interpretation is plausible or has been thought through, and can capture events that have been missed because no physical trace of them survives in the archaeology. For example, it was only when an enactment for a documentary was conducted at Çatalhöyük that the disruption caused by digging a grave in the middle of a house became apparent. The fact that a grave had been dug in the past is an obvious assumption to make given the presence of the burial, but there is no evidence of the event of ‘digging’ the grave nor the pile of soil in the middle of the house. What had been missed was the middle bit: the pile of soil in the middle of the house. Recognizing this raised many new questions: How long would burial take? Would it be a house-centric event or would outsiders have witnessed it? As such, much is missing from the archaeological record that can only be teased out using multiple techniques and approaches; this example indicates the value of ‘experi- mental archaeology.’

Shared-access data A final aspect of reflexive methodology to consider here is the maintenance of data- sharing systems at Çatalhöyük. All data are made available via the project database as quickly as possible: contextual metadata, photos, diaries and videos from the excavations are made available by the time the laboratory teams analyze their material. The material analysts have their own purpose-designed databases and they input their data on a daily basis as well, so all information is accessible in a single integrated database, which, by the end of the season, is linked to updated site plans. This relational database makes available, within fingertip reach, a large and invaluable dataset that can support elaborate searches and complex queries across a myriad of datasets. It must be stressed, however, that the data output of such a system will only ever be as good as the data entered into it, and this depends on what archaeologists decide is pertinent to record, whether comparable attributes are reported for different materials and whether data entry is consistent and validated. To realize the advantages of a shared-access relational database, everyone working on the system must enter comparable information: the system fails if team members use different terminology to record the same technological process, or if different teams excavate in different ways and record incompatible data. These requirements for the construction of a database epitomize the tension between a commitment to support fluidity and diversity on the one hand, and, on the other, the need for systematization; objections to the rigidity of the database correspond to criticisms of the single context recording system. It is precisely this tension that some archaeologists hope to address by exploring the semantic web. The rich corpus of information assembled in the site-wide database is dis- seminated in different ways. On site, data are shared through team seminars in which presentations are made on a particular research angle for group discussion ‘Proportional representation’: multiple voices at Çatalhöyük 75

FIGURE 4.6 Team seminar in preparation for publication, 2010. (Photograph: Jason Quinlan; copyright: Çatalhöyük Research Project.) of a theme – for instance, in presentations on the question of whether ‘house closure by burning’ was deliberate or accidental. Site seminars are more intensive and formalized during study seasons (Figure 4.6), when excavation and material datasets are brought together and interpretations discussed and argued over with the aim of reaching a consensus on interpretation in preparation for publication. If opinions conflict, both sides of the argument are presented. Interpretations are in a constant state of flux as new data, ideas and analyses are introduced, and there is not always a right or wrong answer to focal questions. Unfortunately, not all voices are represented during pre-publication discussions, as excavations must continue and site teams are discouraged from spending valuable excavation time in seminar discussion.

Conclusion: conflict resolution In this chapter I have addressed only some aspects of Hodder’s 12-point plan for implementing reflexive methods at Çatalhöyük; excavation team members are also expected to participate in outreach programs that engage local residents and wider- interest groups. Taken as a whole, a fully reflexive methodology is even more ambitious and demanding than what is described here. Moreover, the Çatalhöyük project grew too large, with many teams undertaking different projects. Although all of these are valid and the attempt to bring them all together at one site is laudable, without cohesion and adequate resources the result is an overwhelming and disorientating diversity of initiatives. Any project director seeking to adopt 76 Shahina Farid

Hodder’s model should be mindful of what it requires and how it will affect the output of a project. Perhaps only some aspects of the model should be adopted to suit the purpose of the research agenda, or perhaps a reflexive methodology should be adopted for delimited tranches of work that can be adequately resourced and managed, with clear objectives and realistic quantifiable deliverables. At Çatalhöyük, time restraints, pressure to achieve certain goals for the season and work overload resulted in compromises, in diary writing, for example, or self- reflexive interpretation (Berggren and Nilsson in prep.). The failure to embrace these methods fully, as well as divisive inter-team tensions, resulted in a great deal of criticism of the project, and ultimately resulted in a wholesale change of its longstanding personnel in 2010–2011. The reason given to lab teams was that they had been ‘getting comfortable with their interpretations and so not challenging each other or the assumptions that were being taken for granted’ (Balter 2010); the excavation teams were told that, while their work was invaluable, they were not wholeheartedly engaging reflexive methods. Conflicts and tensions naturally arise in any workplace, but with careful consideration they can be minimized. Perhaps some problems could have been alleviated if the reflexive methods developed for the interpretive process had been applied to the work process: if the pressures of the workload had been regularly acknowledged and addressed, and expectations modified on a seasonal basis. A better balance in team experience would have relieved some of the pressure on the contract excavators, and explicit articulation of what was expected – clarity about project objectives and about the methods to be employed – would also have addressed some of these issues. In addition, if teams had been mixed, inter-team tensions might have been less divisive. These measures could have led to mutual respect for one another’s expertise rather than fostering an appearance of constant scrutiny that led to a sense of distrust. In short, if the ‘faultlines’ inherent in the project (Hamilton 2000) had been acknowledged and addressed in a reflexive spirit, it may have been possible to ‘empower’ many more team members on a more even basis. In 2012, the revised group began mixing excavation teams and reviewing the recording system. It will be an interesting exercise, in the future, to compare the productivity of the project before and after the major changes implemented in 2010–2011; this watershed will afford an opportunity to assess whether ‘changing windows’ – the use of alternate excavation methods – really does have an impact on interpretations and overall results. Despite the problems I have identified here, overall the methodologies employed at Çatalhöyük between 1995 and 2011 have made it possible to study, at a new level of detail, gradual changes over time on a site that had otherwise stood out for its repetition and continuity. The recognition of a shift from house-based production to specialized household craft production, and therefore to shared economies, has been made possible by attention to detail in the archaeology and by the process of collaboratively building interpretations that draw on a wide range of expertise and analyses. The project has thus produced a dataset that is incomparable, one that will be used for years to come. Its legacy is, for now, second to none. ‘Proportional representation’: multiple voices at Çatalhöyük 77

Acknowledgment Everyone involved in the Çatalhöyük excavations, 1995–2011.

References Andrews, G. A., Barrett, J. C. and Lewis, J. S. C. (2000) ‘Interpretation not record: the practice of archaeology,’ Antiquity, 74: 525–530. Balter, B. (2010) ‘Hodder cleans house at famed Çatalhöyük dig.’ Available at: news.sciencemag.org/2010/09/hodder-cleans-house-famed-Çatalhöyük-dig_. Accessed 30 March 2014. Bartu, A. (2000) ‘Where is Çatalhöyük? Multiple sites in the construction of an archaeological site,’ in I. Hodder (ed.) Towards Reflexive Methods in Archaeology: The Example at Çatalhöyük, Cambridge: McDonald Institute for Archaeological Research and British Institute of Archaeology at Ankara. Bartu Candan, A. (2005) ‘Entanglements/encounters/engagements with prehistory: Çatalhöyük and its publics,’ in I. Hodder (ed.) Çatalhöyük Perspectives: Themes from the 1995–99 Seasons, Cambridge: McDonald Institute for Archaeological Research and British Institute of Archaeology at Ankara. Berggren, Å. (2009) ‘Evaluation of a reflexive attempt, the Citytunnel Project in retrospect,’ Archaeological Review from Cambridge, 24(1): 23–35. Berggren, Å., and Nilsson, B. (in prep.) ‘Evaluating reflexive methodologies at Çatalhöyük,’ in I. Hodder (ed.) Integrating Çatalhöyük, Çatalhöyük Research Project Series Volume 10, British Institute at Ankara Monograph No. 49, Monumenta Archaeologica 32, Los Angeles, CA: Cotsen Institute of Archaeology Press. Boddington, M., (2013) ‘Truth and archaeology: justification in archaeology,’ unpublished PhD thesis, Magdalene College, University of Cambridge. Cessford, C. (2007) ‘History and excavation of Buildings 1 and 5 and summary of phases,’ in I. Hodder (ed.) Excavating Çatalhöyük: South, North and KOPAL Area Reports from the 1995–99 Seasons, Cambridge: McDonald Institute for Archaeological Research and British Institute of Archaeology at Ankara, 345–360. Chadwick, A. (1998) ‘Archaeology at the edge of chaos: further towards reflexive excavation methodologies,’ Assemblage 3. Available at: www.shef.ac.uk/~assem/3/3chad.htm. Last accessed 2013. Davidovic, T. (2010) ‘Practices of archaeological knowledge production at Çatalhöyük, 2010.’ Available at: www.catalhoyuk.com/downloads/Archive_Report_2010.pdf. Accessed 30 March 2014. Farid, S., with contributions from Cessford, C., Berggren, Å., Cooper, A., Turnbull, R., Baysal, A., Leaver, S., Boyer, P., Cawdron, T., Hamilton, N., Hawkes, L., Knight, M., Webb, S. and team members (2000) ‘The excavation process at Çatalhöyük,’ in I. Hodder (ed.) Towards Reflexive Methods in Archaeology: The Example at Çatalhöyük, Cambridge: McDonald Institute for Archaeological Research and British Institute of Archaeology at Ankara, 19–35. Farid, S. (2007) ‘Level IX relative heights, Building 2, Buildings 22 & 16 and Building 17,’ in I. Hodder (ed.) Excavating Çatalhöyük: South, North and KOPAL Area Reports from the 1995–99 Seasons, Cambridge: McDonald Institute for Archaeological Research and British Institute of Archaeology at Ankara, 139–226. Farid, S. (in press a) ‘Building 50,’ in I. Hodder (ed.) Çatalhöyük Excavations: The 2000–2008 Seasons, Çatalhöyük Research Project Series Volume 7, British Institute at Ankara Monograph No. 46, Monumenta Archaeologica 29, Los Angeles, CA: Cotsen Institute of Archaeology Press. 78 Shahina Farid

Farid, S. (in press b) ‘Timelines: phasing Neolithic Çatalhöyük,’ in I. Hodder (ed.) Çatalhöyük Excavations: The 2000–2008 Seasons, Çatalhöyük Research Project Series Volume 7, British Institute at Ankara Monograph No. 46, Monumenta Archaeologica 29, Los Angeles, CA: Cotsen Institute of Archaeology Press. Farid, S. (forthcoming) ‘From excavation to dissemination – breaking down the barriers between archaeology and the public,’ in P. Stone and Z. Hui (eds) Sharing Archaeology: Academe, Practice and the Public, New York: Taylor and Francis. Framework Archaeology, T5 Heathrow. Available at: www.framearch.co.uk/about.html. Accessed 30 March 2014. Hamilton, C. (2000) ‘Faultlines: the construction of archaeological knowledge at Çatalhöyük,’ in I. Hodder (ed.) Towards Reflexive Methods in Archaeology: The Example at Çatalhöyük, Cambridge: McDonald Institute for Archaeological Research and British Institute of Archaeology at Ankara, 119–127. Hodder, I. (1996) On the Surface: Çatalhöyük 1993–95, Cambridge: McDonald Institute for Archaeological Research. Hodder, I. (1997) ‘Always momentary, fluid and flexible: towards a reflexive excavation methodology,’ Antiquity, 71: 691–700. Hodder, I. (ed.) (1999) The Archaeological Process: An Introduction, Oxford: Blackwell Books. Hodder, I. (ed.) (2000) ‘Towards Reflexive Methods in Archaeology: The Example at Çatalhöyük, Cambridge: McDonald Institute for Archaeological Research and British Institute of Archaeology at Ankara. Hodder, I. (2006) Çatalhöyük: The Leopard’s Tale, Revealing the mysteries of Turkey’s Ancient ‘Town’, London: Thames and Hudson. Love, S. (2010) ‘How houses build people: an archaeology of mudbrick houses from Çatalhöyük, Turkey,’ unpublished PhD dissertation, Stanford University. Matthews, W. and Farid, S. (1996) ‘Exploring the 1960s surface – the stratigraphy of Çatalhöyük,’ in I. Hodder (ed.) On the Surface: Çatalhöyük 1993–95, Cambridge: McDonald Institute for Archaeological Research. Nilsson, B. and Berggren, Å. (2010) ‘Evaluation of reflexive methods.’ Available at: www.catalhoyuk.com/downloads/Archive_Report_2010.pdf. Accessed 30 March 2014. Russell, N., Twiss, K. C., Orton, D. C. and Demirerg, G. A. (2014) ‘More on the Çatalhöyük mammal remains,’ in I. Hodder (ed.) Humans and Landscapes of Çatalhöyük Reports from the 2000–2008 Seasons, Çatalhöyük Research Project Series Volume 8, British Institute at Ankara Monograph No. 47, Monumenta Archaeologica 30, Los Angeles, CA: Cotsen Institute of Archaeology Press. Tung, B. (2008) ‘Making place, doing tradition: exploring intimate knowledge at Neolithic Çatalhöyük,’ unpublished PhD thesis, University of California, Berkeley. 5 INTEGRATING DATABASE DESIGN AND USE INTO RECORDING METHODOLOGIES

Michael J. Rains

Archaeologists were early adopters of computing techniques and technologies. The annual Computer Applications and Statistical Methods in Archaeology (CAA) conference was first held in Birmingham, UK in 1973 (caa-international.org). Over the years, many of the papers presented at CAA conferences have described the development and use of databases as part of excavation recording strategies. Early examples include the use of primitive personal or ‘home’ computers, the BBC Micro and Sinclair Spectrum in the UK, for example, usually with ‘hand-crafted’ bespoke software. These systems were developed to tackle specific issues in excavation recording, such as searchable artefact catalogues, or the automated production of stratigraphic matrix diagrams. Over time, advances in both hardware and software, particularly the widespread availability of ‘off the shelf’ database packages such as Access and spreadsheets such as Excel, have encouraged the adoption of computer- based recording strategies throughout archaeology. In recent years, the development of integrated database-based packages, such as Intrasis (www.intrasis.com), ARK (ark.iparchaeology.com) and the Integrated Archaeological Database (IADB), have fostered the much closer integration of database design and functionality into archaeological recording methodologies. This chapter will examine the relationship between research/excavation aims, recording strategies, data collection methodologies and database design and use in archaeological projects. It will look at how constraints such as money, time and logis- tics during the data gathering and recording phase of a project can conflict with the requirements of different end uses (and end users) of the database (for example, post- excavation analysis, wider academic research, archiving, publication and dissemina- tion); and how all of these impact on and are impacted by the site recording strategy. These issues will be examined from the perspective of the Integrated Archaeological Database (IADB) and its use in two very different archaeological environments, one a major academic research excavation project, the other within a commercial developer-funded archaeological contracting unit. 80 Michael J. Rains

It will be shown that the design of the database must be seen as an integral part of the development of the recording strategy and that, as far as possible, both should reflect the purposes to which the database is to be put.

A brief history of the IADB The Integrated Archaeological Database (IADB) (http://www.iadb.org.uk) has developed over many years. In the late 1980s, work began at the Scottish Urban Archaeological Trust (SUAT) in Perth to develop the concept of a computerised integrated database to record and help with the analysis of several large excavation projects on which SUAT was working at the time. The earliest versions of the IADB incorporated earlier work at Durham University and the Scottish Development Department (later to become Historic Scotland) to develop a simple stratigraphic analysis program, a basic context and find cataloging application and an early attempt at a single context plan digitising solution. The original concept of the IADB was to make available digital versions of the various excavation records as an easily accessible integrated resource for use in post- excavation analysis, and to provide a framework within which that analysis would be undertaken. Initially, the IADB only dealt with simple artefact records and strati- graphic unit or context records. Over time, the scope of the IADB has widened to cover other digital resources, including single context plans, photographs, stratigraphy diagrams, etc. Initially the IADB was described as a ‘digital workbench’ or a ‘com- puterised desktop’ (Rains 1995); today we would probably call it a virtual research environment. The IADB was rewritten as a web-based application in 1999. In recent years, development has continued in collaboration with the Silchester Town Life Project at Reading University. Grants from the Arts and Humanities Research Council have funded the development of the IADB as a web publication tool, and grants from the Joint Information Systems Committee have funded the OGHAM and VERA projects for the development of the IADB as a virtual research environment for archaeology, using the Silchester project as a test-bed. Since 2007, the IADB has been made available as open source software and is now in use with a number of major research projects and commercial contracting archaeological organisations in the UK and Europe. It has also been used as a teaching tool at the University of Reading and in a British Academy-funded project in Iran and Iraq.

Structure of the IADB The IADB is a web application using modern AJAX programming techniques. It is accessed through a web browser, such as Mozilla Firefox or Google Chrome, from any Internet-connected computer. No IADB software is installed on the user’s computer. As stated above, the initial concept of the IADB was that it would be a tool for use in post-excavation analysis, and this remains a primary purpose of the IADB. This means that there are essentially two stages in the use of the IADB, although Integrating database design and use into recording methodologies 81 these overlap to some extent. The first stage is the creation of the digital resource, including the entering of context and find records, the digitisation of single context plans, the uploading of digital photographs, etc. In general, the IADB does not impose any order or timetable on the completion of these tasks. The second stage is the use of these resources to produce the post-excavation analysis of the data. Whilst it is probably true that no two excavation recording and analysis systems are identical, most of those in use in UK archaeology today share a common ancestry. The IADB essentially embodies the architecture and methodology developed by the Museum of London’s former Department of Urban Archaeology, and others, in the early 1980s (MOLAS 1994). At the heart of post-excavation analysis is a process of synthesis, combining individual context records, usually on the basis of a combination of stratigraphic, spatial and artefactual dating evidence, into what the IADB terms ‘sets’ (elsewhere often called ‘features’), then combining these into groups (episodes), then into phases. Although different systems may have more or fewer steps, the result is a hierarchical structure, which in a typical project moves from hundreds or thousands of contexts to perhaps ten or twenty phases. At the end of the process, one or more documents will be produced – the site report. All of these processes are completed within the IADB supported at every stage by the primary digital record. At the broadest level, therefore, the IADB design embodies the recording and analysis methodologies of the projects for which it was developed. It is only when we examine much more detailed aspects of the design and use of particular parts of the IADB that we uncover some of the more interesting challenges and potential conflicts that can arise between database design and recording methodologies. Two recent case studies (one a failure, the other a success) will serve to illustrate this.

FIGURE 5.1 The IADB – part of a stratigraphic matrix and a composite plan. (Copyright: author.) 82 Michael J. Rains

CASE STUDY 1: DIGITAL PENS

The Silchester Town Life Project, based at the University of Reading (www. reading.ac.uk/silchester), is a large-scale, long-running and on-going excavation of part of the abandoned Roman town of Calleva Attrebatum at Silchester, which lies approximately 80 kilometres west of London. The Silchester project has used the IADB as its data management system since the start of the project in 1996, and today it is difficult to imagine a project of the size and complexity of Silchester without recourse to a computerised database such as the IADB. The benefits of the IADB to the Silchester project can be divided into on-site and off-site elements. A broadband Internet connection to site allows the IADB to be used directly for data input and retrieval on site. The benefits of this include:

• Much primary data entry can be completed on site during the excavation. • Queries can be used to retrieve data quickly on finds and contexts. • The site director can use current information to develop the excavation strategy as the season progresses. • The site matrices can be worked on in the IADB as they are deciphered on the ground – this can make understanding some other contexts much clearer as the excavation progresses.

Off site, use of a database such as the IADB, which can capture the written and plan record of each context, plus associated photographs, finds and scientific evidence, has eased the complexity of post-excavation analysis in many ways:

• Archaeological information can be retrieved much more quickly than if it was all kept on paper in a storeroom. • Anyone working on the project can access the data from anywhere with an Internet connection. • It is simply not possible to analyse and report such a complicated stratigraphic sequence without such a database – for example, the IADB provides a tool to organise and represent stratigraphic matrices, which would be difficult, if not impossible, on paper. • Information that is created after the excavation can easily be added to the database later to enhance the record by uploading spreadsheets or documents. This can include further information on environmental samples or photographs. • The collaborative writing and web publication tools of the IADB have enabled the Silchester project to create a number of publication websites as the project has progressed. Integrating database design and use into recording methodologies 83

In addition, Amanda Clarke, the site director of the Silchester field school, has developed the use of the IADB as a teaching tool, both on and off site. All students can potentially access the IADB on site, and thus learn about and contribute to the archaeological record. This provides students with a stake in the accurate recording of the excavation, plus a whole new set of skills to add to their CVs. The IADB has also provided another summative assessment tool, with the development of the IADB Quiz. In 2005, the Joint Information Systems Committee (JISC), part of the UK higher education funding framework, announced funding for a number of projects under its Virtual Research Environments programme (www.jisc.ac.uk/ whatwedo/programmes/vre.aspx). The JISC described the purpose of a virtual research environment (VRE) as helping ‘researchers in all disciplines manage the increasingly complex range of tasks involved in carrying out research’ by providing a ‘framework of resources to support the underlying processes of research on both small and large scales’ (JISC 2007). In practice, this translates into providing access, which in current circumstances means online access, to the resources (data) and applications (tools) necessary for research. Although not stated explicitly, the concept of collaborative working is also central to the JISC’s definition of a virtual research environment. The Silchester project was awarded funding by the JISC for an initial two-year phase under the acronym OGHAM (Online Group Historical and Archaeological Matrix). In 2007, a further two years’ funding was awarded for a continuation project entitled VERA (Virtual Environments for Research in Archaeology). One strand of both the OGHAM and VERA projects aimed to improve the flow of data or information throughout the lifetime of the archaeological project, including initial data collection and recording (Rains 2008). As part of this, extensive trials were undertaken in the use of digital pens to speed up the digital capture of site context records and their accessioning into the project database. Previously, these on-site records had been written onto pro-forma context recording sheets, and manually transcribed into the IADB later, normally as part of the post-excavation process. While the digital pens were found to perform well from a technical point of view, it was not clear that their use produced any significant time saving with regard to the production of digitised context records. This was particularly true when the extra resources allocated to the trials were taken into account. In fact it can be argued that the digital pens did not address the real problem. The amount of actual text on a typical context recording sheet is relatively small – it would rarely exceed one hundred words – and so the transcribing of this text was never the most time- consuming part of the process. Another problem was that for many years the Silchester project, in common with most other excavations, has used hand-completed pro-forma context recording sheets (CRSs). Over time, a comprehensive system has been 84 Michael J. Rains

developed for the management and checking of these forms. For various reasons, the digital pens trial opted not to use digital versions of these forms, but to use free-form digital notebooks. This fundamental change to the recording system caused many problems for the management and checking of the data. For example, while it was easy enough to check what had been recorded in the digital notepads, it was much more difficult to check what had not been recorded, whereas empty fields on the context recording sheets were easy to spot. This highlighted the central role of the traditional CRS in excavation recording, to the extent that it would require a brave (or foolhardy) project director to suggest getting rid of it. In developing any new system or application, it is important to keep in view the reason for doing it. What is the problem we are trying to address? Does the solution actually resolve it? With any new technology, hardware or software, it is easy to become blinded by its newness and lose sight of the original reason for introducing it. The problems highlighted above started to become apparent soon after the initial introduction of the digital pens on a trial basis at Silchester, and as a result the trial was not extended and use of the pens was discontinued after two seasons. If there had been clear potential savings in money, time and effort for the Silchester project, and clear benefits for the archaeologists expected to use the pens, there would probably have been more effort to persevere with resolving the organisational, training and implementation issues. In fact, the digital pens were seen as a novel nuisance that site supervisors found difficult to integrate into their established working practices, and from which they saw little or no benefit to themselves or to the wider project. The digital pen trials highlighted an important issue with regards to the introduction of new techniques and technologies into long-running systems such as archaeological excavation projects with well-established methodologies. Any change or development has to be a real solution to a real problem. It will only be accepted if researchers feel that the solution offered truly addresses a need that they themselves perceive. In other words, it is not enough for the solution to address the problem adequately (although it has to do that); the problem itself has to be a real one. Integrating database design and use into recording methodologies 85

CASE STUDY 2: THE FACSIMILE CONTEXT RECORD

As was suggested above, the context record, usually in the form of the context recording sheet (CRS), lies at the heart of virtually all current approaches to excavation recording. Over recent years, this has come to be reflected in the importance of the digitised context record in computerised excavation records created in systems such as the IADB. Within the IADB, the format of the context record has remained largely unchanged since the early days of the system, when it was designed as a full digital transcription of all information from the CRS. It is not a replica of any particular CRS but is similar to many in use in the UK. However, transcription of the entire CRS into the IADB context record is a time- consuming and expensive process, and tends to add to the delay between the completion of the CRS on site and the availability of context data within the IADB. When this is combined with the financial and time pressures of modern contract archaeology, or the online access requirements of large research projects, there is an increasing need for the rapid, accurate creation of a digital context record. It can be argued that much of the transcribed CRS data is of questionable utility (at least in transcribed digital form) in post-excavation analysis and research. Little of it will ever be used to index or search the database; in fact, little of it will ever be referred to again. In practice, the transcribed data is invariably an interpretation of the original CRS record rather than an accurate transcription of it. The IADB does not contain the original CRS record, and so the immediate connection between the IADB record and the original CRS is lost. Although in theory it may be possible for post-excavation analysts and researchers to gain access to the original CRS record, this opportunity is rarely used. In the light of these points, a new approach to the IADB context record was developed, which in simple terms replaces the full transcription of the paper CRS with a scanned image, or facsimile, of the original CRS, plus some basic transcribed metadata. A key consideration in implementing this approach to the digital context record is selecting the appropriate metadata to transcribe. In doing this, we need to keep in mind the original purpose of the IADB as a post-excavation tool that provides access to the complete integrated excavation record. Therefore, the metadata needs to be sufficient to support the normal post-excavation uses of the IADB. From an access point of view, the metadata needs to include sufficiently detailed keywords to support searching, indexing and classification. Stratigraphic and other relationships are essential and provide the linkages within the database that define the integrated aspect of the IADB. However, long textual descriptions of the composition or interpretation of the context are 86 Michael J. Rains

FIGURE 5.2 A facsimile context record in the IADB. (Copyright: author.)

not required – these can be read from the facsimile image. It can be argued that, by creating a restricted range of digital metadata, the transcribed text of the detailed context descriptions and interpretations will not be available for later re-use, such as for copying and pasting into reports, etc. However, in IADB terms, such re-use of database text is most likely to be at the set, group, phase or object level rather than at the context level. In fact, it is likely that such a small percentage of text from all context records would ever be used in this way that it would almost certainly be more efficient to transcribe it, if and when required. Consideration needs to be given to the situation in which there is a discrepancy in the IADB database between the facsimile and the metadata. For example, during metadata entry, it might be noticed that one or more of the stratigraphic relationships recorded on the previously scanned CRS is wrong. One solution would be to correct the CRS and re-scan it. However, this would be time consuming and could potentially lead to loss of the digital record for the original, uncorrected CRS. Another approach would be to allow the discrepancy with the understanding that the metadata overrules the facsimile. An alternative solution has been developed that enables the addition of annotations, in the style of ‘post-it’ notes, to the scanned facsimile. The facsimile mechanism needs to cater for both single and double-sided CRSs in both portrait and landscape format, all of which are in widespread use, as well as for the uploading and display of scanned continuation sheets. Uploading of individual scanned images will be required in certain situations, but this approach is very slow and inefficient when a large number of context records needs to be processed. Therefore, a batch upload facility, linked to the Integrating database design and use into recording methodologies 87

use of a high-speed scanner with a sheet feeder, has been developed. As each facsimile image is uploaded, it is displayed to the user alongside fields for entering the required metadata for that context. This has proven to be a quick and easy-to-learn method for uploading large numbers of context records into the IADB, and the same technique has now been extended to the batch uploading of photos to the IADB. As part of the process of designing and implementing the facsimile-based approach to context recording, the opportunity has been taken to rethink and hopefully improve the layout of the IADB context record. In particular, the introduction of more hyperlinks to enable the launching of keyword searches directly from within a context record provides for improved data integration, which is something that has always been at the heart of the IADB. The opportunity has also been taken to extend some of the new design features to other core IADB records, including finds, sets, groups, phases and objects. From its outset, the development of the facsimile approach has been driven by user needs and feedback. User needs analysis carried out as part of the VERA project highlighted the central role the context recording sheet plays in modern archaeological excavations. The smooth running of any archaeological exca- vation depends on a system where data recorded on the CRS can be effectively managed and manipulated. The facsimile approach appears to offer a streamlined solution to the issue. In order to evaluate how successfully the new system fitted the needs of its users, responses were gathered from some of the people using it. The facsimile approach to context recording has been in use by York Archaeological Trust (YAT) (www.yorkarchaeology.co.uk) and Canterbury Archaeological Trust (CAT) (www.canterburytrust.co.uk) for over three years now, and has been deemed a great success at both organisations.

The differing outcomes of the two case studies described above touch on some of the key considerations of the relationship between database systems and recording methodologies. We can identify at least three significant factors that contributed to the relative failure of the digital pens trial. First, they abandoned the traditional context recording sheet; this was revolutionary rather than evolutionary, and was a step too far for most users. Second, the people who were expected to use the digital pens were not the people who would gain if the trial were successful: it was primarily the site supervisors and their assistants at Silchester who were expected to use the pens on site, whereas it was in post-excavation that any benefit would be felt by avoiding the need to digitise CRSs. The on-site workers therefore had little stake in the new system and no incentive to make it work. Third, and at least partly as a result of the first two factors, the trial could not be shown to have saved money, time or effort, or to have produced a fuller or more accurate digital record. 88 Michael J. Rains

In fact, it could be argued that the score was negative on all of these. On a tightly funded research excavation, the money spent on the trial could probably have been better spent elsewhere. It is interesting to note that none of the reasons for failure were technological – the digital pens performed exactly as expected and, in fact, proved remarkably robust in the harsh environment of an archaeological exca- vation. The key reasons for the relative success of facsimile context recording are in many ways the opposite of the reasons for the relative failure of the digital pens trial. First, the introduction of facsimile context recording required no significant change in the well-established on-site recording methodologies. The CRS is scanned and its metadata added off site by the same people who would previously have transcribed the CRS into the IADB. The fact that a clear saving in time and effort could be demonstrated in what most people would agree can be a boring and repetitive task gave users a great incentive to make the system work even in the face of initial teething problems such as jammed scanners. Improvements brought about by the introduction of facsimile context recording have not been at the cost of de-skilling the job – the same level of archaeological knowledge and experience is required to select the correct metadata for a context as to interpret and transcribe a full context record.

Publication and dissemination Over recent years, there has been a considerable movement in archaeology towards the publication of excavation reports, research papers and other documents on the web. At the forefront of this movement in the UK has been the peer-reviewed web-based journal Internet Archaeology (http://intarch.ac.uk). This move has probably been driven largely by considerations of cost, but it has allowed authors and publishers to explore the enhanced functionality, particularly interactivity, offered by web publication. The Silchester project has aimed since its beginning in 1996 for full publication of both data and reports throughout the life of the project, and not just at the end (Clarke et al. 2004). In recent years development of the IADB has concentrated in this area of web publication with the emphasis on the integration of report publication and database publication. Facilities have been developed to allow what might be termed IADB ‘resources’ (for example, plans, photographs, context records, finds lists, stratigraphic matrix diagrams, etc.) to be linked to or embedded directly within report documents, which are authored directly within the IADB and published directly to the web. Recent examples of this include reports from both the Silchester project (Fulford and Clarke 2010) and York Archaeological Trust (McComish 2010). Canterbury Archaeological Trust has recently launched a searchable online map-based gazetteer of all its projects in the Kent area, built directly within the IADB (www.iadb.co.uk/i3/item.php?ID=CAT:GAZ). In all of these cases, the reader or user is able to ‘drill down’ through the various synthetic analytical levels of the IADB to primary data, such as individual context and find Integrating database design and use into recording methodologies 89 records. In fact, each IADB project, each tagged list, each individual resource record (find, context, image, structure diagram, etc.), and each SQL query in an IADB project database has a unique web address or URI which allows links to individual IADB resources to be included in web publications, external databases, etc. As well as requiring a technical solution, such as the IADB, to make this integrated database/report publication possible, these developments also place the onus on archaeologists to ensure that their recording methodology, including its checking processes, produces publishable basic records. This means records that are not only complete and accurate but that also adhere to a certain professional standard, in terms of grammar and style, for example.

Curating the archive In the early years of the IADB, it could have been argued that, as it was a closed system designed purely to assist in post-excavation, it was only necessary to digitise those parts of the primary excavation archive that would be directly useful to the post-excavation analyst. However, as the scope of the IADB has widened and has become a web application, it can now be seen as the primary archive of a project. The online archive (in the IADB) becomes the definitive archive. Very few people will ever refer to the primary paper archive, mainly because distance makes that impossible for most people, but even those few for whom access to the primary paper archive is possible are unlikely to make the extra effort involved given that they can simply engage with the online archive. This means that, to all intents and purposes, if something is not in the IADB, then it does not exist – and whatever it says in the IADB is the truth, even if it is not. This puts a great burden on archaeologists to ensure that, however they are creating it and whatever its structure and functionality, their digital archive is both as complete and as accurate as possible. They also have a responsibility to ensure that both the data and its interface survive. Many of the issues surrounding long-term data archiving and preservation – the use of metadata, the selection of appropriate archival file formats, the durability of storage media and questions of technological obsolescence, to name but a few – are now well appreciated and understood, if not completely resolved. This is in large part due to the excellent work over recent years of bodies such as the UK Archaeology Data Service (http://archaeologydataservice.ac.uk). On the other hand, software sustainability, if considered at all, has tended to be viewed mainly at the infrastructure level, rather than at the application level of the user interface in systems such as the IADB. Infrastructure frameworks on which web applications are built include LAMP (Linux, Apache, MySQL and PHP, as used by the IADB) and the Microsoft .Net framework. At the top of any such framework lies the actual web application. Neither the digital pens trial nor the introduction of facsimile context recording required any major changes to the IADB database structure – rather, both were developments of the functionality of the system. This emphasis on user interface, 90 Michael J. Rains and particularly its functionality (rather than just its look and feel), over recent years has led to a position where it is increasingly difficult to separate data from interface. For any IADB project, a simple dump of the raw data is of much less utility and value than the same data accessed through the IADB interface. Indeed, many of the ninety-plus tables within the IADB database are there only to support the functionality of the interface, and have little intrinsic meaning or value when taken out of that context. Examples of this include the multi-context plan, created in real time from a combination of spatial and stratigraphic data, and the stratigraphic hierarchy diagram, which combines stratigraphic and context attribute data with other IADB data (in this case images and plans). These are both interface constructs or visualisations of a range of data that would have little immediate value or meaning if it were only available as the output of an SQL query dump. In these cases, the whole truly is greater than the sum of its parts. It can be argued, therefore, that simply archiving the IADB database into a series of CSV files, for example, does not adequately address the problem, although it might be part of the solution. York Archaeological Trust has digital records dating back over forty years. It is noteworthy that those that remain most accessible today are the ones that have never been formally archived, but, because they are part of the larger organisation- wide database (the IADB), they have been carried forward through each successive hardware and software migration and, most importantly, interface development. For example, basic find records from a 1972 excavation in York that have never been archived are still instantly available through the IADB, whereas other records that, with the best of intentions, were ‘archived’ to floppy disk or tape are now largely inaccessible or indecipherable. This may be the result of technological obsolescence, media degradation or even a simple lack of adequate metadata. It can be argued, therefore, that instead of preservation through formal archiving, the aim should be to preserve through continuous refreshment of the data. It can also be argued that preservation of the interface is just as important as preservation of the data and its metadata, which the interface in many ways embodies and represents. In fact a continuous process of interface development will lead to a kind of virtuous circle, in that maintaining and developing the interface will ensure that all necessary data refreshment and migration takes place. How this long-term maintenance and development of the interface is to be managed and funded is not clear.

Conclusion The case studies described above, along with the work on web publication of archaeological projects and the questions concerning archiving and preservation of archaeological resources, make clear the need for true integration of the design and development both of excavation and recording strategies and methodologies, and of database structures and functionalities. It is clear that the design and structure of the database in terms of tables and records must reflect closely the conceptual structure implicit within the recording methodology. Just as fundamental though, Integrating database design and use into recording methodologies 91 and probably more difficult, is to ensure that the functionality and procedures of the database system mesh well with the methodology and procedures of excavation and recording and analysis. Just as it has long been recognised within the IADB that the links between records are at least as important as the records themselves, so it is the links between, the points of interaction between – the integration of – database and recording methodology that will define the success or failure of the system as a whole.

Acknowledgements With thanks to Michael Fulford and Amanda Clarke of the Silchester Town Life Project at the University of Reading; Claire Fisher, formerly of University College London; Peter Clark of Canterbury Archaeological Trust; and Martin Stockwell and Peter Connolly of York Archaeological Trust.

References Clarke, A., Fulford, M. and Rains, M. (2004) ‘Nothing to hide – online database publication and the Silchester Town Life Project’, in M. Doerr and A. Sarris (eds) The Digital Heritage of Archaeology: Computer Applications and Quantitative Methods in Archaeology, Proceedings of the 30th CAA Conference, Heraklion, Crete, April 2002, 401–404. Fulford, M. and Clarke, A. (2010) ‘Silchester: the city in transition – the mid-Roman occupation of Insula IX, c.AD125–250/300’, a report on excavations undertaken since 1997. Available at: www.silchester.reading.ac.uk/cit/index.htm. Accessed 17 September 2013. JISC (2007) Virtual Research Environments Programme. Available at: www.jisc.ac.uk/what wedo/programmes/vre2.aspx. Accessed 17 September 2013. McComish, J. (2010) ‘Roman, Anglian and Anglo-Scandinavian activity and a medieval cemetery on land at the junction of Dixon Lane and George Street’, The Archaeology of York, Web Series AYW9. Available at: www.iadb.co.uk/i3/item.php?ID=IADB:1307: D3476. Accessed 17 September 2013. MOLAS (1994) ‘Museum of London Archaeology Service “Archaeological Site Manual”, 3rd edn,’ Available at: www.museumoflondonarchaeology.org.uk/NR/rdonlyres/ 056B4AFD-AB5F-45AF-9097-5A53FFDC1F94/0/MoLASManual94.pdf. Accessed 17 September 2013. Rains, M. (1995) ‘Towards a computerised desktop’, in J. Huggett and N. Ryan (eds) Proceedings of the Twenty-Second CAA Conference held at Glasgow University, Glasgow, 1994. Oxford: BAR International Series. Rains, M. (2008) ‘Silchester – a virtual research environment for archaeology’, in A. Posluschny, K. Lambers and I. Herzog (eds) Layers of Perception: Proceedings of the Thirty-Fifth International Conference on Computer Applications and Quantitative Methods in Archaeology (CAA), Berlin, Germany, 2–6 April 2007. 6 THE TYRANNY OF TYPOLOGIES

Evidential reasoning in Romano- Egyptian domestic archaeology

Anna Lucille Boozer

. . . logical typologies are not always useful, and useful typologies are not always logical. (Adams and Adams, 1991: 8)

While excavating in an underexplored region of Egypt, my project encountered the archaeological remains of a house that did not correspond to the dominant domestic typologies for Roman Egypt. For example, our structure lacked clear archaeological evidence of a roof over a central room in the house, despite a high degree of preservation elsewhere in the structure. We found evidence of food preparation within the house, rather than in an exterior courtyard. The house also appeared to have a single story rather than two stories. We were left with a dilemma. Did this evidence – and lack of evidence – signify a new house type? Or should we question our archaeological evidence because our structure differed from pre-established house typologies? These debates brought to the fore a long-standing issue in archaeology: how and when should we use, modify, and abandon typologies? And how does academic discourse impact our use of typologies and archaeological evidence? This contribution explores these questions in order to understand how archae- ologists weigh competing clues between typologies, taphonomies, and depositional episodes in our evidentiary reasoning. I begin by addressing typologies within archaeological thought, particularly focusing on architectural rather than artifactual typologies. Typological analysis has been a long-standing component of the archae- ological tool kit and both pragmatic and theoretical approaches to this analysis have changed over time; therefore, a review of the history of typological analysis situates the current predicament. I then draw from some recent critiques of typological analysis that are relevant to the present inquiry into the use and abuse of typologies in Roman Egypt. In the following section, I focus upon Karanis, a Romano- Egyptian village located in Egypt’s Fayum region, as a case study for the tyranny The tyranny of typologies: Romano-Egyptian domestic archaeology 93

Nonjudgmental Nonjudgmental Nonjudgmental Nonjudgmental NonjudgmentalNonjudgmental

Nonjudgmental

Nonjudgmental Nonjudgmental Nonjudgmental

NonjudgmentalNonjudgmental Nonjudgmental Nonjudgmental Nonjudgmental NonjudgmentalNonjudgmental

Nonjudgmental Nonjudgmental

NonjudgmentalNonjudgmental

Land above 500 m Land below sea level 0300 kms Roads Land above 200 m National boundary

MAP 6.1 Map of Egypt (M. Matthews). of typologies (Map 6.1). Karanis has long dominated perceptions of Romano- Egyptian housing traditions, and it is necessary to re-examine both Karanis itself and its role within academic discourse to explain why this dominance is undesirable. Finally, I draw together these inquiries in order to assess a potential way forward for the use of typologies within Romano-Egyptian domestic archaeology, as well as more broadly.

The tyranny of typologies The practice of archaeology is paradoxical: it draws from partial, complex, and ambiguous evidence, but it is also tangible and “real.” This paradox entails that 94 Anna Lucille Boozer major discoveries in archaeology can include both unexpected material evidence as well as reassessing theoretical assumptions within contemporary archaeological thought (Wylie 2011: 372–373). Archaeological writings often conceal this paradox by erasing the ambiguities inherent in tangible archaeological evidence (Gero 2007). In this contribution, I focus on one way in which archaeologists conceal this evidentiary paradox – what Joan Gero calls cleaning the data – in order to make data appear homogeneous (Gero 2007: 320). One common way in which archaeologists clean their data is by classifying house forms, objects, and social practices (Costin 2007: 144–145) so that the data can be homogenized. This practice involves the use of typologies, which stabilize the assumption that there are homogeneous meanings, forms, and types across different time periods and geographic terrains. Typological analysis can be understood as the “systemic arrangement of material culture into types based on similarities of form, construction, decoration or style, content, use, or some combination of these” (Aldenderfer 1996: 727). Typologies serve the process of archaeological concealment inasmuch as they can erase differences between uncertain evidence through classification. Despite drawbacks associated with their application, typologies are a vital and long-term component of archaeological practice (see Whallon and Brown 1982; Cowgill 1990; Adams and Adams 1991). The primary concern of this chapter is to explore how typological analyses are used and abused within archaeological thought.

Typological analysis: functionality and tyranny It is worthwhile exploring the history of typological research more deeply in order to understand how typological analysis became entrenched within archaeology. Typologies are purpose-built. As a heuristic device, typologies can serve two different functions within various stages of archaeological research. First, in the early stages of research, typologies make it possible for archaeologists to describe messy data concisely and give it some order. This order can take shape as chronological, as well as formal, functional and cultural, groupings. Second, in later stages of research, typologies provide a material context for new finds, and they function as a crucial medium for communication between researchers. This stage, and the communicative shorthand produced, is where the tyrannical aspect of typologies becomes most evident. This section reviews how these two major phases of typological analysis developed in archaeological thought over time.

Functionality: early stages of research and present-day uses Classificatory phases usually occur fairly early in the history of every discipline. There comes a point at which the haphazard accumulation of raw data demands a major effort at systematization before any further progress can be made (Adams and Adams 1991: 266). The primary function of archaeological classification in the early days (c. 1880–1920 for North America) was to predict dating and interpret The tyranny of typologies: Romano-Egyptian domestic archaeology 95

“culture” (Adams and Adams 1991: 310–311). These typologies helped archaeolo- gists to identify patterns and diachronic changes, as well as make the most of partial data sets. As such, typological analysis was an essential component of archaeological research before the advent of absolute dating techniques and was employed globally in archaeology by prescient pioneers such as Oscar Montelius,1 Flinders Petrie, Nels Nelson, and A. V. Kidder. It is worth highlighting V. Gordon Childe’s publications of the 1920s and 1930s. Childe was an early proponent of culture-historical archaeology and Marxist archaeology. In particular, he was known as “the Great Synthesizer” for his efforts to synthesize regional typologies into a broader understanding of Near Eastern and European prehistory. The regional typologies that Childe developed were strongly grounded in a spatial dynamic rather than purely a temporal one (see Childe 1929). Although the old typological objectives remained, these loftier goals began to take root within archaeological thought. By the 1940s, archaeologists began to look to artifact types in order to understand belief and behavioral systems also. The so- called “Typological Debate” arose at this time because archaeologists were uncertain if previously established typologies could be adapted to new purposes, such as answering questions about social life (Adams and Adams 1991: 310–311). The Typological Debate raged between major archaeological figures of the 1930s, 1940s, and 1950s.2 One of the most critical questions of this debate was whether typologies were a priori or if they were arbitrary constructs created by archaeologists to address specific contexts encountered in the archaeological record (Wylie 2002b). Spaulding, in particular, argued that typologies were inherent in the archaeological material and could be subjected to rigorous statistical analyses (Spaulding 1953). One of the most persistent critics of conventional typologies was Lewis Binford, who did not agree with such statements (e.g. Binford 1972: 195–207, 252–294). Binford’s work can be characterized as strongly anti- metaphysical and scientifically oriented. Binford’s “New Archaeology,” as espoused in An Archaeological Perspective (1972), exposes an archaeological positivist who characterized his view of archaeology as a struggle against “traditionalists” such as Robert Braidwood, James Griffin, and (notably) Albert Spaulding. Despite this phase of archaeological anxiety about typologies, they remain a cornerstone of archaeological practice today. Most current archaeologists under- stand typologies to be instruments for measurement and for testing hypothesized relationships. Typologies are designed to measure what we believe are the relevant set of relationships between variables or attributes, helping archaeologists to test specific hypotheses. Typologies of this kind arise from procedural steps of logical thinking and are designed for specific purposes. In this sense, typologies are not used to define types by induction only, but instead represent one of the most important analytical steps we traverse when we evaluate the quality of our analytical process and theoretical outcomes. Evaluation is vital for determining how well our theories account for our data, and they are also critical to theory building. Ideally, we should formulate our assertions, deduce new hypotheses, and construct new typologies to measure the 96 Anna Lucille Boozer hypothesized relationships. Used in this way, typological analysis takes on a more interactive role in the feedback loop between method and theory. Such feedback is necessary because theories are never complete and require constant reformu- lation. Theory building is cumulative rather than a simple one-step process (Vierra 1982: 173). Today, typologies also serve as pragmatic tools of communication among archaeologists. Useful typologies require consistency, and most archaeologists recognize that consistency is different from objectivity or accuracy. We will never know how closely our type concepts correspond to some external reality, but we can discover and measure how closely the concepts of one person correspond to those of another (see Fish 1978; Ziman 1978). There is no absolute right or wrong way to classify anything, but there are better and worse ways of pursuing specific purposes, once we have decided what those purposes are. Many archaeologists understand “better” to be more precise, more communicable, and more affordable (Adams and Adams 1991: 4).

Typological tyranny: critiques of typological analysis Although there are clear benefits to typological analysis, this traditional practice is more of an art than a science. Moreover, it is noteworthy that the “Typological Debate” raged almost entirely at the level of theory. In many regions and for many purposes the tangible practice of artifact classification operates today much as it did fifty years ago (Adams and Adams 1991: 266). For example, Ian Hodder’s famous volume The Domestication of Europe successfully re-examined old data, based upon old typologies, through the lens of new theoretical vantages (Hodder 1990). Other archaeologists have followed suit, leaving the old typologies unexamined. Unfortunately, the intuitive approach most archaeologists have employed in typological analysis lacks scientific rigor and introduces arbitrary emphases in the choice of attributes singled out as diagnostic. The construction and use of typologies can never be an automatic or a wholly objective process. Typologies are created to serve human purposes, and these purposes strongly affect the ways in which they are made and used (Adams and Adams 1991: xvi). Further scrutiny of human factors would be worthwhile. For example, archaeologists are affected deeply by their own experiences excavating sites, which entails that they “read” new sites through the lens of old sites, a process that Bell calls xeroxing (Bell, Chapter 3 this volume). Moreover, the timing of discovery may entail that sites discovered early on in a region may have an undue influence on the interpretation of subsequent dis- coveries. A peek behind the typological curtain would reveal more about the procedural thinking hidden in archaeological publications. Although the spatial and temporal ordering of data is not (is not intended to be, and should not be mistaken for) explanation (contra Fritz and Plog 1970), the analytical steps involved in typology creation and communication are not always made clear in excavation reports and other publications. This lack of transparency entails that archaeologists can move The tyranny of typologies: Romano-Egyptian domestic archaeology 97 smoothly between knowns and unknowns in a disconcerting fashion in their published accounts of excavations. Returning to Gero’s exploration of data cleaning, there are two ways of data cleaning that are relevant to this critique of typological analysis: (1) stretching the data; and (2) providing visual focus. First, stretching the data involves finding the general in the specific, or making sweeping claims and conclusions on the basis of specific cases and sites (Gero 2007: 321). In the case of typologies, stretching of data entails erecting typologies for regions and time periods from single sites, thereby “demonstrating” a wider pattern or process. The problem with this approach is that data stretching can eventually become so entrenched in scholarship that it is difficult for scholars to realize that it is problematic. For example, some attributes may appear to be significant on the basis of an assemblage discovered early on in a region before broader statistical patterning could be accomplished. Early discoveries often have undue influence upon later typologies and patterns identified by scholars, and this process must then be undone by careful scholarship to peel back the layers of misattribution (see Bowes 2010: 12–13; Ruffini 2012). For example, Allison’s groundbreaking article in 1999 argues that labels mapped from texts onto archaeological material led to many decades of misunder- standing and misinterpretation (Allison 1999). Her careful study of archaeological material from a modern, scientifically grounded standpoint allowed for more rigorous interpretations of Pompeiian households. This painstaking scholarship often results in significant paradigm shifts within local arenas in archaeology. Second, providing visual focus entails the redundant use of visual templates as graphic displays (Gero 2007: 321–322). Visual attention focuses upon specific images so frequently that ambiguous outlying or competing interpretations and images become obliterated and ignored. Over time, scholars begin to limit their analytical spectrum and confuse dominant images with reality. For example, Eckhardt notes that many papers in her volume on Roman diasporas refer to the Regina and Barates tombstones from South Shields, denoting the limited examples that are repeatedly re-examined (Eckardt 2010: 8). The Romano-Egyptian site, Karanis, has become emblematic of these two facets of data cleaning and returns us to the problem introduced at the beginning of this chapter. By virtue of its high degree of preservation, the house plans and artifacts from this site are often used to illustrate “typical” daily life in Roman Egypt (visual focus). Moreover, by virtue of its excavation timing, Karanis house forms have become the “standard” type of house found in Roman Egypt (data stretching). One now has to “prove” that excavated houses do or do not look like Karanis houses, since this one house type has come to dominate domestic archaeology in Roman Egypt.

The case of Karanis Karanis is the most dominant example of an over-cited domestic site from Roman Egypt. Many scholars consider Karanis to be a well-understood site within Roman 98 Anna Lucille Boozer

Egypt and frequently employ Karanis evidence in order to fill in knowledge gaps. Unfortunately, there are two major problems with this approach, which are relevant to the generalities surrounding typological debates: (1) the way in which scholars use the data is often unexamined; and (2) the data itself is less secure than many scholars may realize. Despite these concerns, archaeologists today must rigorously demonstrate differences between their discoveries and Karanis for differences to be believed.

Karanis within academic discourse Typologies make it possible to build up “conjuctions” because they are a repository of interpretive insight (at their best), and the basis for producing an entire domain of second-order data, such as facts about the relationships between types of material and their distribution, comparisons, and patterning in the archaeological record (see Taylor 1983 [1948], on conjunctive archaeology). Unfortunately, the initial description and systematization phase of data can itself be problematic. This initial phase can be partial, impressionistic, heavily biased and even empirically false. Problems with an initial description can have palpable downstream effects on archaeological thought. The Romano-Egyptian town site, Karanis, serves as an excellent illustration of the ways in which ill-conceived and empirically inadequate typologies can undermine future work. Karanis appears in academic literature as the Romano- Egyptian domestic type site, both implicitly and explicitly. These issues relating to the dominance of Karanis are separate but linked and dependent upon the way in which scholars employ Karanis in their arguments. First, some authors implicitly exploit Karanis as a type site. This implicit use of Karanis is particularly common among scholars making an argument that primarily employs papyrological and historical evidence rather than material evidence. In this respect, Karanis serves as an illustrative device, rather than an examined example or case study. Images, plans and brief references to the site often accompany arguments asymptotic to the materiality of the site itself. This implicit use of Karanis provides visual focus because it involves the redundant use of visual templates as visual displays (see Gero 2007: 321–322). Visual attention returns to Karanis in an almost subconscious manner that reduces the ambiguity of outlying or competing interpretations and images. Specific images from Karanis are often employed as illustrative devices (e.g. Malouta 2012: 295). For example, the same photograph of a passageway in front of Karanis house C123 was employed in the two most significant summative accounts of Roman Egypt (Bagnall 1993: cover photo; Bowman 1986: Figure 92). Likewise, the same image of a courtyard with milling equipment from House C113 is reproduced in two significant Roman Egypt volumes (Bowman 1986: Figure 60; Rowlandson 1998: Figure 22b). Wilfong recently published a small selection of the numerous unpublished images of Karanis and argues that these images give us a more complex and somewhat different perspective on Karanis than the common assumptions, which are based The tyranny of typologies: Romano-Egyptian domestic archaeology 99 on the frequently published corpus of images (Wilfong 2012). Wilfong’s publica- tion underscores the need to re-examine extant images of these early excavations in order to create fresh insights into Karanis. Second, authors employ Karanis explicitly as a typological example. These studies often use Karanis as “filler” when desirable archaeological evidence is lack- ing. This scenario is particularly common among archaeologists and scholars dis- cussing archaeological data. The case of Hermopolis Magna, which was destroyed by digging of various types, serves as a good example of this practice:

No private house of the Roman period has been recorded archaeologically at Hermopolis Magna, although hundreds must have been cleared away by sebakh diggers, papyrus hunters and excavators. We must imagine them to be like those excavated at Karanis, several with many stories, the lower rooms often falling out of use with rising road level. (Bailey 2012: 195)3

The suggestion that Hermopolis Magna once bore similarities to Karanis was not based upon material evidence, but upon the dominance of Karanis as a domestic type site. Statements such as the above entrench Hermopolis Magna in arguments that the Karanis house forms were pervasive across Roman Egypt. Gero would consider this second issue with the use of Karanis to be stretching the data, that is, finding the general in the specific, or making sweeping claims and conclusions on the basis of particular cases and sites (Gero 2007: 321). This stretching of data entails that scholars “demonstrate” a wide pattern of a particular typology without proper data scaffolding to support such a demonstration. These two approaches to Karanis domestic architecture illustrate the ways in which the site has become enmeshed within current Romano-Egyptian domestic understandings. As a result of Karanis exemplifying the “norm,” we are left with a long history of academic research that replicates problems with the site (cf. Allison 1999). Indeed, the site itself is not without problems of its own, which makes it unstable ground upon which to erect arguments.

Karanis background The University of Michigan excavated two of the most famous Fayum sites in the early twentieth century: Karanis (Kom Aushim) (1924–1934) and Soknopaiou Nesos (Dimê) (1931–1932). Karanis has become the type site of Romano-Egyptian domestic architecture, due to its high degree of preservation and the care with which it was excavated compared to previous work on Romano-Egyptian houses. Since the 1930s, Karanis has appeared in numerous publications as a representative Romano-Egyptian settlement. By contrast, Soknopaiou Nesos is less familiar to the Romano-Egyptian community of scholars. Additional houses from other Romano- Egyptian sites are virtually unknown to many scholars of Roman Egypt. Generally, it seems that buildings at Karanis were aligned into blocks of habita- tion (see Figure 6.1; see also Husselman 1979: 10, Maps 4–5). Although hundreds 100 Anna Lucille Boozer of buildings were excavated, only a few were described and drawn. These houses were made predominately of mud brick with only small amounts of wood used (Husselman 1979: 34). Flat roofs were common, except in cellars, which were vaulted. Most houses were elongated and had multiple stories, with cooking taking place in courtyards that were either private or shared. Often, the walls were plastered and covered with a thin lime wash. A black wash was most common, with white accents painted horizontally across the mud brick courses. Decorations were minimal and were usually found in niches; they often represented religious scenes, and were typically painted in maroon and black (Husselman 1979: 35–36, plates 18, 19a, 21b, 22a, 24a, 24b, 25, 102a).

Fun

Fun Fun Fun

Fun Fun Fun Fun

Fun Fun

Nonjudgmental

FIGURE 6.1 Karanis insula. Plan VI, Section F10 in A.E.B. Boak and E.E. Peterson (1931) Karanis: Topographical and Architectural Report of Excavations during the Seasons 1924–28, Kelsey Museum of Archaeology, University of Michigan. (Courtesy of the Kelsey Museum.) The tyranny of typologies: Romano-Egyptian domestic archaeology 101

The University of Michigan also excavated houses at Soknopaiou Nesos, but the results of this mission were not fully published.4 Only the coins, papyri, and specific architectural elements received attention in the single, slim, published excavation report (Boak et al. 1935). The houses appear to have been built contiguously, and each had a courtyard to support domestic cooking needs. These structures were built directly onto bedrock. The largest house excavated had an internal courtyard (aithrion). All of the structures had a central pillared stairway leading to the upper floor(s), and often also to cellars. Poor-quality wall paintings were found in some of the structures (Boak et al. 1935: 9–10, Figures 4, 6, 7). There are three major caveats that must be kept in mind when employing Karanis houses as comparanda.5 First, one of the primary reasons why the University of Michigan archaeologists began excavations at Soknopaiou Nesos is that they were confused by the stratigraphy at Karanis (Boak et al. 1935: v). As a result, the Michigan archaeologists excavated more carefully at Soknopaiou Nesos than they had in the eight years prior at Karanis. They hoped to improve their understanding of the geology and stratigraphy in the Fayum through more careful excavations at Soknopaiou Nesos. The Michigan team only returned to Karanis for two seasons after refining their methodologies at Soknopaiou Nesos. Sadly, stratigraphic issues at Karanis were never resolved, and to this day it is uncertain how the phasing of the site should be understood (Wilfong 2012: 228). Second, the Karanis houses have never been published fully. General reports on the stratigraphy, topography, and architecture were published in the 1930s, but these lacked full analyses of each structure and did not attempt to interpret the findings (cf. Boak and Peterson 1931; Boak 1933). A subsequent report attempted to fill in the gaps left by these prior publications (Husselman 1979) but it did not provide contextual explorations of the material, and the accompanying maps and illustrations are difficult to connect to particular structures and occupational phases. Third, the publications that do exist for Karanis houses analyzed material categories rather than contexts (e.g. individual houses). Many of the objects from the excavations went to the Kelsey Museum at the University of Michigan in Ann Arbor. Over the years, specialist publications on exhibitions and particular categories of material have appeared (e.g. Gazda and Hessenbruch 1978; Thomas 2001; Haatveldt and Peterson 1964; Harden 1936; Higashi 1990; Husselman 1952, 1953, 1971; Johnson 1981; Shier 1978; Yeivin 1934; Youtie and Pearl 1939, 1944). Unfortunately, we still do not have a full publication on the architectural layout of most of the Karanis houses. Moreover, only the largest and best-preserved houses were singled out for publication (see Husselman 1979: 67–73). This practice makes it difficult to discern the range of house types once present at Karanis and obscures the most common houses occupied by typical households at Karanis. Twenty years ago, Peter van Minnen, a papyrologist, urged archaeologists to explore Karanis domestic material by context rather than by material category (van Minnen 1994). Unfortunately, this type of analysis has not yet been accomplished for more than a tiny portion of this site. To date, it is impossible to connect artifacts 102 Anna Lucille Boozer to the houses from which they came, except by consultation of the excavation records held in the Kelsey Museum. It is difficult to determine which types of objects, texts, and architectural features co-occurred with one another and what the distribution of house types, objects, and texts looked like across the site.6 This brief overview of the methodological, phasing, classification, and inter- pretation issues for Karanis indicates that the site is an unstable foundation upon which to erect arguments about domestic life in Roman Egypt. Although the data from Karanis are still useful, caution is essential when employing it as a type site for comparison with new, modern excavations.

The tyranny of typologies for excavators Returning to excavations of the present era, the problem of Karanis persists. When we encountered House B2 at Trimithis (Roman Amheida) (see Map 6.1 and Figures 6.2 and 6.3), it was clear that the preserved remains did not correspond to the published house types from Karanis. Even so, the dominance of the Karanis house type entails a heavy burden of demonstration to allow for a different house form. In order to situate our excavations at B2, it is worthwhile reviewing a bit about the broader site context (for a recent overview of Trimithis, see Boozer 2013). Trimithis is generally very well preserved, with mud brick walls clearly visible for one to two stories above the site surface. Early excavations at House B1 suggested great depths of preservation, up to approximately 4 meters above floor level (see Boozer 2010). House B2 was preserved to only 1.82 meters above floor level, in contrast to House B1. Within the world of Egyptology and Trimithis itself, the preserved depth of House B2 suggested that it was preserved poorly. Despite the deflation of the House B2 walls, we recovered a wealth of artifacts and ecofacts, including soft organics, which are preserved poorly in the location of House B1. This depositional evidence suggested to me that the taphonomy of the B2 structure could be attributed to wind erosion as well as other wear against the mud brick (e.g. the rough edges of the many pot sherds in the area). The preserved anthropogenic layers appeared to be little damaged by wetness and disturbance by rodents or illicit digging. Preservation cannot be based upon depth alone and it appeared that B2 might be better preserved than its depth suggested. Our excavations revealed the clear plan of a house that generally fit a group of houses from nearby Kellis: B2 had barrel vaults over its small rooms and a bread oven; it lacked any evidence of roofing material over its large central room. Given that bread ovens are invariably placed in open areas (Depraetere 2002), and that there was no evidence of any type of roof over room 7, it seemed logical to interpret this room as an aithrion (an internal courtyard). It should be noted that the domestic comparanda from Kellis were excavated to a modern standard and are well published in preliminary reports (e.g. Hope 2005, 2007; Hope and Kaper 1992; Hope et al. 1993). Even so, the known houses from Karanis overshadowed and negated the Kellis comparanda for some individuals on our excavation team. The tyranny of typologies: Romano-Egyptian domestic archaeology 103

Outgoing Outgoing

Outgoing Outgoing

OutgoingOutgoing Outgoing Outgoing

Outgoing Outgoing Outgoing Outgoing

FIGURE 6.2 House B2, Trimithis (Roman period Amheida). (Graphic: N. Warner; copyright: author.)

Debates about a proposed reconstruction for B2 ensued both on site and in subsequent presentations of the material. Some of the excavation participants wanted to revise interpretations of the localized taphonomy, although the new arguments for water seepage did not correspond to wet meter readings or the preservation of other soft organics in B2. Other data, such as Depraetere’s study of bread ovens across Egypt (Depraetere 2002), were discounted because they suggested that B2 might be a house type different from the well-known examples at Karanis. The dominance of Karanis led to the suggestion from some individuals that we discount the depositional, taphonomic, and localized comparanda in published accounts of the site. Instead, it was argued that we should publish B2 as a multi-story, fully roofed structure following the Karanis model. 104 Anna Lucille Boozer

Discussion Having explored the theoretical and practical concerns involved in the use of typologies, as well as the particular example of Karanis and how this use impacts active excavation today, we can tease out several different ways in which ill- conceived and empirically inadequate typologies can undermine future work. The second goal of this discussion is to explore how we might try to retain typologies while recognizing uncertainties and ambiguities in the archaeological record. First, the Karanis/Trimithis case study demonstrates the ways in which the initial phase of systematization can itself be problematic. The initial phase of research on Romano-Egyptian houses at Karanis was partial and impressionistic, which entails that it was simply empirically false. These early impressions of houses had palpable downstream effects because subsequent researchers have employed the Karanis domestic archive without significant consideration to its time and method of creation. The negative impact upon archaeological reasoning can be summed up as follows:

1 Negating the value of archaeological research: the interpretation of archaeological sites ought to be strongly grounded in the archaeological data itself. If we revise our archaeological data based upon typologies alone, what is the purpose of further archaeological excavations? This approach will simply reproduce what has already been excavated and published in the past. 2 Undermining recent typologies: attempts to undermine recent typologies in favor of entrenched concepts impede the development of archaeology as a discipline. In the Karanis/Trimithis example, Depraetere’s recent study of bread ovens from Roman Egypt generated typologies based upon contemporary, well- grounded, and internally coherent archaeological research. It was a struggle to ensure that Depraetere’s research be heeded by archaeologists devoted to old typologies. 3 Cleaning the data: publishing revised archaeological data without showing the relevant analytical scaffolding has dangerous implications for future research. Data cleaning entrenches erroneous typologies deeper within the discipline. Moreover, it is a falsification of data, disguised within what has become acceptable archaeological publication practice.

The Karanis/Trimithis case study has illuminated three broad issues associated with typological tyranny and it is likely that additional extrapolations can be gleaned from further exploration upon this topic. Second, how might we retain typologies within archaeological research while preserving ambiguities within the archaeological record? The feminist practice of archaeology has advocated for an archaeology that interrogates the past and preserves ambiguity within archaeological evidence instead of advancing homo- genized interpretations of the past (Conkey and Gero 1991; Gero 1993; Kus 2006; Meskell 2000; Wylie 2006). This agenda has become mainstream in much archaeological thought, although the feminist roots are still an integral component The tyranny of typologies: Romano-Egyptian domestic archaeology 105 of the drive to maintain ambiguity rather than smooth it over. It is only by openly acknowledging the ambiguity inherent in messy data sets that we can build a richer understanding of the past. What would an archaeology that honors ambiguity look like? An archaeology less restricted by typologies would include a range of interpretations that could follow from the “same” data. Data could also be made accessible to other scholars by providing the database out of which summative monographs were generated. This fluidity would promote more original interpretation rather than privileging the single view of the author or archaeological director. Enabling and even promoting ambiguity, heterogeneity, and complexity in archaeological interpreta- tions has real advantages for understanding the past and it also enables us to work more proactively with typologies. My own decision for the House B2 publication was to compare House B2 to a wider range of Romano-Egyptian houses than just Karanis (see Boozer forth- coming). I suggested that archaeologists should expect that a spectrum of house options once existed and try to report variability among finds more clearly. I also provided full depositional, taphonomic, and typological evidence for my proposed reconstruction (Figure 6.3). The project database will be fully available to the public

FIGURE 6.3 House B2, reconstruction, Trimithis (Roman period Amheida). (Graphic: N. Warner; copyright: author.) 106 Anna Lucille Boozer so that researchers can explore the data on their own terms and come to their own conclusions about the reconstruction (www.amheida.org). It should be noted that this procedure was followed successfully for a volume published on the ostraka from Trimithis (Bagnall and Ruffini 2012). I hope that this open-access policy will encourage additional interpretations of House B2, which would be most welcome to myself and other scholars of Romano-Egyptian domestic contexts. Moreover, it would be desirable to increase open-access data in archaeology more broadly to enable new and altogether different interpretations of sites among scholars and the public alike. Ultimately, it is neither practical nor desirable to dispense with typologies. Despite their indespensibility, typologies must be employed with caution so that archaeologists can organize and communicate their findings. Exploring a spectrum of options and being open to new discoveries is essential for archaeological under- standings to move forward. We also must be open to revising our past inter- pretations and what we think we know about sites. If archaeological data does not fit easily into “known” typologies, it is perhaps best to try something new. As the opening Adams and Adams quote states, “. . . logical typologies are not always useful, and useful typologies are not always logical” (Adams and Adams 1991: 8).

Notes 1 On the development of nineteenth-century Scandinavian typological analysis and archaeology more broadly, see Gräslund 2008 [1987]. 2 For a comprehensive review and analysis of the Typological Debate, see Dunnell 1986. 3 It is also noteworthy that Bailey cites Husselman (1979), a monograph, and provides no guiding page numbers (Bailey 2012). 4 For an overview of this excavation as well as other archaeological research at Soknopaiou Nesos, see Davoli (1998: 39–71). 5 Davoli summarizes the work at Karanis, as well as some of the issues with the University of Michigan methodologies and publication practices (Davoli 1998). 6 Still, many scholars consider the data from Karanis to be very clear (e.g. Bowman 1986: 148).

References Adams, W. Y. and Adams, E. (1991) Archaeological Typology and Practical Reality: A Dialectical Approach to Artifact Classification and Sorting, Cambridge: Cambridge University Press. Aldenderfer, M. (1996) “Typological analysis,” in B. M. Fagan (ed.) The Oxford Companion to Archaeology, Oxford: Oxford University Press, 727–728. Allison, P. M. (1999) “Labels for ladles: interpreting the material culture of Roman households,” in P. M. Allison (ed.) The Archaeology of Household Activities, London: Routledge, 57–77. Bagnall, R. S. (1993) Egypt in Late Antiquity, Princeton: Princeton University Press. Bagnall, R. S. and Ruffini, G. R. (2012) Ostraka from Trimithis, Volume 1: Texts from the 2004-2007 Seasons, New York: Institute for the Study of the Ancient World and New York University Press. Bailey, D. M. (2012) “Classical architecture,” in C. Riggs (ed.) The Oxford Handbook of Roman Egypt, Oxford: Oxford University Press, 189–204. The tyranny of typologies: Romano-Egyptian domestic archaeology 107

Binford, L. (1972) An Archaeological Perspective, New York: Seminar Press. Boak, A. E. R. (1933) Karanis: The Temples. Coin Hoards, Botanical and Zoological Reports, Season 1924–31, Humanistic Series 30. Ann Arbor: University of Michigan. Boak, A. E. R. and Peterson, E. E. (1931) Karanis: Topographical and Architectural Report of Excavations During the Seasons 1924–28, Humanistic Series 25, Ann Arbor: University of Michigan. Boak, A. E. R, Peterson, E. E. and Haatveldt, R. A. (1935) “Soknopaiou Nesos: the University of Michigan excavations at Dimê in 1931–32,” Ann Arbor: University of Michigan Press. Boozer, A. L. (2010) “Memory and microhistory of an empire: domestic contexts in Roman Amheida, Egypt.” in D. Boric´ (ed.) Archaeology and Memory, Oxford: Oxbow Books, 138–157. Boozer, A. L. (2013) “The archaeology of Amheida (Egypt),” in C. Smith (ed.) The Encyclopedia of Global Archaeology, New York: Springer. Boozer, A. L. (forthcoming) A Romano-Egyptian House in Transition: Amheida House B2, New York: ISAW & NYU Press. Bowes, K. (2010) Houses and Society in the Later Roman Empire, London: Duckworth. Bowman, A. K. (1986) Egypt after the Pharaohs 332 BC – AD 642: From Alexander to the Arab Conquest, Berkeley, CA: University of California Press. Childe, V. G. (1929) The Danube in Prehistory, Oxford: Oxford University Press. Conkey, M. W. and Gero, J. M. (1991) “Tensions, pluralities, and engendering archaeology: an introduction to women in prehistory,” in J. M. Gero and M. W. Conkey (eds) Engendering Archaeology: Women and Prehistory, Cambridge, MA: Blackwell, 3–30. Costin, C. L. (2007) “Thinking about production: phenomenological classification and lexical semantics,” in Z. Hruby and R. Flad (eds) Rethinking Craft Specialization in Complex Societies. Archaeological Papers of the American Anthropological Association, 17: 143–162. Cowgill, G. L. (1990) “Artifact classification and archaeological purposes,” in A. Voorips (ed.) Mathematics and Information Science in Archaeology: A Flexible Framework, Bonn: Holos, 61–78. Davoli, P. (1998) L’archeologia Urbana nel Fayyum di età ellenistica e romana, Volume Monografie 1, Bolognia: Generoso Procaccini. Depraetere, D. D. E. (2002) “A comparative study on the construction and the use of the domestic bread oven in Egypt during the Graeco-Roman and Late Antique/Early Byzantine period,” Mitteilungen des Deutschen Archäologischen Institutes Abteilung Kairo, 58: 119–156, Plates 15–16. Dunnell, R. C. (1986) “Methodological issues in Americanist artefact classification,” Advances in Archaeological Method and Theory, 9: 149–207. Eckardt, H. (2010) “Introduction,” Roman Diasporas: Archaeological Approaches to Mobility and Diversity in the Roman Empire, in H. Eckardt (ed.) Portsmouth, RI: Journal of Roman Archaeology: 7–11. Fish, P. R. (1978) “Consistency in archaeological measurement and classification: a pilot study,” American Antiquity, 43(1): 86–89. Fritz, J. M. and Plog, F. T. (1970) “The nature of archaeological explanation,” American Antiquity, 35(4): 405–412. Gazda, E. K. and Hessenbruch, C. (1978) “Sculptures in stone,” in E. K. Gazda (ed.) Guardians of the Nile: sculptures from Karanis in the Fayoum (c.250 BC – AD 450), Kelsey Museum of Archaeology, University of Michigan, Ann Arbor, 14 October – 17 December, Ann Arbor: Kelsey Museum, 18–21. Gero, J. M. (1993) “The social world of prehistoric facts: gender and power in Paleoindian research,” in H. D Cros and L. Smith (eds) Women in Archaeology: A Feminist Critique, 108 Anna Lucille Boozer

Department of Prehistory, Research School of Pacific Studies, Australian National University, 31–40. Gero, J. M. (2007) “Honoring ambiguity/problematizing certitude,” Journal of Archaeological Method and Theory, 14: 311–327. Gräslund, B. (2008 [1987]) The Birth of Prehistoric Chronology: Dating Methods and Dating Systems in Nineteenth-Century Scandinavian Archaeology, New Studies in Archaeology, Cambridge: Cambridge University Press. Haatveldt, R. A. and Peterson, E. E. (1964) Coins from Karanis: The University of Michigan Excavations 1924–1935, Ann Arbor, MI: Kelsey Museum of Archaeology. Harden, D. B. (1936) Roman Glass from Karanis Found by the University of Michigan Archaeological Expedition in Egypt, 1924–29, Ann Arbor, MI: University of Michigan Studies. Higashi, E. L. (1990) “Conical glass vessels from Karanis: function and meaning in a Pagan/Christian context in rural Egypt,” PhD, History of Art, University of Michigan, Ann Arbor, MI. Hodder, I. (1990) The Domestication of Europe, Oxford: Blackwell. Hope, C. A. (1993) “Ismant el-Kharab 1991–92,” The Journal of the Society for the Study of Egyptian Antiquities, 19: 1–22. Hope, C. A. (2005) “Report on the excavations at Ismant el-Kharab and Mut el-Kharab in 2005,” Bulletin of the Australian Centre for Egyptology, 16: 35–37. Hope, C. A. (2007) “Brief report on the 2007 excavations at Ismant el-Kharab,” Bulletin of the Australian Centre for Egyptology, 18. Hope, C. A. and Kaper, O. E. (1992) “Excavations at Ismant el-Kharab – 1992,” Bulletin of the Australian Centre for Egyptology, 3: 41–49. Hope, C. A., Kaper, O. E., Bowen, G. E. and Patten, S. F. (1993) “Ismant el-Kharab 1991–92,” Journal of the Society for the Study of Egyptian Antiquities, 19: 1–22. Husselman, E. M. (1952) “The granaries of Karanis,” Transactions and Proceedings of the American Philological Association, 83: 56–73. Husselman, E. M. (1953) “The dovecotes of Karanis,” Transactions and Proceedings of the American Philological Association, 84: 81–91. Husselman, E. M. (ed.) (1971) Papyri from Karanis, third series, Cleveland, OH: Case Western Reserve University Press. Husselman, E. M. (1979) Karanis Excavations of the University of Michigan in Egypt, 1928–1935: Topography and Architecture – A Summary of the Reports of the Director, Enoch E. Peterson, Ann Arbor, MI: University of Michigan. Johnson, B. (1981) Pottery from Karanis: Excavations of the University of Michigan, Ann Arbor, MI: The University of Michigan Press. Kus, S. (2006) “In the mist of moving waters: material, metaphor, and feminist archaeology,” in P. Geller and M. Stockett (eds) Feminist Anthropology, Philadelphia, PA: University of Pennsylvania, 105–114. Malouta, M. (2012) “Families, households, and children,” in C. Riggs (ed.) The Oxford Handbook of Roman Egypt, Oxford: Oxford University Press, 288–304. Meskell, L. M. (2000) “Feminism in archaeology,” in L. Code (ed.) Encyclopedia of Feminist Theories, London: Routledge, 26–27. Rowlandson, J. (ed.) (1998) Women and Society in Greek and Roman Egypt: A Sourcebook, Cambridge: Cambridge University Press. Ruffini, G. R. (2012) Medieval Nubia: A Social and Economic History, Oxford: Oxford University Press. Shier, L. A. (1978) Terracotta Lamps from Karanis, Egypt: Excavations of the University of Michigan, Ann Arbor, MI: University of Michigan Press. The tyranny of typologies: Romano-Egyptian domestic archaeology 109

Spaulding, A. C. (1953) “Statistical techniques for the discovery of artifact types,” American Antiquity, 18: 305–314. Taylor, W. W. (1983 [1948]) A Study of Archaeology, Carbondale, IL: Southern Illinois University. Thomas, T. K. (2001) Artifacts of Everyday Life: Textiles from Karanis, Egypt in the Kelsey Museum of Archaeology, Ann Arbor, MI: Kelsey Museum of Archaeology. van Minnen, P. (1994) “House-to-house enquiries: an interdisciplinary approach to Roman Karanis,” Zeitschrift für Papyrologie und Epigraphik, 100: 227–251. Vierra, R. K. (1982) “Typology, classification, and theory building,” in R. Whallon and J. A. Brown (eds) Essays on Archaeological Typology, Evanston, IL: Center for American Archaeology, 162–175. Whallon, R. and Brown, J. A. (eds) (1982) Essays on Archaeological Typology, Evanston, IL: Center for American Archeology Press. Wilfong, T.G. (2012) “The University of Michigan excavation of Karanis (1924–1935): images from the Kelsey Museum Photographic Archive,” in C. Rigss (ed.) The Oxford Handbook of Roman Egypt, Oxford: Oxford University Press, 223–243. Wylie, A. (2002a) “Bootstrapping in the un-natural sciences – archaeology for example,” in A. Wylie (ed.) Thinking from Things: Essays in the Philosophy of Archaeology, Berkeley, CA: University of California Press, 179–184. Wylie, A. (2002b) “The typology debate,” in A. Wylie (ed.) Thinking from Things: Essays in the Philosophy of Archaeology, Berkeley, CA: University of California Press, 42–56. Wylie, A. (2006) “Afterword: on waves,” in P. Geller and M. Stockett (eds) Feminist Anthropology, Philadelphia, PA: University of Pennsylvania, 167–175. Wylie, A. (2011) “Critical distance: stabilizing evidential claims in archaeology,” in P. Dawid, W. Twining and M. Vasilaki (eds) Evidence, Inference and Enquiry, Proceedings of the British Academy, 171: 371–394. Yeivin, S. (1934) “Notes on the Northern Temple at Karanis Aegyptus,” Revista italiana di Egittolgia e di Papirologia, 14: 41–79. Youtie, H. C. and Pearl, O. M. (1939) Tax Rolls from Karanis, Ann Arbor, MI: University of Michigan Press. Youtie, H. C. and Pearl, O. M. (1944) Papyri and Ostraka from Karanis, Ann Arbor, MI: University of Michigan Press. Ziman, J. (1978) Reliable Knowledge, Cambridge: Cambridge University Press. This page intentionally left blank PART II Cross-field trade: archaeological applications of external expertise and technologies This page intentionally left blank 7 THE ARCHAEOLOGICAL BAZAAR

Scientific methods for sale? Or: ‘putting the “arch-” back into archaeometry’

Mark Pollard and Peter Bray

Whether or not one regards archaeology as a science or as a humanities-based discipline that calls upon scientific methods to aid in the interpretation of evidence, there can be little doubt that archaeology and the natural sciences have been intertwined since their very inceptions in the seventeenth century. Indeed, up until the nineteenth century, prehistory was seen as a branch of natural science (along with palaeontology, geology and palaeobiology). Archaeology of the (European) classical world took a different course during the eighteenth century, dominated by the desire to corroborate the narratives presented in the classical texts and the perception of archaeological objects as ‘art’ rather than as loci of archaeological information – a division which to some extent continues today, but less clearly demarcated. Early science in archaeology was primarily the application of rigorous measurement – evidenced by the earliest mapping and recording of prehistoric monuments such as Avebury and Stonehenge by antiquarians such as John Aubrey (1626–1697) during the 1660s, which coincided with, and contributed to, the foundation of the Royal Society. During subsequent work at Stonehenge in 1720, Edmond Halley (1646–1742), whilst considering the antiquity of Stonehenge (from astronomical and magnetic observations), collected some samples of stone. These were later examined at the Royal Society in 1740 by William Stukeley (1687–1765), who undertook the earliest petrological examination of these stones (Johnson 2012: 64). The evolution of archaeological excavation from ‘treasure hunt’ to scientific recovery and recording followed during the late nineteenth century, and was attributed to Augustus Henry Lane-Fox Pitt-Rivers (1827–1900) in England and Sir William Matthew Flinders Petrie FRS (1853–1942), working in Egypt. Given this intimate relationship for over 300 years, it is worth examining why there is still a perceived division in some quarters between science and archaeology in the twenty-first century. Or, put more positively, how can we further integrate 114 Mark Pollard and Peter Bray science and archaeology? At one end of the spectrum, this happens already, with scientific applications being integrated into archaeological research designs from the outset, and with no artificial demarcation between ‘science’ and ‘archaeology’. This is the ideal, and the goal for which we should be aiming. At the other end, however, there is still archaeology carried out in which it is glaringly obvious that a certain amount of scientific input could contribute positively to the debate. It is perhaps invidious to be specific, but issues relating to discrepancies between the conventional chronologies based on pottery typology for parts of the Mediterranean and the absolute dates of events such as the eruption of Thera spring to mind. A programme of direct dating of such ceramic chronologies would, we think, be both important and informative. Still at the same end of the spectrum, but seen from the other side, there continues to be a proportion of scientific work carried out on archaeological material, the consequence of which is, to say the least, trivial. This has been elsewhere exemplified in an exaggerated fashion as using a synchrotron to demonstrate that pottery is made of clay, or more generally paraphrased as showing that ‘things are made of stuff ’ (Pollard 2013a). Such work clearly fails the important ‘so what?’ test. Less extreme, but perhaps more difficult to deal with in the long run, is the tendency of some major excavations to produce numerous ‘specialist reports’, describing isolated applications of an apparently random selection of scientific techniques, the purpose of which seems to be more to tick boxes than actually to answer real questions. This gives rise to the perception that the scientific analysis of archaeological material can be regarded as a self-service buffet: select what you fancy, constrained by overall budget, but not necessarily focused on any specific archaeological question. So how should we do it better? Many articles have been written pointing out these problems, mostly concluding that we need to work together more effectively, but without specifically saying how. This chapter will attempt to break with tradition and show how it can be done in the contested field of metal provenance. As part of the ‘Age of Enlightenment’, the newly developed science of chemical analysis was applied to archaeological objects during the last decade of the eighteenth century, primarily in revolutionary Paris (Pollard 2013b). This consisted partly of the enthusiastic examination of the contents of the many ‘cabinets of curiosities’ using the new tool of analytical chemistry. Initially these early analyses were primarily directed at finding out from what the archaeological objects were made, although it would be a mistake to see them purely as examples of ‘curiosity- driven’ research (sadly a term that in some circles, at least in the UK, is seen as derogatory, but that history shows has yielded many unexpected returns). There is clear evidence of an eighteenth-century ‘question-driven research framework’, certainly for the early chemical analysis of copper alloy objects. For example, the Abbé Mongez (1747–1835) instigated the analysis of a set of Roman, Greek and Gaulish coins in order to understand what the alloying metal in bronze was – partly, it would appear, to refute the ideas previously put forward by the Comte de Caylus (1692–1765), who thought that the alloying metal in the bronze swords of the ‘ancients’ was iron. Apart from what might be seen in this case as simple academic The archaeological bazaar: scientific methods for sale? 115 progress by robust contest, there is also evidence of practical forces at work behind such endeavour. The main impetus for devising a method for determining the chemical nature of bronze during the French Revolution was to enable reduction of the tin that was known to be in bell metal (which was in plentiful supply, given the de-Christianization of France at the time) in order to produce cannon (then in very short supply). The fact that such knowledge also fostered the beginnings of what we now might call ‘chemical archaeometallurgy’ was a fortuitous by-product, perhaps akin to the development of neutron activation analysis from the activities of the Manhattan Project during the Second World War.

The quest for provenance These relatively simple early questions soon gave way to more complex archaeological questions, however, which remain largely unchanged today. By 1842, the idea had clearly emerged that the chemical analysis of archaeological copper alloy objects could provide not only evidence of the geological source of copper, but, much more significantly, evidence of contact between geographically remote peoples (see Pollard et al. in press). This became enshrined in the concept of provenance, an idea clearly summarised by the French geologist Damour in 1856: ‘when an object on which the hand of man marks his work, and the matter from which it is made is foreign, we infer that it, or the matter from which it was formed, was the subject of transportation’. Not only did this mark the beginning of the quest for provenance in a wide range of materials, but it indirectly encapsulated the emergence of the concept of artefact as information repository rather than (or in addition to) as work of art. In the sphere of the chemical analysis of inorganic objects for use as a proxy for ‘contact’ (however that term might be construed) between peoples in the past, little has changed theoretically in the last 150 years. Of course, analytical techniques have changed (from gravimetry in the late eighteenth century to LA-ICP-MS (laser ablation inductively coupled plasma mass spectrometry) and the like in the twenty- first century), and the range of elements reported has grown to reflect the increasing sensitivity of the analytical methodology, but even with the advent of new concepts of analysis, such as the precise measurement of isotopic ratios, the theoretical framework has not changed. It is still based on the concept of the chemical or isotopic ‘fingerprint’ – some characteristic of the source of the raw material that passes through to the finished object and allows a relationship to be established between the two. For some materials, such as obsidian and lithic materials, this concept may seem completely reasonable, since no modifications made to the form of the tools are likely to affect this chemical or isotopic fingerprint. We have pointed out elsewhere (Pollard et al. in press), however, that even for such materials the time taken for an object to move from A to B is an important and oft-neglected parameter in such studies. For other materials, however, it is less clear that these principles apply, at least in such simple terms. In particular, for objects made from easily recyclable material (such as copper alloys, precious metals and glass, which 116 Mark Pollard and Peter Bray can be melted, mixed and recast), it is obvious that the process of recycling might result in the blurring of any chemical or isotopic fingerprints, particularly if materials from different geological sources are mixed. The term recycling is one that urgently requires further definition in an archaeological context, because it has been widely argued that recycling of metal was not practised in Europe until well into the Middle Bronze Age. This may be true if we interpret recycling purely in the modern sense – the gathering of ‘scrap’ material, grouped together and re- melted to make a new object or objects, which may bear no relationship typologi- cally or symbolically to the ‘scrap’ material. However, we may conceive of other modes of recycling, where a single object is remodelled to produce the same object but of a different shape (perhaps the remodelling of an axe to meet local concepts of ‘axe-ness’), for example, or where several objects of the same type are re-melted to produce one or more of the same objects (implying a deliberate conservation of form). Crucially, such patterns of recycling would be largely invisible archaeologi- cally (except, perhaps, for the random appearance of casting moulds, divorced from any evidence of primary metal production), in contrast to ‘conventional recycling’, which might be evidenced by ‘founder’s hoards’. Despite being invisible in the archaeological record, it is clear that such practices might have a significant effect on a conventional provenance study, to the point where such work could become seriously misleading. Large analytical campaigns have been undertaken on European Bronze Age metalwork (e.g. Otto and Witter 1952; Junghans et al. 1960, 1968–1974 (also known as the SAM project); Ottaway 1982; Krause 2003), and later programmes of analysis of lead isotopes, but the impacts of such studies on mainstream archaeological thinking are difficult to quantify. Undoubtedly a number of widely accepted definitions of specific copper sources have emerged, such as a recognition of the Alpine origins of so-called Singen and Ösenring metal (Harding 2000: 204), or the distinctive copper (high in antimony and silver) produced by the Cabrières mines in Southern France (Ambert 1995). But if we apply the key question – has any of this work changed the overall trajectory of archaeological thinking? – we may conclude that the outcomes have so far been somewhat disappointing. Controversially, we have argued that this is because the fundamental assumptions of provenance – particularly, the lack of mixing and recycling – have been violated from the time of the earliest use of metal. Thus, traditional approaches (both chemical and isotopic) may work well in defining metal circulation close to the source, where one may expect mixing and recycling to be minimal, but as distance from source (and time elapsed) increases these simple characteristics disappear, and interpretations become either more uncertain, or, worse, illusory.

An alternative approach It is one thing to identify flaws in current methodology, but completely another to devise a model that offers some improvement. In terms of interpreting trace element data in copper alloys, we have proposed a methodology that, we think, The archaeological bazaar: scientific methods for sale? 117 offers an improvement over existing interpretations – principally because it assumes that human intervention (such as mixing, recycling and reshaping) is the norm rather than something to be ignored. The consequence of this is that the chemistry of a particular unit of metal is not fixed – it can change as it progresses through its ‘life’, either because of mixing, or as a result of going through (potentially) repeated cycles of melting and casting or reworking, during which the more volatile trace elements (especially arsenic, but also antimony) are lost due to oxidation. These observations of oxidative loss have been known in the archaeometallurgical literature since at least the 1970s (e.g. McKerrell and Tylecote 1972), but the contemporary discussions were phrased in terms of such behaviour ‘washing out’ or obscuring provenance signals. Rather than seeing this chemical mutability as an obstacle to interpretation, we view it as an additional source of information about the biography of the unit of metal, and the objects into which it is turned. We focus on the elements arsenic (As), antimony (Sb), silver (Ag) and nickel (Ni) in the copper because they have long been seen as diagnostic and have featured in previous classification schemes by several authors, and are consequently commonly reported. As has also been widely observed, these elements are regarded as diagnostic of ore source, since they tend to be either present or absent in the ores that are known to have been used in antiquity. Nor do they suffer from pro- nounced partition problems in copper, unlike bismuth (Bi), which is consequently unevenly distributed and is therefore an unreliable marker. As we have shown (Bray and Pollard 2012), these elements have a range of different chemical and thermodynamic properties and therefore appear to be useful in teasing out socio- technological histories. In order to do this, we need to make a number of assumptions. The first is that the actual concentration of any of these particular trace elements is only of secondary importance – the key initial categorization is done purely on the presence/absence of these elements, and only subsequently do we use averages of absolute values to help interpret the data further. Second, we must not over- interpret the chemical analysis of a single object; we need to synthesise a range of other datasets to draw secure conclusions (e.g. is it primary metal? is it obviously recycled?). Our interpretations are based on type averages from a large number of analyses – the types being defined on the basis of geography, chronology and typology. To make an analogy with statistical mechanics, we can make no prediction about the state of any particular component of the system, but we can specify the condition of the assemblage. This approach has the advantage that it makes use of the vast quantities of published analyses of archaeological metalwork without re-sampling or fresh analyses, but it does have to deal with differences in accuracy, precision, levels of detection and reporting conventions over two centuries of analysis. We note also that some data remain surprisingly difficult to track down and access. In brief, the methodology is simple. It is based primarily on large-scale mapping – our database is linked to a GIS, which allows us to plot chemical patterns in objects over time and space. We consider two discrete and independent levels of 118 Mark Pollard and Peter Bray chemical information: that of the alloys that are used; and also the ‘type’ of copper used. This ‘type’ is defined by the presence or absence of arsenic, antimony, silver and nickel, which are either passed from the ore into the smelted metal, or introduced as the result of mixing different types of copper. As explained below, however, it is important to appreciate immediately that ‘copper type’ and ‘copper source’ are not synonymous. The major elements alloyed with copper are well known: tin; lead; and zinc (arguably arsenic and also antimony should also be included, but in general the chemical evidence for the occasional deliberate, separate, addition of these elements is very distinctive). We wish to allow for the possibility of copper metal and alloying elements having separate patterns of existence, with alloying episodes taking place at secondary centres rather than being a single event that only occurs at the primary source of the metal. We also see the possibility of ‘re-alloying’ – the addition of further quantities of alloying metal to boost the required properties, most likely colour. We do not therefore necessarily regard alloy composition to be a fixed property of the metal unit. Levels of the primary alloying metal can be deliberately boosted, or diluted by the addition of fresh metal, and other alloying elements can be added. A key decision is of course at what level does a particular alloying element become a ‘deliberate’, as opposed to an ‘accidental’ component, perhaps coming in from impurities in the primary ore. In a heavily recycled world this becomes a false dichotomy, because the assumption has to be that significant levels of any alloying element could be introduced by particular choices in the recycling pattern employed. We note, for example, that in any large assemblage of copper alloy objects, such as Roman brooches or medieval brass, most objects in fact contain significant (several per cent) levels of tin, lead and zinc, even though they may be described primarily as being of either bronze or brass. This observation itself suggests that recycling was widely practised in the past, unless one is inclined to believe that small quantities of secondary alloying elements were deliberately added to enhance a particular property, even though it is highly unlikely that such additions produced any detectable results. By plotting the location, typology, date and ubiquity of alloys containing tin in the Early Bronze Age, for example, we can use this method to track the spread of tin bronze in any particular region. In order to look more closely at the sources of copper, we have devised a semi-quantitative method, which allows for the fact (as we see it) that the actual level of the trace elements in the copper might be variable, but that the presence or absence of such elements is highly significant. First, we mathematically strip out the alloying elements (tin, lead and zinc), and re-normalise the remaining elements to 100 per cent, to give what we describe as the com- position of the copper base. We realise that this is somewhat arbitrary (since, as discussed above, the concept of an alloying element is ambiguous), and that complications might arise, for example, with the addition of argentiferous lead bringing in silver that is unrelated to the source of the copper (or indeed, vice versa, with lead being added to de-silver the copper, resulting in a lower silver value), but it is a necessary step to avoid ‘dilution effects’ when significant levels of alloying The archaeological bazaar: scientific methods for sale? 119 elements have been added. Having calculated a ‘raw copper composition’ for each object, we can now assign each object to a ‘metal type’ on the basis of presence/ absence of the four trace elements, As, Sb, Ag and Ni. We simply allocate the metal composition to one of 16 bins, on the basis of presence/absence (Y/N) of each trace element, ordered for convenience as listed in Table 7.1. Thus, a metal with As but nothing else would be YNNN (assigned type 2 in our numbering system), whereas ‘Ag only’ would be NNYN (type 4). Again, an arbitrary decision is needed about when ‘presence’ becomes ‘absence’. This can be determined by a close examination of the data to estimate minimum detectable levels (mdl) for each element, but in mixed data (i.e. data from different methods of analysis and different dates of determination) it is often only possible to use an arbitrary cut-off of 0.1 per cent. We then proceed as described above, using mapping options to determine the extent, movement and timing of the circulation of particular metal types. An example is shown in Figure 7.1, where we can see type 2 metal (YNNN) originating in southern Iberia and moving up towards the Atlantic coast of France and then into the UK. We must emphasise that these metal types do not necessarily correspond to ‘sources’ or ‘mine sites’. Whilst there is undoubtedly information about source within these descriptions, we do not make any assumptions about allocating a particular copper type to a specific copper source, known or unknown. A single metal type may contain copper from one or more discrete geological sources – all we can say is that these sources must have the same characteristic pattern TABLE 7.1 The definition of copper types based on the presence (Y) or absence (N) of, of trace elements. For example, in order, arsenic (As), antimony (Sb), silver the As-Sb-Ag metal of type 12 is (Ag) and nickel (Ni). An element is taken to associated with mines in Ross Island be present at above 0.1 per cent by weight. of County Kerry, Ireland and also in the Alps. We can, however, offer Copper type Definition some observations on the location 1 NNNN of these potential sources by plot- 2 YNNN ting the distribution of the ubiquity 3 NYNN of particular metal types (taking into 4 NNYN account, where possible, typology 5 NNNY and chronology) and identifying 6 YYNN geographical concentrations where 7 NYYN a metal type becomes dominant in 8 NNYY 9 YNYN the assemblage. It is then reasonable 10 NYNY to assume that the source is located 11 YNNY somewhere within this region. For 12 YYYN example, we can predict that the 13 NYYY source of the type 2 metal in the 14 YYNY EBA shown in Figure 7.1 must be 15 YNYY somewhere in Iberia. At this point, 16 YYYY it becomes useful to seek additional FIGURE 7.1 Distribution of type 2 copper (YNNN, ‘Arsenic only metal’) during the Early Bronze Age. The high proportion of objects made from this metal in Iberia suggests that the source is located there, and the relatively high values up the coast of Atlantic Europe suggest that this metal is traded along the coast. The higher values in Anatolia and Bulgaria seem to show another independent source of this metal coming in from further east. (Underlying base map after Krause 2003.)

Not enough data

<5%

5 to 24.9%

25 to 49.9%

50+ % The archaeological bazaar: scientific methods for sale? 121 supporting information (archaeological or geological) to confirm and identify the possible location of a metal source within the highlighted region. It is important to note that this method is therefore independent of any prior knowledge of copper sources, and is thus a powerful means of testing previous assumptions. Not only does this approach give a simple way of looking independently at the spread of alloying practices and the circulation of copper, it also offers a means of identifying and interpreting recycling and mixing patterns. We have argued above that the actual levels of specific trace elements are variable in an assemblage of objects, in part due to the loss of the more volatile elements (As and Sb) during high temperature re-processing, and also because the mixing of copper of different types can dilute or add trace elements. In terms of oxidative losses – in general we would expect As to be lost quickly and Sb to be lost less quickly, but that neither Ag nor Ni will be lost – we can produce a ‘decay table’ for each metal type, which shows how one metal type will transform into another as the more vulnerable elements are lost (an example is shown in Table 7.2, but other pathways are possible). For example, type 1 (NNNN, in which all four trace elements are below the cut-off value) will not change type under oxidative loss, whereas type 12 (YYYN) will become type 9 (YNYN) if antimony drops below our cut-off point, or type 7 (NYYN) if arsenic is lost. The end point for a metallurgical system that continually re-melts what was originally classed as type 12 would be type 4 (NNYN), as only the more noble silver would remain. Early Bronze Age Ireland provides a case study for the interplay of this ‘family’ of related metal types. The Ross Island mine produced an As-Sb-Ag-rich copper (type 12: YYYN) that dominated the Irish and British Early Bronze Age metal assemblage. However, after around 1900 BC the mine flooded, fresh production from the site became impossible and reclaimed and re-melted metal therefore came to be increasingly more important to local metallurgy. Table 7.3 summarises the complete chemical assemblage for Irish Early Bronze Age metal and clearly shows the importance of ‘prime’ type-12 Ross Island metal and its possible ‘daughter’ compositions (types 7, 9 and 4). At the beginning of the sequence (mid-third millennium BC) the assemblage is dominated by type-12 metal (86.1 per cent of all known analyses). The proportion of this type-12 copper drops rapidly after c.2000 BC, to be replaced by a growth in types 9 and 4 (which we argue are recycled ‘daughter’ composi- tions), and also a replacement by type-1 metal from a new source, possibly the Mount Gabriel mines, which have been excavated and absolutely dated to the early second millennium BC. Without data from the archaeological excavation of the mine, combined with ore petrology, a geographical and typo-chronological understanding of the artefacts and absolute dating, we would not be able to understand the range of chemical signals recorded by the laboratory scientists. Even worse, we may confuse the cluster of data provided by recycled ‘arsenic and antimony-depleted metal’ (types 4 and 9) as a heretofore missing source. To draw an analogy from another area of archaeological science, the use of stratigraphical, typological and geographical data has revolutionised the precision and accuracy of dating models within radiocarbon. TABLE 7.2 Modelled depletion families for the 16 possible combinations of presence/absence of arsenic (As), antimony (Sb), silver (Ag) and nickel (Ni).

Prime type T 1 T 2 T 3 T 4 T 5 T 6 T 7 T 8 T 9 T 10 T 11 T 12 T 13 T 14 T 15 T 16

NNNN YNNN NYNN NNYN NNNY YYNN NYYN NNYY YNYN NYNY YNNY YYYN NYYY YYNY YNYY YYYY

Possible 1 1 2 1 3 1 4 4 5 5 3 2 4 4 8 8 4 4 5 5 5 5 9 7 8 8 11 10 8 8 15 13 secondary types

TABLE 7.3 The ubiquity of copper types in Irish Early Bronze Age copper alloy assemblages.

Type Type Type Type Type Type Type Type Type Type Type Type Type Type Type Type 1234 56 7 8 9 10 11 12 13 14 15 16

Mid-third mill. BC 1.9 1.5 0.8 0.4 0.0 3.5 3.1 0.0 1.9 0.0 0.4 86.1 0.0 0.0 0.0 0.4

2200–2000 BC 8.8 0.6 0.0 0.6 1.2 2.9 6.5 0.0 0.0 0.0 0.0 77.1 0.6 0.6 0.0 1.2

2000–1700 BC 35.5 17.5 1.1 7.0 3.0 1.1 3.8 0.0 16.7 0.0 1.9 11.8 0.0 0.0 0.0 0.8

1700–1400 BC 34.8 10.6 1.4 10.6 7.1 2.8 1.4 0.0 14.9 0.0 6.4 6.4 0.0 2.1 0.7 0.7 The archaeological bazaar: scientific methods for sale? 123

We feel that a similar process is feasible for the work of archaeological chemistry. It is possible – indeed, essential – to move beyond static concepts of provenance to explore the whole socio-technological arena.

Putting the ‘arch-’ back into archaeometry? A great deal of work over the last century has gone into identifying and interpreting patterns in the chemical composition of copper artefacts, and some of what has been found remains uncontroversial. Our point is that where these ‘traditional’ approaches (we would say ‘static’ interpretations) have not worked well is where mixing and cycling have been important, which is more likely to be at a distance from the sources, although not exclusively so. In these cases we would argue that our approach is the most appropriate, but, because it also works well in unmixed/ unrecycled systems, it is a better overall approach. There are, however, a number of corollaries if this idea is to be adopted. Foremost is the observation that cluster analysis and related numerical techniques are not appropriate, unless some sort of ‘fuzzy’ system is adopted, in which the absolute values of trace or alloying elements are given very limited or unequal categorical weight. If we accept that chemical composition is susceptible to change and drift over the time of a unit of metal’s existence as it potentially changes shape, alloy composition and so on, then the aim of drawing circles around these shifting sands seems arbitrary and counter- productive, and often appears to exclude human-level information about tech- nology and social decisions. Exceptions, where cluster analysis works, are of course to be expected, but we would see them as just that – exceptions, more likely to be appropriate in regions close to an active and productive source, where fresh metal is in ready supply, and mixing/recycling likely to be at a minimum. The large majority of archaeological metalwork does not fit this criterion. Many pages have been devoted to the advantages and shortcomings of the use of lead isotopes for provenance studies of metals in general and copper alloys in particular, and only a short summary of these arguments can be presented here (see Pollard 2009 for further discussion). In brief, criticisms have centred largely on the definition of the extent and shape of the isotopic ‘field’ for a specific ore source, since isotopic values can vary considerably within a single deposit, and there is no reason to expect a normal distribution of the values. Thus the practice of drawing a ‘95 per cent confidence ellipse’ around the few measured isotopic values in a particular region is not likely to capture fully the true extent of that ore source, and any assignments of object to source made on this basis are likely to be suspect. Kernel density estimation (‘KDE’, Baxter et al. 1997) offers the prospect of building up an estimate of the true shape and size of an isotope field whilst making few extra assumptions about the nature of the data, if sufficient data are available, but this is rarely applied. Indeed, when sufficiently large numbers of ore samples are measured, it often appears that many sources overlap and are therefore indistinguishable. In their study of the British sources of copper in the Bronze Age, Rohl and Needham (1998) showed that, despite the fact that copper mineralisation in the British Isles 124 Mark Pollard and Peter Bray occurs in four different geological environments, the isotope data are essentially overlapping. Consequently, these authors present the data from all British and Welsh ore sources as ‘EWLIO’ – the ‘English and Welsh Lead Isotope Outline’. They also show how this EWLIO relates to Scottish, Irish, French and German sources – in all cases, these data overlie EWLIO, although they often extend beyond it. Therefore, the prospects for fine-scale differentiation between potential sources across a large area are not very good, apart from occasional exceptions, such as the high local uranium values in Cornwall, leading to extreme lead isotopic values. To the dispassionate observer, all of this would suggest that the goal of unequivocally linking object to source via lead isotope data is doomed to failure, apart from in a few special cases. Furthermore, if our mixing/recycling models have any reality, then it is now easy to understand why inferences made from lead isotopes have been (on the whole) disappointing at best, and completely inexplica- ble at worst. Indeed, in the few cases where the interpretation of lead isotope ratios have been augmented by the object’s chemical composition, the impact of metal mixing can sometimes be clearly seen. Rohl and Needham (1998) came close to demonstrating this conclusion, but needed to go one step further. They argued that much of the metalwork from the Taunton Phase of the British Middle Bronze Age is chemically and isotopically similar, assigning them the labels impurity-lead isotope types (IMPLI) 12 and 13. Within this type of artefacts, which correspond to our types 11 and 14 (YNNY, YYNY), we note that those with lower nickel levels plot towards the edge of the overall lead isotope distribution for Taunton Phase objects (Figure 7.2). We interpret this as the result of mixing copper from a high-nickel source with that from a low-nickel source, which pulls the isotope ratios away from their ‘prime value’ (that of the high-nickel source) towards the values for the low-nickel source. Ironically, through their use of lead isotopes and chemistry to identify complex technological behaviour, Rohl and Needham provide excellent evidence against attempting to define discreet, bounded IMPLI types. Instead we must consider mixing, recycling, the recovery of old metal and other processes as creating lead isotope ratio patterns that are only tangentially linked to a simplistic notion of ‘source’. This does not mean that lead isotope studies in archaeology are worthless – just that, as with trace elements, the simple goal of linking object to source is unlikely to be successful. Hindsight suggests that we may have been asking the wrong question all along. By focusing more on the changes that show up in the chemical and isotopic record in the metals themselves, we may see something that directly reflects changes in some aspect of human behaviour. This might be the result of changes in patterns of ore exploitation, but it may also be a technological change in extractive metallurgy, or changes to the pattern of metal circulation and recycling. Whatever the cause, the result is perhaps of more direct archaeological interest, ultimately, than knowing that the metal from which a particular object was made came from a particular mine site. This overall approach to the chemical composition of copper alloys described above has been, unsurprisingly, not welcomed in all circles, since it negates or The archaeological bazaar: scientific methods for sale? 125

2.12 Ni > 0.5% Ni < 0.5% 2.11

2.1

2.09 208Pb/206Pb

2.08

2.07 0.825 0.835 0.845 0.855 0.865 0.875 207Pb/206Pb

FIGURE 7.2 Lead isotope ratio values for some Taunton Phase metalwork (data from Rohl and Needham 1998). This assemblage comprises the IMPLI 12 and 13 objects (see definition in Rohl and Needham 1998: 96), all of which we classify chemically as copper types 11 or 14 (see Table 7.1), divided into high-nickel and low-nickel types. Those objects with lower nickel seem to occupy the extremities of the distribution, which we interpret to be the result of mixing. contradicts much (although certainly not all) that has gone before. Apart from being condemned as setting back archaeometallurgy by several decades, it has been specifically described as ‘unscientific’. If incorporating the information provided by rigorous pursuits such as thermodynamics, chronology, typology and geography is deemed to be unscientific, then it is indeed ‘unscientific’. We prefer to see it as better science. By bringing together as much data, explanatory power and context as possible to answer technological questions, it fits extremely well with discussions of technology in other arenas over the last 30 years, such as science, technology and society (STS) studies, agency, materiality and so on. These schools of thought are based on the realisation that technology is not merely a set of material processes but also encompasses the human decisions and structures that surround them. By giving equal weight to a thorough discussion of chemistry, materials science and the archaeological context, we aim to produce a model of past metallurgy that is stronger than a narrowly defined ‘scientific’ approach. The term ‘archaeometry’ was coined by Christopher Hawkes in Oxford in the mid-1950s to denote the application of the physical sciences to archaeological problems, and became the title of a journal. If a separate term is necessary, many now prefer ‘archaeological science’ to denote the broader application of the sciences to archaeology. Sadly, in some circumstances, the ‘-metry’ has come to dominate the ‘arch-’ – a situation common in modern Western society, where science has 126 Mark Pollard and Peter Bray come to be regarded (and occasionally to regard itself) as a superior form of knowledge. One key issue in archaeology, which is of course a fundamentally multidisciplinary activity, is how an ‘archaeologist’ evaluates the information provided by a ‘scientist’, especially if it comes in a form that does not invite question or dialogue. The answer is simple. If the information so proffered does not fit with all the other archaeological information, then it should be regarded as suspect – at best, put in the box labelled ‘interesting if true’. Most people involved in archaeology agree that working together across the ‘science–humanities’ divide is desirable, and indeed archaeology provides some of the best fora for doing so. At the practical level, however, beyond handwringing and saying we should work together more effectively, there has often been little real movement in that direction. Our new concepts for interpreting archaeological metalwork – based on the expectation that humans have manipulated metal a number of times between making and burying, and using the principles of material science, chemistry, biography, prosopography, typology, chronology and mapping – are, we suggest, a step in that direction.

References

Ambert, P. (1995) ‘Les mines préhistoriques de Cabrières (Hérault): quinze ans de recherches. État de la question’, Bulletin de la Société Préhistorique Française, 92: 499–508. Baxter, M. J., Beardah, C. C. and Wright, R. V. S. (1997) ‘Some archaeological applications of kernel density estimates’, Journal of Archaeological Science, 24: 347–354. Bray, P. J. and Pollard, A. M. (2012) ‘A new interpretative approach to the chemistry of copper-alloy objects: source, recycling and technology’, Antiquity, 86: 853–867. Damour, A. (1865) ‘Sur la composition des haches en pierre trouvées dans les monuments celtiques et chez les tribus sauvages’, Comptes Rendues Hebdomadaires des Séances de l’Académie des Sciences’, 61: 313–321, 357–368. Harding, A. (2000) European Societies in the Bronze Age, Cambridge: Cambridge University Press. Johnson, A. E. (2012) Solving Stonehenge, London: Thames and Hudson. Junghans, S., Sangmeister, E. and Schröder, M. (1960) Metallanalysen kuperzeitlicher und frühbronzezeitlicher Bodenfunde aus Europa, Studien zu den Anfangen der Metallurgie 1, Berlin: Gebr. Mann. Junghans, S., Sangmeister, E. and Schröder, M. (1968–1974) Kupfer und Bronze in der Frühen Metallzeit Europas, 3 vols, Katalog der Analysen nr 10 041–22 000 (mit Nachuntersuchungen der Analysen nr 1–10 040), Berlin: Verlag Gebr. Mann. Krause, R. (2003) ‘Studien zur kupfer- und frühbronzezeitlichen Metallurgie zwischen Karpatenbecken und Ostsee’, Vorgeschichtliche Forschungen, Bd 2, Berlin: Freie Universität. McKerrell, H. and Tylecote, R. F. (1972) ‘Working of copper-arsenic alloys in the Early Bronze Age and the effect on the determination of provenance’, Proceedings of the Prehistoric Society, 38: 209–218. Ottaway, B. (1982) ‘Earliest copper artifacts of the northalpine region: their analysis and evaluation’, Heft 7, Bern: Schriften des Seminars für Urgeschichte der Universität Bern. Otto, H. and Witter, W. (1952) Handbuch der altesten vorgeschichtlichen metallurgie, Leipzig: Mitteleuropa. The archaeological bazaar: scientific methods for sale? 127

Pollard, A. M. (2009) ‘What a long strange trip it’s been: lead isotopes in archaeology’, in A. J. Shortland, I. C. Freestone and T. Rehren (eds) From Mine to Microscope – Advances in the Study of Ancient Technology, Oxford: Oxbow Books, 181–189. Pollard, A. M. (2013a) ‘“Comin’ in on a wing and a prayer”: archaeological chemistry since 1790’, in R. A. Armitage and J. H. Burton (eds) Archaeological Chemistry VIII, American Chemical Society Vol. 1147, Washington, DC: American Chemical Society, 451–459. Pollard, A. M. (2013b) ‘From bells to cannon – the beginnings of archaeological chemistry in the eighteenth century’, Oxford Journal of Archaeology, 32: 333–339. Pollard, A. M., Bray, P. J. and Gosden, C. (in press) ‘Is there something missing in scientific provenance studies of prehistoric artefacts?’, Antiquity. Rohl, B. and Needham, S. (1998) ‘The circulation of metal in the British Bronze Age: the application of lead isotope analysis’, Occasional Paper 102, British Museum, London. 8 RADIOCARBON DATING AND ARCHAEOLOGY

History, progress and present status

Sturt W. Manning

The radiocarbon revolution: beginnings The invention of the radiocarbon (14C) method (Libby et al. 1949) provided the first opportunity for practical absolute dating of a range of organic materials from the recent to the distant past (to around 50,000 years ago). In its original conception, it was assumed that the concentration of 14C in the atmosphere had remained constant in the past. Dates were run on a variety of known age (or approximately known age) samples from around the world in the early years of the method (e.g. Arnold and Libby 1949, 1951; Libby 1951). At first it seemed, within the (large) measurement errors then possible, that the 14C dates on samples of known age (tree rings), or of historically accepted age, yielded satisfactory dates – as expressed in the famous ‘curve of knowns’ (Arnold and Libby 1949: Figure 1). The early 14C dates on non-known-age prehistoric test samples provided the first non-guesswork dates for a myriad of topics worldwide, ranging – to give a few examples – from the termination of the last Ice Age, to the timescale of the prehistory of North America (Libby 1961: 626–627), to the first direct dates for a number of prehistoric contexts ranging from sites in Europe (Arnold and Libby 1951: 112–113) to Australia (Libby 1951: 295). This was sensational; to take just the last case, 14C dates rapidly and progressively demonstrated the very considerable time-depth of Australian prehistory before European contact (e.g. Gillespie 2002). Radiocarbon dating was a revolution, which entirely restructured the practice and understanding of prehistoric archaeology around the world (e.g. Renfrew 1973; Taylor 1987) – it provided an independent universal timeframe. Radiocarbon was exciting, but safe, as it largely filled a vacuum (there was a short-lived debate over the Neolithic in the Near East as dates, notably for Jericho, were much earlier than some anticipated, but this quickly died away and everyone accepted that dates had previously been more or less guesswork before about 3000 BC – Renfrew 1973: 61–62). Where there had been nothing (no direct timescale for prehistory), there Radiocarbon dating and archaeology: history, progress and present status 129 were now approximate dates creating a chronological framework and scale for world prehistory; where there were supposed historical chronologies then the very approximate radiocarbon dates initially available – the error ranges were typically plus or minus several centuries or more – could find approximate confirmation, but did not challenge the existing dates and culture-historical syntheses. However, the initial situation changed quite rapidly, both regarding the science and the archaeology. In a paper of 1958 de Vries observed from analysis of known- age tree rings that 14C production seemed in fact to vary over time (a result confirmed by Willis et al. 1960). A revised ‘curve of knowns’ in Libby (1961: Figure 4) saw 50 per cent of the samples placed as somewhat too old, with conjecture (Libby 1961: 625) over whether the half-life of 14C needed to be slightly longer (see also Kohler and Ralph 1961: 357–358) or a note that ‘there are other possible explanations of a deviation of the curve of knowns from the theoretical curve’). Libby two years later observed that 14C dates on Egyptian samples, while acceptable to around 4,000 years ago, were too recent before this; and similarly, that while dates on tree-rings were acceptable to around 2,400 years ago, they also were too recent – 3,000–4,000 years ago (Libby 1963). By the mid-1960s, the story started to become clear: data from several sources, and repeated measurements, demon- strated that the original assumption was incorrect, and instead that the 14C content of the atmosphere varied over time (e.g. Suess 1965, 1970). And so, 14C dates had to be ‘calibrated’ against a record of known past atmospheric 14C levels in order to achieve the correct calendar date estimate. Meanwhile, as more 14C dates appeared, they started to fill voids in many areas of the world, but, in those areas with relatively elaborate existing archaeological chronologies, there were now challenges: this applied especially in the Old World. Daniel captures the situation well:

But there is this great difference between dates in the New World and in the Old. In America dates are being produced for cultures and events which were at one time thought to be undateable, except by a rough and ready guesswork calculation back from dendrochronological dates. In Europe and the Near East the problem is quite a different one; it is that the radiocarbon dates are in many instances producing a chronology which, to put it mildly, seems at variance with archaeological dating. (Daniel 1959: 238)

Prehistoric Europe became the focus of debate. Here, based on different views (of routes and speed of diffusion from an Old World core in the east Mediterranean-Near East), there had been estimated dates placed against the known archaeological sequences (Thomas and Ehrich 1969: 145), with the stronger scholarly tradition leaning towards a relatively short chronology (e.g. Milojcˇi´c 1949) – although it is important to remember how very flexible those dates always were and the assumptions, constructs and questionable logic transfers involved (even a friendly contemporary critic – Childe – highlights this and writes that dates ‘are 130 Sturt W. Manning much more elastic than Milojcˇic´ thinks’ (Childe 1950: 291)). The new 14C dates indicated much earlier dates (by a millennium), and in each case favoured a higher (i.e. earlier) chronology. This led to a famous critique of the 14C method by Milojcˇi´c (1957), and a vigorous debate over several years, as summarized by Renfrew (1973: 62–65). New archaeological finds, indicating more, and more extensive, cultural phases than previously recognized, also suggested the need for longer chronologies, and so earlier dates (Thomas and Ehrich 1969: 147). This uncertainty of the 1950s–1960s was ultimately also resolved by the calibration of the 14C timescale, using known-age tree-rings. It was then clearly demonstrated which 14C ages equated with which calendar years, and, however carefully and lovingly crafted, the outcome was that the previous short European prehistoric chronologies were swept away (e.g. Ferguson et al. 1966, 1976; Renfrew 1970, 1973). To give one extreme example of how times have changed: as Renfrew (1973: 64 and n.28) summarizes, Piggott’s (1954) pre-14C chronology placed the start of the English Neolithic after 2000 BC, and Piggott (in)famously declared that early 14C dates in the twenty- seventh century BC are ‘archaeologically inacceptable’ (Piggott 1959: 289; Daniel 1959: 239), yet the recent large-scale and comprehensive 14C-based study of Whittle et al. (2011) now places the beginning of the (complicated and now much better understood) Neolithic of Britain and Ireland as starting just before 4000 BC in the South-East, and spreading to all areas before about 3700 BC!

Calibration The need for a secure known-age record of atmospheric 14C ages was clear, and the obvious source was known-age tree-rings. These could be used to build up calibration time-series of calendar-placed 14C measurements: ‘tree-ring cellulose is a direct sample of the atmospheric 14C level in the year of growth, preserving the information reliably over millennia-long intervals of deposition’ – and there has consequently been a close relationship between 14C and dendrochronology for six decades, with the development of long Holocene dendrochronologies in Europe especially, spurred by the need for 14C calibration back in time as far as possible (Kromer 2009: 15 (quote)). The recognition of variations in atmospheric 14C content led to a focus on, and progressive elucidation of, the causes and mechanisms involved: over the longer term, natural processes such as geomagnetic variations, solar variability, ocean processes and, in the recent period, fossil fuel release of old carbon (the so-called ‘Suess effect’) and impact of atmospheric nuclear explosions. As a result, not only have the relevant calendar dates from 14C been better defined, but 14C has become a highly relevant dataset for research on solar activity, climate, the carbon-cycle and other Earth systems (e.g. Stuiver 1961; Damon et al. 1978; Stuiver and Quay 1980; Stuiver et al. 1991; Stuiver and Braziunas 1993; Bond et al. 2001; Kromer 2009: 16–17; Vieira et al. 2011). Such science is now standard and can inform assessments of past climate and history (e.g. McCormick et al. 2012; Manning 2013b), but it is easy to forget that it was not yet all understood and clear even into the later 1960s (cf. Thomas and Ehrich 1969: 150–151). Radiocarbon dating and archaeology: history, progress and present status 131

Initially, it was unclear what form the calibration curve should take: was it a relatively smooth curve (and perhaps monotonic), or irregular and wiggly (non- monotonic)? The measurement precision available in the 1960s to early 1970s was not capable of resolving whether regular minor variations were an artefact of the measurement process, or real. Thus, whereas Suess (e.g. 1970) drew an irregular calibration curve trying best to fit the available data points by hand (his ‘cosmic schwung’), other smoother curves were proposed and could also be regarded as appropriate given the data then available (e.g. Clark 1975). It was only with the development of higher-precision 14C dating and the systematic measurement of times-series of known-age wood samples – and not just for the recent two millennia, where variations happen to be relatively modest when compared to other periods in the past few thousand years (Pearson et al. 1977) – that it finally became clear the answer was a wiggly, irregular, calibration curve (de Jong et al. 1979). The tree-ring calibration of the 14C timescale became the second radiocarbon revolution (Renfrew 1973). The first calibration curves relied on wood from North America, especially from Bristlecone pines (e.g. Suess 1970), but the very long- lived Bristlecone pines have typically very narrow rings (making it difficult to supply large samples), there are often many locally absent (missing) rings in any sample and they come from very high elevations (which, as later work has shown, may introduce small differences in the atmospheric 14C recorded versus low- elevation trees and plants as relevant to the majority of archaeological investigations – e.g. Dellinger et al. 2004). Thus the next major turning point and full expression of the second radiocarbon revolution was the development and use of European oak tree-ring records (see summary, Kromer 2009: 16), which could provide the large sample sizes needed for high-precision conventional (pre-accelerator mass spectometry, AMS) 14C dates. Allied with a focus on laboratory procedures and technology, in order to increase the measurement precision possible at what came to be called ‘high-precision’ 14C laboratories, this made it possible to build long high-precision 14C calibration curves (for reviews of calibration progress, see Hassan 1986; Taylor et al. 1996; Bronk Ramsey et al. 2006). A secure basis now existed for estimating the timescale of prehistory, and with much increased precision. Beyond about 2000 years ago, the calibrated dates provided noticeably older calendar ages than first thought from radiocarbon (and even more so from most pre-radiocarbon guesstimates). Figure 8.1 shows the divergence between the original Libby model and the current IntCal13 Northern Hemisphere 14C calibration curve (Reimer et al. 2013): by sheer chance there is little offset (or merely variation around no apparent offset) for the most recent two millennia (see Figure 8.1, inset), but then, progressively, there is a larger and larger significant deviation. Figure 8.2 shows, as an example of an early radiocarbon calibration curve, the Damon et al. (1974) calibration curve versus the current IntCal13 calibration curve, and both versus the Libby model. The calibration curve of almost 40 years ago shows the clear deviation versus the Libby model beyond about 2000 years ago. It also mainly either includes the 2013 calibration curve, or is not too far away (the Damon et al. record tends to be a slight under-estimation 132 Sturt W. Manning

of the 14C age typically when compared against IntCal13 – if this does not reflect a small laboratory offset/bias issue, then it may reflect the short, late, summer growing season for the very high elevation Bristlecone pines versus the overall later spring-through-summer growing season of the oaks in Europe employed for IntCal13 – compare with the findings and discussion of Dellinger et al. 2004). Thus the Damon et al. (1974) curve is relatively accurate. However, what has changed dramatically in the succeeding four decades is the availability of many more, and more precise, measurements, and hence today we have a considerably more precise (narrow) calibration curve, and so a greater ability to resolve calendar age ranges. The archaeological reception of the second radiocarbon revolution varied. Much depended on a proper application of 14C dating: focusing on what the samples actually dated and how this related (or did not) to archaeological contexts of interest (Waterbolk 1971), and the appropriate use of (and thinking about) the dating information available and the archaeological questions (Whittle 1988: 12–22). In areas and periods before some form of text-based history, the new dates were rapidly accepted by many, and, excitingly, were seen to push back the prehistoric record. The weight of evidence from all sides overwhelmed the pre- 14C positions. The general diffusionist paradigm, which considered the spread of

Outgoing

Outgoing Outgoing Outgoing Outgoing Outgoing Outgoing OutgoingOutgoing Outgoing OutgoingOutgoing Outgoing Outgoing Outgoing Outgoing

Outgoing Outgoing Outgoing

Outgoing OutgoingOutgoing Outgoing Outgoing Outgoing Outgoing Outgoing

FIGURE 8.1 Comparison of the original Libby 14C age expectation model versus the reality observed from known-age archives reflecting the variations in the levels of atmospheric 14C over time, as modelled in the current IntCal13 Northern Hemisphere 14C calibration dataset (Reimer et al. 2013). Inset: detail of the period from 3000 BC to AD 1950. Radiocarbon dating and archaeology: history, progress and present status 133

Outgoing Outgoing Outgoing Outgoing Outgoing Outgoing Outgoing Outgoing OutgoingOutgoing Outgoing OutgoingOutgoing OutgoingOutgoing Outgoing Outgoing Outgoing Outgoing

Outgoing Outgoing Outgoing Outgoing Outgoing OutgoingOutgoing

FIGURE 8.2 Comparison of the Damon et al. (1974) 14C calibration curve (from measurements of known-age Bristlecone pine) versus the IntCal13 14C calibration curve of four decades later (Reimer et al. 2013), and both versus the Libby model.

technology and influences in terms of a wave of change emanating from a core region, such as Egypt and the Near East, was overturned as it became clear that many other areas and developments were as old as, or older than, those in the east Mediterranean (Renfrew 1971, 1973). In particular, long-standing chronologies and entire archaeological histories and understandings based around ceramic (or other object) typologies and comparisons – and, inevitably, scholarly biases in assembling such interpretations – were up-ended. Today an independent timeframe for prehistory and proto-history from 14C is possible (sometimes supplemented or improved by other absolute dating methods, like dendrochronology), against which the archaeological evidence can be assessed. A nice example is the radical rethinking of the Carpathian Basin area, comparing the ceramic-based pre-14C situation versus assessment via recent sophisticated 14C dating and analysis – not only is the timescale very different (longer) than previously thought, but, critically, the material culture story is much more complicated (see Raczky and Siklósi 2013). The second radiocarbon revolution led directly to much new (and independent) thinking, especially in European-Mediterranean archaeology – in particular, emphasis was placed on the apparent autonomy or independence of key cultural changes in Europe, versus simplistic diffusionist thinking (e.g. Renfrew 1969, 1973: 150–271) – since, as Renfrew states, ‘. . . the basic links of the traditional chronol- ogy are snapped and Europe is no longer directly linked, either chronologically or 134 Sturt W. Manning culturally, with the early civilizations of the Near East’ (Renfrew 1973: caption, Figure 21). This was important (and a corrective), but, allied with the other associated themes of the ‘new’ or processual archaeology of the period, it also led to a generation of scholarship largely predisposed against looking at migrations, connections and influences and the related socially entangled interpretations of material culture and societal change (anything that could be labelled as diffusion). This was a negative, as the roles of human entanglements with material culture and ideas, and their stories of these, of connectivities and of movements (of things, ideas and people) are, on timescales short and long, also central to history (variously, see Helms 1988; Anthony 1990; Hodder 2012; Broodbank 2013). Sophisticated radiocarbon analysis can indeed even hope to trace such engagements and processes (Whittle et al. 2011). Meanwhile, in the Aegean and east Mediterranean, there began to be problems. Whereas, pre-calibration, the first 14C dates – e.g. Kohler and Ralph 1961 – were of interest, but were no challenge to conventional thinking and timeframes, calibration entirely changed this status. Now science was starting to encroach on long-standing scholarly orthodoxy linked to historical chronology. The second radiocarbon revolution thus highlighted a temporal disconnect with the east Mediterranean-Egypt-Near East region versus elsewhere, and so exacerbated the growing method–theory divide between much classical and Near Eastern archae- ology as against general anthropological archaeology during this period (e.g. Renfrew 1980; Snodgrass 1985). Even amongst those working with 14C in this period, it was assumed that 14C could not challenge the long-established historically based or derived chronologies in the Aegean-east Mediterranean-Near East. Rather, comparisons with these established historical-archaeological chronologies were held, at best, to provide approximate support for the validity of calibrated 14C dating as to be employed elsewhere (Clark and Renfrew 1973). If there was any disagreement between the 14C timescale and, for example, the historical chronology of Ancient Egypt, then it seemed obvious that 14C must be in error (and rarely at this time were 14C dates sufficiently precise to be a cause for concern). Renfrew indicated this situation and mode of thought with what became his infamous chronological faultline (1973: 115–117, Figures 20, 21):

there is a line running round the Aegean and east Mediterranean. Inside this arc, the dates in the third millennium B.C. are not much altered by the radiocarbon chronology and its calibration. But outside this line . . . every- thing is shifted several centuries earlier. (Renfrew 1973: 115–117)

However, this supposed stability in the Aegean was a question of scale. Yes – no millennial to many centuries changes were apparent as in Europe, but in fact, in the third millennium BC, the 14C dates did support a high chronology (compare Thomas and Ehrich 1969: 147–148); it was just that the adjustments were smaller and seemed compatible with a ‘high’ reading of the ancient Near Eastern semi- Radiocarbon dating and archaeology: history, progress and present status 135 historical evidence. The start of the Aegean Early Bronze Age had been placed after 2800/2700 BC (down to 2300 BC) by a range of the major scholars, whereas 14C- based assessments indicated a date around or before 3000 BC (Manning 1995: Table 1). And this was genuine pre-history – no one pretended there were good, close linkages between the Aegean and the historical worlds of Egypt and Mesopotamia at this time. Things could thus be stretched a little in the third millennium BC, just as long as nothing was moved from about 2000 BC onwards. Here it was stated that ‘the radiocarbon dating evidence for Aegean chronology after about 2000 BC is for the most part less precise than dates obtainable from the Egyptian correlations’ (Warren and Hankey 1989: 127). Nonetheless, as 14C dates accumulated during the 1970s, it increasingly appeared that there might be some conflicts between 14C dates and historical chronologies even in the Aegean and east Mediterranean region and into the proto-historical periods (e.g. Betancourt and Weinstein 1976; Mellaart 1979) – leading to the first hints of the start of a real and sharp divide between those prepared to consider the possibility that the 14C dates indicated a different but valid chronology for some parts of the Old World, and those who rejected such sug- gestions outright and instead maintained the conventional positions. This was a new mini-version of the radiocarbon revolution in Europe surveyed by Renfrew (1973) – just at the scale of decades to centuries, not many centuries to millennia. However, in general at this stage, neither the precision of the 14C calibration curve, nor the accuracy and precision of 14C measurements on archaeological samples, nor the relatively restricted availability of suitable large samples (as required for 14C in the era before the advent of accelerator mass spectrometry (AMS) dating) permitted such finegrained issues to be settled.

Better calibration, smaller samples The general acceptance of the need for the calibration of the 14C timescale (relevant to several fields, not just archaeology) prompted large-scale efforts to achieve a progressively better-defined tree-ring calibration for the past several thousand years (e.g. Damon et al. 1974; Pearson et al. 1977, 1983; de Jong et al. 1979; Stuiver 1982) – leading to the creation of a general Northern Hemisphere high-precision 14C calibration curve in 1986, which has in turn been progressively improved and extended to the present day (e.g. Stuiver et al. 1998; Reimer et al. 2004, 2009, 2013); see Figures 8.1 and 8.2. A separate Southern Hemisphere 14C calibration curve has also been developed (most recently Hogg et al. 2013). To achieve age estimates beyond the known-age tree-ring record, other near absolutely dated archives were brought to bear. Until recently, this primarily meant non-terrestrial sources, e.g. marine deposits and speleothems for the period before the Holocene. This was less than ideal for archaeology, since the 14C records from such sources are not the same as contemporary atmospheric, and so terrestrial 14C records – due to reservoir effects – and there were differences between the various datasets, which led to uncertainty and debate (e.g. van der Plicht et al. 2004; van Andel 2005; Mellars 2006a, 2006b; Turney et al. 2006; Bronk Ramsey et al. 2006). Thus a key 136 Sturt W. Manning breakthrough was the recent publication of a long, well-defined terrestrial 14C record reaching back just over 50,000 years (Bronk Ramsey et al. 2012) – this now forms the best available basis for archaeological dating prior to the tree-ring record. Today, as I write, the state-of-the-art global assessment of 14C calibration for the past 50,000 years is to be found in the set of papers in the IntCal13 special issue of the journal Radiocarbon (Volume 55, No. 4, 2013). The IntCal13 14C calibration curve is shown in Figure 8.3, with the insets illustrating some periods of time in more detail. Note the wiggly shape, reflecting the varying natural 14C levels through time. The current IntCal13 curve is compared in Figures 8.3 and 8.4 against several of the previous high-precision calibration curves (1986, 1998, 2004, 2009). For the last few thousand years, there has been little substantive change in the calibration record in most periods over the past quarter-century, despite additional data. However, in some earlier periods, and especially before the start of the Holocene (and the established tree-ring record), there have been some quite substantive changes even in the past few years. As

Outgoing(c) Outgoing (a) Outgoing Outgoing Outgoing Outgoing Outgoing Outgoing OutgoingOutgoing OutgoingOutgoing Outgoing (d) Outgoing Outgoing (b) Outgoing Outgoing OutgoingOutgoing Outgoing Outgoing OutgoingOutgoing OutgoingOutgoing Outgoing Outgoing Outgoing

OutgoingOutgoing Outgoing Outgoing Outgoing OutgoingOutgoing FIGURE 8.3 (a) IntCal13 (Reimer et al. 2013) (1σ errors) 14C calibration curve 0–50,000 years BP (from AD 1950) shown in years BC/AD (resultant dates cal. BC or cal. AD), versus IntCal09 (Reimer et al. 2009) with 1σ errors versus IntCal04 (Reimer et al. 2004) with 1σ errors versus IntCal98 (Stuiver et al. 1998) with 1σ errors versus the 1986 initial international high-precision calibration datasets (Pearson and Stuiver 1986; Pearson et al. 1986) with 1σ errors. In the Holocene period (so, to c.10,000 BC), the datasets largely entirely overlap each other; differences are more evident before this. (b) Comparison for the period 35,000 to 25,000 BC. (c) Comparison for the period 12,000 to 8500 BC. (d) Comparison for the period 1700 to 1200 BC. Radiocarbon dating and archaeology: history, progress and present status 137 illustrated in Figures 8.3b and 8.4, changes of 500–1,000 14C years have been made to the period between 25,000–35,000 BC – which means quite large changes (centuries to a millennium) in estimated calendar dates for the archaeological record at this time, with one example shown in Figure 8.5. This can have major impli- cations, and may require some reassessment of previous arguments (e.g. Mellars 2006a, 2006b; Turney et al. 2006) made on the basis of earlier and much less accurate (or secure) calibration curves. The period around the onset to the Holocene is another time when century-scale changes have been made (Figures 8.3c and 8.4). The Southern Hemisphere 14C calibration curve is generally similar in shape, but exhibits small, varying, differences – smallest at times of general cooling and increased 14C production (steep slopes downwards in the 14C calibration curve) and largest at periods of warming and decreased 14C production (reverses and plateaux in the calibration curve) – see Figure 8.6 – indicating an ice–ocean feedback is at work. In addition to an increasingly refined, accurate and precise 14C calibration curve, the 1980s saw the advent of AMS dating (Gove 1992). This permitted dating of dramatically smaller samples (and of specific fractions of samples – for example, of bones: McCullagh et al. 2010; Nalawade-Chavan et al. 2012), and especially made it at last routinely possible to date the often small amounts of short-lived sample matter (e.g. seeds) capable of offering directly relevant dating evidence for

Outgoing Outgoing Outgoing Outgoing Outgoing

Outgoing Outgoing

Outgoing Outgoing Outgoing OutgoingOutgoing Outgoing

OutgoingOutgoing Outgoing Outgoing OutgoingOutgoing Outgoing OutgoingOutgoing Outgoing OutgoingOutgoing Outgoing Outgoing Outgoing OutgoingOutgoing

FIGURE 8.4 Comparison of the differences of the mid-point values of the previous 14C calibration datasets (IntCal09, IntCal04, IntCal98 and the 1986 dataset) versus IntCal13 – data from Figure 8.1. Inset: Detail for the period 3000 to 1 BC. 138 Sturt W. Manning

OxCal V4.2.3 Bronk Ramsey (2013); r:5 32000 (b)

(a)(a) lntCal13 atm o spheric curve R _ D a te :’9 9 00 _i:i Straightforward. (Reim er e t a l : *013) or down-to-earth 31500

-32123 t >-31932 (8.2% Positive-32216 15 -attitude31840 <5.4% )

P f t_Date 2 9! I00_09

B 31000 (

n io t a

in -3277:5 t o -32621 68.2°/» m r Positive-3287 I t o -3 2attitude5 40 95.4°/ > te 30500 e d

n 33500 33000 32500 32000 31500

o IntCalOS iitm ospheric curve \ b r (Reimer e t al 2009) Calibrated date (calBC) ca io

d R _D ate 29 90 0 _ 13 __ 30000 Ra

R _D ate 29900_09

29500

29000

Straightforward3350Straightforward0 33000 or 32down-to-earth50 or0 down-to-earth32000 3150 0 Calibrated date (calBC)

FIGURE 8.5 The (a) calibrated calendar probability distributions and (b) calendar age ranges at 68.2 per cent and 95.4Positive per cent attitudeprobability for the 14C age of 29,900 ± 35 BP, comparing between the IntCal13 and IntCal09 14C calibration curves. archaeological find contexts or objects (Waterbolk 1971). This was a(nother) great revolution in 14C dating (for histories of 14C through to AMS, see Taylor 1997, 2000). Previously the large sample requirements usually meant dating large charcoal samples from archaeological sites (whether one piece or several pieces combined), and inevitably this meant dating tree-rings that were older or much older than the archaeological context of interest (the issue of the ‘dated event’ versus the ‘target event’ (Dean 1978), and the general ‘old wood’ problem (Schiffer 1986)). The small-sample revolution, however, has provided a means (improved and now near high precision (Bronk Ramsey et al. 2004)) routinely to date archaeological contexts of interest directly, or at least to minimize the terminus post quem by dating only outer tree-rings or bark from charcoal or wood samples. Today, most archaeo- logical dating projects, prehistoric to historic, therefore focus where possible on dating suitable short-lived sample material that is contemporary with the archaeo- logical context of interest, or on dating the most chronologically specific samples available, ideally as part of a sequence of such data – stratigraphically or historically defined – at a site or set of linked sites/contexts (e.g. Manning et al. 2006; Bayliss et al. 2007; Higham and Higham 2009; Bronk Ramsey et al. 2010; Whittle et al. 2011; Lorentzen et al. 2014). For a brief, comprehensive summary of current 14C dating methods, analytical procedures (how to get calendar dates) and some of the issues and potentials, see Bronk Ramsey (2008). Radiocarbon dating and archaeology: history, progress and present status 139

In tC a M 3 ± la S H CaM 3 ± la 3000 (a)(a) 1100 1000 900 2500 - 800 (c)(c) 700 600 ) 2000 ■ 500 P

B 400

s 1000 1100 1200 1300 1400 1500 r a e 1500 (b)

Y (b) (

e 11000 g A 1000

C ' 10500-

500 10000

-11000 -10500 -10000

0 —|— i—

-1000 -500 ■1/1 500 1000 1500

Calendar Date BC / AD FIGURE 8.6 (a) Comparison of the IntCal13 (Northern Hemisphere: Reimer et al. 2013) and SHCal13 (Southern Hemisphere: Hogg et al. 2013) 14C calibration curves 1000 BC to AD 1940, comparing the 1σ ranges of each. (b) Comparison of IntCal13 versus SHCal13 (1σ) for the maximum (2σ limits) period of the Younger Dryas climate interval as represented in Greenland Ice Core records (Blockley et al. 2012: Table 1). (c) Detailed view of the period AD 1000–1500.

As a result of these two developments – higher-precision calibration curve and AMS dating of small samples – it was now beginning to be plausible to address seriously the question of whether or not 14C and historically based chronologies in the east Mediterranean and Near East are compatible. Initially, the new high- precision calibration curves were employed to calibrate the existing relatively low- precision 14C dates. This indicated in some cases that the (higher-precision) conventional historically derived or estimated dates were compatible with 14C – for example, for the third millennium BC and the Early Bronze Age in the Aegean (Manning 1995) and Cyprus (Manning and Swiny 1994), and for the end of the Late Bronze Age in the Aegean and Cyprus (Manning and Weninger 1992; Manning et al. 2001). However, it also continued to indicate apparent problems in some cases – with the date of the early Late Bronze Age volcanic destruction level buried by the Thera/Santorini volcano in the southern Aegean perhaps the main focus of debate (e.g. Betancourt 1987; Manning 1988, 1995: 200–216). Subsequently, programmes to acquire new higher-precision 14C dates from contexts in Egypt, the Levant, the Aegean and the central-west Mediterranean have started to produce much new data where there seem to be issues. Some examples include: Ancient Egypt (Bronk Ramsey et al. 2010; Dee et al. 2013); Early–Middle Bronze 140 Sturt W. Manning

Age Italy (Alberti 2013); the chronology of Iron Age Israel (Finkelstein and Piasetzky 2010; Finkelstein 2011; Mazar 2011; Lee et al. 2013); the chronology of the Aegean and central-west Mediterranean Iron Age (van der Plicht et al. 2009; Fantalkin et al. 2011; Bruins et al. 2011; Toffolo et al. 2013; Wardle et al. 2014), and the chronology of the Early through Late Bronze Age in the Aegean (e.g. Manning et al. 2006; Wild et al. 2010; Renfrew et al. 2012), to which we return in the next section. While there is by no means resolution of all debates, new data have crystallized the issues, and in several cases narrowed or clarified gaps. The truly notable aspect of recent work, despite a few holdouts (e.g. Bietak 2003), is that sophisticated 14C analysis, integrated with the archaeological evidence, is now being used by all sides in many debates as the primary basis for chronologies (and as the framework of history) – even inside the area of Renfrew’s supposed chronological faultline – as opposed to archaeological linkages with Egypt. Thus European-Mediterranean archaeological timescales have been reunified for the first time since the 1950s (another revolution).

Bayesian chronological modelling and the latest radiocarbon revolution The most recent key development in 14C chronology has been the development of Bayesian chronological modelling methods and tools (e.g. Bayliss 2009; Bronk Ramsey 2009a; Buck and Millard 2004). This approach allows the integration of archaeological and other information about context, especially a site (or wider) stratigraphic or other sequence, with the 14C dates and the 14C calibration curve, in order to estimate the overall relevant calendar dating probabilities as precisely and accurately as possible. These dating probabilities pertain to the actual 14C data in the model, but also to other periods, intervals or events within the archaeological sequence – such as the time between dated contexts, or the span represented by a set of dates. Data quality (detection of outliers or other problems) and coherence may also be quantified and tested (Bronk Ramsey 2009b). Bayesian chronological modelling therefore permits us to develop an archaeologically (and socially) informed fine-grained chronology, integrating places, people and things (e.g. Bayliss et al. 2007; Whittle et al. 2011) – see Bayliss and Whittle, this volume (Chapter 12). Bayesian chronological modelling is starting to be widely employed to test, improve, challenge and replace pre-existing chronologies where these were based either on less comprehensive 14C analyses, or on problematic archaeological, historical or other interpretations (for some examples, see Manning et al. 2006; Bayliss et al. 2007; Higham and Higham 2009; Bronk Ramsey et al. 2010; Kutschera et al. 2012; Ogburn 2012). It also enables the tightly resolved quantification and investigation of temporal relationships (and possible linkages) between archaeological contexts, change horizons and causal factors like climate- change episodes (e.g. Weiss et al. 2012; Manning 2013a). Bayes’s theorem had been around since the eighteenth century, but it was only with the availability of affordable, high-performance computing that it could be applied to archaeology, starting with initial publications from the early 1990s (Buck et al. 1991; Bronk Ramsey 1995), through to its now standard and powerful role in radiocarbon dating analysis and chronology building in archaeology today. I review two case studies Radiocarbon dating and archaeology: history, progress and present status 141 from the Aegean to illustrate the power and potential of such analyses, but also some of the issues (for other examples, see Bayliss and Whittle, this volume: Chapter 12). The analyses here employ the OxCal software (Bronk Ramsey 1995, 2009a) and the IntCal13 14C calibration dataset (Reimer et al. 2013).1 OxCal terms like Phase, Boundary, etc. are capitalized.

EARLY CYCLADIC CASE

The chronology of the Early Cycladic period – third millennium BC southern Aegean, well known since Renfrew (1972) – offers one example of recent work that shows potentials but also some problems as we increase our chronometric resolution. Renfrew et al. (2012) presented a set of new 14C dates in a stratigraphic- typological sequence, and, combined with some previously published dates, constructed a Bayesian model to produce a refined Early Cycladic chronology. The Bayesian modelling roughly reduces the dating range from the non-modelled dates by about half (cf. Renfrew et al. 2012: Figure 7 vs Figures 8 and 9). But this new precision also brings other issues into focus. One particular topic of interest is the date of the end of Dhaskalio Phase C and the transition to the next phase – since this represents the end of the Kastri Group Phase in the Cyclades, and potentially is the time of a much-debated ‘gap’ in the Cycladic sequence (Rutter 1983; there is much discussion of the topic in a set of seven papers in honour of Jeremy Rutter – see American Journal of Archaeology, 117 (2013): 527–597). Renfrew et al. (2012: Table 6) offer a date of about 2300 BC and minimize the evidence for the subsequent ‘alleged “gap”’ (Renfrew et al. 2012: 157). But it is important to consider exactly what is being dated and the likely archaeological contexts and assumptions involved. Sometimes even quite small changes in a Bayesian model can have substantial effects (for one discussion and some illustrations, see Mazar and Bronk Ramsey 2008). Three dates are on seeds – from a single group of seeds – and so should offer ages contemporary with their context and may be combined to give a best date estimate; seven dates are on animal bones (not stated, but likely sheep/goat) and so likely have only a few years of in-built age, but the rest are on charcoal. Renfrew et al. state that the charcoal is ‘often from pruned branches and twigs . . . We think therefore that the growth of the original wood should be within a decade or two of the deposition of the charcoal’ (Renfrew et al. 2012: 149, 151). But they also note that ‘the possibility that some of the olive wood samples may have derived from roof timbers cannot be excluded’. However, a number of the samples comprise Olea europaea or Juniperus (assumed phoenicea) and both of these can be (very) long-lived trees, sometimes the latter can have substantial in-built age with narrow rings, even in a branch. It is noticeable if one compares the ages on the two charcoal samples from shorter-lived species (R7, Hedera helix, common ivy; R51, Ericaceae, the heath/heather family of plants) that they are among the more recent for their contexts (see Figure 8.7). This does indicate 142 Sturt W. Manning

OxCal V4.2.3 Bronk Ramsey (2013) r:5 IntC all3 atmospheric curve fReimer et al 2013 Ea ly Cycladic S ta rt M o rk i — [Ffiase Kanpos Group — C. xA-329'/ a tr ------C xA-400i ab - - - CxA-400a ab — ; Kampos Group to Keros-Syms- ase Ker DS-Syros; xA-3292Z ab — xA-3294A abb ------xA-329,t ab — xA-3296 42, OxA 22754 45, OxA 22757 ros-Syros to Earlier Kast ase Ear ier Kastfi 4, OxA-.22745 7, OxA:22746 (Hedera helix) 8, OxA-.2 2747 41, OxA -22753 43, OxA 22755 44, OxA 22756 rlier Ka$tri to Dh askalio Phase C ase Phi se C 11, OxA 22748 13, OxA -22749 R 15, OxA -22750 R 19, OxA -22751 R 26, OxA -22761 R 46, OxA -22758 R51, OxA -22760 (Ericaea e) Phase C fcj> early fid d le Cy 4adic Pti.ase Mid die Cycl^i' dic O xA-118 19 seeds O xA-1217 3 seeds CxA-1211 4 seeds End Mode,

3600 3400 3200 3000 2800 2600 2400 2200 2000 Modelled date (BC)

FIGURE 8.7 Early Cycladic sequence as published in Renfrew et al. (2012: Figure 9). The dark-shaded histograms are the modelled age ranges (the lines underneath indicate the 68.2 per cent and 95.4 per cent probability ranges respectively); the light-shaded histograms are the non-modelled individual calibrated age ranges. All dates are on long-lived tree charcoal unless stated otherwise. The ‘ab’ samples are animal bones. The two dates on short/shorter- lived plant matter are indicated in bold and the arrows point to the calendar date ranges – note especially that these two light-shaded non-modelled ranges are among the most recent versus the other charcoal dates in each phase. Radiocarbon dating and archaeology: history, progress and present status 143

that at least some of the other charcoal samples contain substantial in-built age (either a number of tree-rings, or they comprise inner tree-rings, or they are re- used/residual material). Thus one might reasonably consider that the charcoal samples on the long-lived tree charcoal (i.e. all but samples R7 and R51) are typically something like perhaps 20±20 years older than their use contexts (this is merely a plausible average estimate). Further, we can assume that such charcoal samples may include some much older material (inner tree rings, re- used wood), but all should be at least slight terminus post quem estimates for their use dates. This situation can be modelled with the Charcoal Outlier model in OxCal (Bronk Ramsey 2009b: 1030–1032). If we now re-run the Renfrew et al. (2012: Figure 9) model, adjusting the long-lived tree charcoal dates as noted and employing the Charcoal Outlier model, we get the modelled age range shown in Figure 8.8b for the modelled Boundary Dhaskalio Phase C to Middle Cycladic, versus the range from the Renfrew et al. (2012: Figure 9) analysis shown in Figure 8.8a. We see that the Boundary quantifying the time period from the end of Dhaskalio Phase C through to the subsequent Middle Cycladic no longer centres c.2300 BC, but rather starts at 2337 BC and runs to c.2168 BC (95.4 per cent probability range), with most probability in the twenty-third century BC, and that the most likely (70.6 per cent) probability range (the most likely sub-range of the 95.4 per cent probability range) for the early Middle Cycladic sample set is 2214–2137 BC. Even use of a very modest 10±10 years’ adjustment finds 69 per cent of the dating probability lying after 2300 BC, and, indeed, even making no adjustment but simply using the Charcoal Outlier model yields a >64 per cent probability of a date after 2300 BC. In all cases the data are more consistent with a twenty- third century BC transition. What about the putative Rutter Gap? Renfrew et al. (2012: Figure 9), by placing all the data in one coherent sequence and not anticipating any time interval between the Kastri Phase and the Middle Cycladic Phase, tends to push the two phases towards each other and effectively finds what it supposes: no gap. In contrast, if one assumes there might be a missing time interval – the Rutter Gap – and re-runs the original or revised model, then, of course, one can find just such a possible gap period in the twenty-third century BC (see Figures 8.9a and 8.9b). The outcome of these analyses is to illustrate the analytical and interrogative potentials of Bayesian chronological modelling, but also the need to create an archaeologically appropriate model, and to ask apposite questions. In this Cycladic case, it appears that what our available information highlights is that there is a period of uncertainty in the twenty-third century BC. The Middle Cycladic Phase needs not start until around or after 2200 BC. This situation is not unexpected. Across the east Mediterranean and Near East there is a pattern of endings, decline and destruction in the twenty-third century BC, lasting to about 2200 BC (Wiener 2013), and of gaps or new beginnings after 2200 BC in 144 Sturt W. Manning

(a) (a) OxCal v4.2.3 Bronk Ramsey (2013); r:5______Phase C to early Middle Cycladic Boundary 68.2% probability

2337 (68.2% ) 2261BC 95.4% probability

2384 (95.4% ) 2193BC

Probability b e fo re : after y t i s n 0.01 e d

y it

il 0.005 b a b ro

P 0 I______I I______I

2400 2300 2200 2100

Modelled date (BC) ((b)b) OxCal v4.2.3 Bronk Ramsey (2013); r:5 Phase C to early Middle Cycladic Boundary 68.2% probability

2318 (57.4% ) 2249BC

2233 (4.7% ) 2222BC 2216 (6.2% ) 2202BC

95.4% probability

2337 (95.4% ) 2168BC

ity 0.01 Probability b e fore : after 2300BC = 17% : 83% dens

ty i 0.005 l i b a b ro

P 0

2400 2300 2200 2100

M odelled date (BC) FIGURE 8.8 The calibrated age distributions for the Phase C to early Middle Cycladic Boundary from (a) Renfrew et al. (2012: Figure 9), and (b) a re-run of the Renfrew et al. model but, as described in the text, adjusting the dates on charcoal from long-lived trees by 20±20 14C years, employing the Charcoal Outlier model (for the charcoal dates), and combining the three seed dates all on seeds from the same group of seeds into a weighted average. Radiocarbon dating and archaeology: history, progress and present status 145

(a)(a) (b)(b) OxCa V4.2.3 Bronk Ramsev (2013): r:S lntCall3 atmospheric curve IReimer et al 2013) Oxpa v4.2.j BrQnk Ramjgy (2p1?); r;5 Intg^li? a|mojph?ri<; wrvs (Reimer ?t al 2p1g)

Phase Kam os Group Phase Kami 10s Group

Boundary Krnpos Group Boundary Ki impos Gr oup o Keros-■ >yros o Keros-: iyros

Ph ase Keros-Syros Phase Keros-Syros

Be unclary K iros-Syro; to Eartie r Kastri Be undary K iros-Syro> to Eariie r Kastri

Phase Earlier Kastri Phase Earlier Kastri

Be undary E irtier KasI ri to Dhas kalio Pha ;e C Be undary E 1 riier Kast ri to Dhas kalio Pha ;e C

Phase Phas 5 C Phase Phasa C

Be undary Ei id Phase C Be undary Ei id Phase C

Boundary R itter Gap' ’ Be undary R itter Gap ’

Be undary Si art Middle1 Cycladic Be undary Si art Middle Cycladic

Phase Midd e Cycladi: Ph ase Middle Cycladi;

Be undary Ei id Model Be undary E id Model

3600 3400 3200 3000 2800 2600 2400 2200 2000 3600 3400 3200 3000 2800 2600 2400 2200 2000 Modelled date (BC) Modelled date (BC)

FIGURE 8.9 (a) Re-run of the Renfrew et al. (2012: Figure 9) model but adding in a putative Rutter Gap period between the end of Dhaskalio Phase C and before the start of the Middle Cycladic phase. (b) Re-run of the revised model, as in Figure 8.8b but adding in a putative Rutter Gap period between the end of Dhaskalio Phase C and before the start of the Middle Cycladic phase.

the twenty-second century BC. Cyprus offers one such example (Manning 2013a). This pattern is likely related to the well-known 2200 BC climate change episode (see, e.g., Weiss et al.1993; Weiss et al. 2012: 185–187; Weiss 2014; Dalfes et al. 1997; Staubwasser and Weiss 2006: 380–383; Salzer et al. 2013), and the associated end of the Anatolian trade network world (S¸ahogˇlu 2005, 2011) that included both the Aegean (Kastri group) and Cyprus (Philia facies) as peripheral elements.

AEGEAN EARLIER LATE BRONZE AGE CASE AND THE DATE OF THE THERA/SANTORINI ERUPTION

The date of the beginning of the Late Bronze Age in the Aegean has been a topic of much debate for several decades, centred especially around the date of the enormous eruption of the Thera volcano, which occurred (archaeologi- cally) in the later Late Minoan (LM) IA ceramic period and buried a whole civilization, Pompeii-like, on Thera (main site Akrotiri) (e.g. Doumas 1983; 146 Sturt W. Manning

Friedrich 2009; Manning 2014). The conventional (pre-14C) chronology built from interpreting material culture and stylistic linkages and associations between the Aegean and Egypt (and linking the Egyptian contexts with the Egyptian historical chronology) placed the eruption about 1500 BC, but 14C indicated a date around 100 years earlier (Betancourt and Weinstein 1976; Manning 2014). Which position is correct? This is a matter of key relevance to the correct synchronization, and so history, of the civilizations of the Aegean and east Mediterranean in the mid-second millennium BC. Let us look first at the published 14C dates on short-lived plant material from the volcanic destruction level at Akrotiri on Thera, most of which should offer age estimates very shortly before the volcanic eruption that buried the site. Up until recently, the method of analysis was to note that the samples all came from the same approximate archaeological horizon, and to treat a weighted average value as the best age estimate (the assumption being all the data were random estimates of the same real age). In most cases this is likely reasonable, especially for the stored seeds, since preservation circumstances and palaeo- botanical investigation suggest they had not been stored more than a short time (e.g. between zero and two years; see Panagiotakopulu and Buckland 1991; Sarpaki 2001: 28–30). However, it is also possible that a little more time is represented overall by some of the samples (including a short-lived twig), and Bayesian chronological modelling offers an alternative and comprehensive analytical route. We can treat all the samples – so far 43 dates have been published on short-lived material from the volcanic destruction level (see Manning and Kromer 2011: Figure 1, number 3) run at various times over four decades and of varying accuracy and precision – as a Phase (a group of events but not ordered or necessarily all the same age). But since it is reasonable to assume that, although some of the samples may be older, most relate closely to the end of the Phase just before the eruption, we can employ a Tau_Boundary paired with a Boundary in the OxCal software to define an exponentially distributed set of events just before the volcanic destruction (Bronk Ramsay 2009a). And we can apply an outlier model (in this case I use the general outlier model of Bronk Ramsey 2009b) to assess data quality. The outcome of the analysis is shown in Figure 8.10. We see that just six dates (the three oldest and three youngest) are outliers and the set shows a strong central tendency. The end Boundary ‘volcanic destruction’ defines the end of the Phase of events comprising the volcanic destruction level, and so gives the date estimate for the volcanic eruption. Figure 8.10b shows this for the set of 43 dates run in Figure 8.10a, and Figure 8.10c shows the result for a re-run of the model excluding the six outliers. We see that the Akrotiri data provide a closely defined late-seventeenth-century BC date range. In addition to its ability to offer appropriate comprehensive analysis, the other great strength of Bayesian chronological modelling is to integrate a time Radiocarbon dating and archaeology: history, progress and present status 147

(a) OxCal V4.2.3 Bronk Ramsev (20131: r:1 IntCa 13 atmnosoheric curve (Reimer tal2013) (a) 'Sequence Tau_Bourdary Start OxCal V4.2.3 Bronk Ramsev (20131: r:1______(b)(b )' [Phase R_Date( 3980,70) [0:9 3/5] - Thera Volcanic Destruction Boundary R_Date( 3800,50) [0:9 3/5] Outliors 68.2% probability R Datef 3610,51) [0:91/5] 1629 (68.2%) 1611BC R_Date( 3490,80) [0:4/ 5] 95.4% probability R_Date( 3460,80) [0:3/ 5] 1642 (95.4%) 1595BC R_Date( 3453,52) [0:4/ 5] R_Date( 3437,54) [0:3/5] R_Date( 3430,90) [0:2/ 5] ■2- R_Date(3415,70) [0:2/5] c 0.04 r R_Date( 3400,31) [0:2/ 5] R_Date( 3395,65) [0:1/5] 2- : R_Date( 3390,65) [0:1/5J 1 0 02 ’ R_Date( 3390,32) [0:1/5] 2 i R_Date(3380,60) [0:1/5] o- 0 - R_Date( 3372,28) [0:1/5] R_Date(3370,70) [0:1/5] R_Date( 3367,33) [Orh 5]------1650 1600 1550 1500 1450 R_Date( 3360,60) [0:1/5] R_Date( 3348,31) [0:1/5] Modelled date (BC) R_Date( 3340,65) [0:1/ 5] R_Date( 3340,55) [0:1/ 5] R_Date( 3340,60) [0:1/5] R_Date( 3339,28) [0:1/ 5] (c)(c) R_Date(3336,34) [0:1/5] OxCal v4.2.3 Bronk Ramsev (20131: r:1______R_Date( 3336,28) [0:1/ 5]------Thera Volcanic Destruction Boundary R_Date( 3335,60) [0:1/5] 68.2% probability R_Date( 3330,50) [0:1/ 5] 1630 (68.2%) 1614BC R_Date( 3322,33) [0:1/5] 95.4% probability R_Date(3321,32) [0:1/5] R_Date( 3320,50) [0:1/5] 1640 (95.4%) 1605BC R_Date(3318,28) [0:1/5] R_Date( 3317,28) [0:1/5] .§• . R_Date( 3315,31) [0:1,5] R_Date( 3310,65) [0:1/ 5] | 0.04 i- R_Date(3310,70) [0:1/5] R_Date( 3310,60) [0:1/5] Jo.02- R_Date( 3300,50) [0:1/ 5] R_Date( 3280,65) [0:2 5] £ 0L R_Date( 3250,50) [0:4/ 5] R_Date( 3245,65) [0:3/ 5] R_Date( 3140,70) [0:31/5] 1650 1600 1550 1500 1450 R_Date( 3130,50) [0:9 3/5] Oul .______Modelled date (BC) Boundary Thera Volcanic Destruction 4000 3500 3000 2500 2000 1500 1000 Modelled date (BC)

FIGURE 8.10 (a) Analysis of the 43 14C dates published on short-lived material from the Akrotiri volcanic destruction level grouped as a Phase with a Tau_Boundary as the start and a Boundary as the end (= estimate of the date of the immediately subsequent volcanic eruption). Six dates are identified as outliers employing the General Outlier model of Bronk Ramsey (2009b), with Posterior values greater than the Prior of 5. (b) Detail of the end Boundary ‘Thera volcanic destruction’ in (a). (c) The end Boundary ‘Thera volcanic destruction’ from a re-run of the model in (a) but excluding the six outliers. Calibration curve resolution set at 1 year.

series of archaeologically sequenced data (or other sequences, such as a tree- ring sequence; see Galimberti et al. 2004) into one comprehensive Sequence analysis. In the case of the earlier Late Bronze Age Aegean, we can consider such an analysis of data in a sequence from before, around the time of and after the Thera eruption, both to estimate the date of the eruption and to test the finding made above considering solely the dates from the Thera volcanic destruction level. The study of Manning et al. (2006) presented such an analysis to try to achieve greater clarity. The integration of the archaeological sequence with the 14C data enabled much greater chronological resolution than is possible by looking merely at the individual dates or at the individual contexts. We can now revise the original OxCal model for the Late Minoan IA through Late Minoan II periods, adopting the following steps: 148 Sturt W. Manning

1 Update the analysis to the current 14C calibration dataset: IntCal13 (Reimer et al. 2013). 2 Add in data available from another Late Minoan IB site (Mochlos; see Soles 2004). 3 Refine the intra-Late Minoan IB sequence according to the sub-phasing of the period by Rutter (2011). 4 Where dates were run on pairs of identical samples, use the weighted average values. 5 Employ Tau_Boundary paired with Boundary models for the two Late Minoan IB cases where the short-lived samples do not come from a clear single defined site destruction – but may be assumed mainly to date close to the end of the archaeological phase at the site (and hence an exponential distribution of the data for the Phase towards the end of each site Phase is the appropriate model). 6 Use the 37 date non-outlier set of Akrotiri VDL data and the Tau_Boundary paired with a Boundary model as in Figure 8.10 as a good estimate of the date of the Akrotiri volcanic destruction level (noting that we would get a very similar result if we use a qualitatively more selective sub-set; see Manning 2014: Figures RE9, RE36A).

Figure 8.11a shows the calendar age probability distributions for the Late Minoan IA to II periods. The integrated model – the archaeological sequence – allows much greater resolution and resolves what were otherwise ambiguous overlaps of the non-modelled data: compare the non-modelled light-shaded histograms with the modelled dark-shaded histograms. Figure 8.11b shows some details, including the placement of the later Late Minoan IA volcanic destruction level at Akrotiri in the late seventeenth century BC (note also that time intervals between the 14C dated episodes or events in a sequence can be quantified, thus the ‘Boundary earlier Late Minoan II’ quantifies the period between the Late Minoan IB final destruction at Mochlos and the close of Late Minoan II destruction at Knossos – i.e. it dates most of the actual Late Minoan II Phase – despite our having no direct 14C dates for this period). Figure 8.11c gives an example of potential queries in such models: quantifying the period of time between the Akrotiri volcanic destruction level (VDL, the eruption) and the Late Minoan IB late destruction at Myrtos-Pyrgos on Crete. The Figure 8.11 analysis suggests the need to consider a new (14C-based) timeframe for the earlier part of the Aegean Late Bronze Age (for more details, see Manning 2014), whereas both 14C and the conventional pre-14C chronology suggest fairly similar date ranges for the Aegean and east Mediterranean periods 1400–1200 BC when there are many material culture linkages (e.g. Manning et al. 2001; 2009; 2013). Subsequently, however, after 1200 BC (end Late Bronze Age to start Iron Age), Radiocarbon dating and archaeology: history, progress and present status 149

i&Cahd.Z.S-BronK Ram^Y.i2Q121.i:5JatQallS.alina5Bliedg-Quiy.eJEeimeLe.Lal-2QlSj___ (Set uence Ll\ IA-II (a(a)) ( (b)b) 1629-1614 BC (68.2%) (PI ase LMTA “ Boundary Thera 1639-1605 BC (95.4%) Pnase Mile os LMIA bone samples Eruption to LMIA Fi ?_Dafe OcA-11954 — ?_DateO:A-11951_____ — - l"ria i^ ------[Sequence/ ikrotiri anil Thera VI5E Boundary Chania L MIB Phase Pre-VDL LMIA Destruction or TAQ R_Comb ne M4N0I ' 3 ------ft— ------R_Combi ne 65/NOl <1/12______— ------Phase late LMIAanc VDL R_Combine Myrtos Pyrgo: R_Comb ne Triandu short-livt d late LM A twig ■ —------LMIB Late Destruct on Sequeno i Akrotiri yDL (After Mil 5tos Thera eruption FPQ [ =Fetliri( Miletos I VM______-- ^ — ______Boundary Mochlos Des truction — Tau_Boi ndary Start VDL Phase | or TAQ (Phase T iera VDL laTer LMI/ Boundaiy Thera Eruption to LMIA Fina I______- - ^ 1 ------Be undary Post-VDL LMIA Final o earlierLMIB Boundary earlier LA 111 fPf ase LMIB Late-Final Destructons Sequence Chania (LMIB to Destruction) I - rau_Boun iary Start Chania Pnase _ ’hase Ch; inia data R_Combine Knossos LMII— R_Comb ne Chania 1 ------— .^g.— Destruction R_Comb ne Chania 2 ------—— ^ R_Comb ne Chania 3 -g R_Comb ne Chania 4______— — —: n ^ 1900 1800 1700 1600 1500 1400 1300 boundary Chania Lt\ 1IB Destri ction or TAQ______- (5 equence l MIB Late to LMIB F inal Modelled date (BC) R_Combir e Myrtos- Dyrgos LA 'IB Late D e s t r u c t i o n ------Sequence Mochlos i LMIB Fin; 1) Interval from Boundary Thera Eruption to Tau_Boundary Start Mochlos Phase ^ Myrtos-Pyrgos LMIB Late Destruction ((c)c) Phase Mochlos 68.2% probability R_Date 35991 - —* (68.2%) 104-156 R_Date 35992 — 95.4% probability (95.4%) 91-171 R_Date 115890 — ® 0.02. R_Date 129765 ------— < £• R_Date 151768 - - - ~ S 0.01 ■ Boundary Mochlos Destructie n or TAQ______— : 6c undary ea rlier LMII I »■ R_ Combine Knossos i .Mil Desti notion - 0 50 100 150 200 2200 2100 2000 1900 1800 1700 1600 1500 1400 1300 Interval (yrs) Modelled date (BC)

FIGURE 8.11 (a) Calibrated age distributions from a re-run and revised version of the Manning et al. (2006) sequence analysis of the Aegean early Late Bronze Age (see text). Calibration curve resolution set at 5 years. (b) Some details of the modelled ranges. (c) Example of query: length of time (years) from the Akrotiri volcanic destruction level Boundary = eruption date to the Late Minoan IB Late Destruction of Myrtos-Pyrgos.

when there is again less archaeological clarity, there is again some debate over possible discrepancies between conventional dates and detailed radiocarbon- based chronology (e.g. Wardle et al. 2014; cf. Toffolo et al. 2013). Why was the conventional chronology incorrect as regards the early Late Bronze Age Aegean? If one critically examines the archaeological linkages and stylistic associations between the Aegean and Egypt, it in fact turns out that there are few and relatively vague data and arguments as regards the time around the Thera eruption (Middle Minoan (MM) III and Late Minoan (LM) IA) (Höflmayer 2012a; Manning et al. 2014). In contrast, there are multiple connec- tions sometime earlier in the nineteenth to eighteenth centuries BC, a period for which a number of Minoan ceramic finds (MM IB–II) are known from Egypt and the Levant. There are also a number of connections for the late LM IB, II and III periods (and their Late Helladic Greek mainland contemporaries), from fifteenth-century to twelfth-century contexts in Egypt and the Levant (see the 150 Sturt W. Manning

clear and still valid statement of Cadogan 1978; for details, see Höflmayer 2012b). For these periods before and after the Thera timeframe, the conven- tional chronology appears to be more or less accurate and agrees with 14C evidence where available. It is just the Thera period, Middle Minoan III, Late Minoan IA and earlier Late Minoan IB that need a new chronology – to replace what was in fact just a poorly based best guess (Manning 2014; Manning et al. 2014). Furthermore, recent work on Egyptian chronology – both using extensive sophisticated 14C dating and modelling (Bronk Ramsey et al. 2010) and critically reviewing the historical documentation from Egypt (Aston 2012–2013) – indicates that it is likely that the dates for the early New Kingdom (Eighteenth Dynasty) rulers of Egypt may be towards the higher end of the historical estimates, contrary to the past generation of work, which has tended to select lower dates. This makes it much easier to see how the archaeological linkages can be synchronized in line with the Aegean 14C chronology (Manning 2014; Manning et al. 2014). For many years it was held that the historically estimated dates for the archaeological strata at the massive site of Tell el-Dab’a in the Nile Delta (capital of the Hyksos rulers of Egypt during the Second Intermediate Period and a port city with connections around the region including with the Aegean) disproved the Aegean ‘high’ 14C-based chronology. However, indicative of changing times, this obstacle is now entirely overturned. An extensive 14C dating programme at Tell el-Dab’a has shown major problems in the dating of the site, and possibly errors (to older dates) of – no coincidence – about a century (Kutschera et al. 2012). At the same time, excavations at Tell Edfu in Upper Egypt have found strong evidence that the sequence of Hyksos kings and their dates need radical revision and that some kings – notably the prominent Hyksos king Khayan, whose name was on a lid imported to Knossos on Crete likely in the late Middle Minoan period – may be (again likely no coincidence) up to a century older than thought (Moeller and Marouard 2011). This news, too, requires a radical rethink of the Tell el-Dab’a sequence and dates (Manning et al. 2014).

Discussion From its invention onwards, there have been a number of (proclaimed) radiocarbon revolutions – undoubtedly too many for easy digestion (Bronk Ramsey 2008: 249), and inevitably, ‘as in most revolutions . . . the aftermath has proved to be rather confused’ (Whittle 1988: 12) in many instances. Nonetheless, looking at the bigger picture and overall history, the ongoing development of 14C dating technology, practice and application in archaeology has revolutionized the subject. Radiocarbon-based research has progressively brought the possibility and potential of greater accuracy and precision in archaeological dating – replacing the previous unquantifiable conceptual and chronological models based on ‘typologically based Radiocarbon dating and archaeology: history, progress and present status 151 cross-dating with its web of a priori assumptions’ (Whittle 1988: 12) – and has extended our confident dating reach further back in time. Thanks to recent work, there is now a well-defined terrestrial 14C calibration record for the past 50,000 years. Each revolution has of course brought some challenges for previous archaeological chronologies and derived interpretative frameworks, and there have been various settling-in periods in scholarship over the past seven decades. Today 14C has become the main basis for the majority of chronologies for prehistoric to protohistoric archaeology around the world. Most recently, Bayesian chronological modelling provides an analytical framework and tools to fully integrate archaeologi- cal information with 14C dating in order to address archaeologically relevant (and therefore socially relevant) timescales and episodes. It offers exciting scope both to re-address old problems and to investigate new questions in the analysis and writing of human history.

Note 1 OxCal is available from http://c14.arch.ox.ac.uk/oxcal.html (last accessed February 2014).

References Alberti, G. (2013) ‘Issues in the absolute chronology of the Early–Middle Bronze Age transition in Sicily and southern Italy: a Bayesian radiocarbon view’, Journal of Quaternary Science, 28: 630–636. Anthony, D. W. (1990) ‘Migration in archaeology: the baby and the bathwater’, American Anthropologist, 92: 895–914. Arnold, J. R. and Libby, W. F. (1949) ‘Age determinations by radiocarbon content: checks with samples of known age’, Science, 110: 678–680. Arnold, J. R. and Libby, W. F. (1951) ‘Radiocarbon dates’, Science, 113: 111–120. Aston, D. (2012–2013) ‘Radiocarbon, wine jars and new kingdom chronology’, Ägypten und Levant, 22–23: 289–315. Bayliss, A. (2009) ‘Rolling out revolution: using radiocarbon dating in archaeology’, Radiocarbon, 51: 123–147. Bayliss, A., Bronk Ramsey, C., van der Plicht, J. and Whittle, A. (2007) ‘Bradshaw and Bayes: towards a timetable for the Neolithic’, Cambridge Journal of Archaeology, 17: 1–28. Betancourt, P. P. (1987) ‘Dating the Aegean Late Bronze Age with radiocarbon’, Archaeometry, 29: 45–49. Betancourt, P. P. and Weinstein, G. A. (1976) ‘Carbon-14 and the beginning of the Late Bronze Age in the Aegean’, American Journal of Archaeology, 80: 329–348. Bietak, M. (2003) ‘Science versus archaeology: problems and consequences of high Aegean chronology’, in M. Bietak (ed.) The Synchronisation of Civilisations in the Eastern Mediterranean in the Second Millennium BC II. Proceedings of the SCIEM 2000 – EuroConference Haindorf, 2nd of May – 7th of May 2001, 23–33. Contributions to the Chronology of the Eastern Mediterranean II. Wien: Verlag der Österreichischen Akademie der Wissenschaften. Blockley, S. P. E., Lane, C. S., Hardiman, M., Rasmussen, S. O., Seierstad, I. K., Steffensen, J. P., Svensson, A., Lotter, A. F., Turney, C. S. M., Bronk Ramsey, C. and INTIMATE members (2012) ‘Synchronisation of palaeoenvironmental records over the last 60,000 152 Sturt W. Manning

years, and an extended INTIMATE event stratigraphy to 48,000 b2k’, Quaternary Science Reviews, 36: 2–10. Bond, G., Kromer, B., Beer, J., Muscheler, R., Evans, M. N., Showers, W., Hoffmann, S., Lotti-Bond, R., Hajdas, I. and Bonani, G. (2001) ‘Persistent solar influence on North Atlantic climate during the Holocene’, Science, 294: 2,130–2,136. Bronk Ramsey, C. (1995) ‘Radiocarbon calibration and analysis of stratigraphy: the OxCal program’, Radiocarbon, 37(2): 425–430. Bronk Ramsey, C. (2008) ‘Radiocarbon dating: revolutions in understanding’, Archaeometry, 50(2): 249–275. Bronk Ramsey, C. (2009a) ‘Bayesian analysis of radiocarbon dates’, Radiocarbon, 51(2): 337–360. Bronk Ramsey, C. (2009b) ‘Dealing with outliers and offsets in radiocarbon dating’, Radiocarbon, 51(3): 1,023–1,045. Bronk Ramsey, C., Buck, C. E., Manning, S. W., Reimer, P. and van der Plicht, H. (2006) ‘Developments in radiocarbon calibration for archaeology’, Antiquity, 80: 783–798. Bronk Ramsey, C., Dee, M. W., Rowland, J. M., Higham, T. F. G., Harris, S. A., Brock, F. A., Quiles, A., Wild, E., Marcus, E. S. and Shortland, A. J.( 2010) ‘Radiocarbon-based chronology for Dynastic Egypt’, Science, 328: 1,554–1,557. Bronk Ramsey, C., Higham, T., and Leach, P. (2004) ‘Towards high-precision AMS: progress and limitations’, Radiocarbon, 46(1): 17–24. Bronk Ramsey, C., Staff, R. A., Bryant, C. L., Brock, F., Kitagawa, H., van der Plicht, J., Schlolaut, G., Marshall, M. H., Brauer, H. F., Payne, R. L., Tarasova, P. E., Haraguchi, T., Gotanda, K., Yonenobu, H., Yokoyama, Y., Tada, R. and Nakagawa, T. (2012) ‘A complete terrestrial radiocarbon record for 11.2 to 52.8 kyr B.P.,” Science, 338: 370–374. Broodbank, C. (2013) The Making of the Middle Sea: A History of the Mediterranean from the Beginning to the Emergence of the Classical World, London: Thames and Hudson. Bruins, H. J., Nijboer, A. J. and van der Plicht, J. (2011) ‘Iron Age Mediterranean chronology: a reply’, Radiocarbon, 53(1): 199–220. Buck, C. E., Kenworthy, J. B., Litton, C. D. and Smith, A. F. M. (1991) ‘Combining archaeological and radiocarbon information: a Bayesian approach to calibration’, Antiquity, 65: 808–821. Buck, C. E., and Millard, A. (2004) Tools for Constructing Chronologies: Crossing Disciplinary Boundaries, London: Springer. Cadogan, G. (1978) ‘Dating the Aegean Bronze Age without radiocarbon’, Archaeometry, 20: 209–214. Childe, V. G. (1950) ‘Review of Chronologie der jüngeren Steinzeit Mittel- und Südosteuropas by Vladimir Milojcˇic´’, Gnomon, 22: 290–292. Clark, R. (1975) ‘A calibration curve for radiocarbon dates’, Antiquity, 49: 251–266. Clark, R. M. and Renfrew, C. (1973) ‘Tree-ring calibration of radiocarbon dates and the chronology of ancient Egypt’, Nature, 243: 266–270. Dalfes, H. N., Kukla, G. and Weiss, H. (eds) (1997) Third Millennium BC Climate Change and Old World Collapse, NATO ASI Series I, Vol. 49, Berlin: Springer. Damon, P. E., Ferguson, C. W., Long, A. and Wallick, E. I. (1974) ‘Dendrochronologic calibration of the radiocarbon timescale’, American Antiquity, 39: 350–366. Damon, P. E., Lerman, J. C. and Long, A. (1978) ‘Temporal functions of 14C: causal factors and implications’, Annual Review of Earth and Planetary Science, 6: 457–494. Daniel, G. (1959) ‘Editorial’, Antiquity, 33: 237–240. de Jong, A. F. M., Mook, W. G. and Becker, B. (1979) ‘Confirmation of the Suess Wiggles: 3200–700 BC,’ Nature, 280: 48–49. Radiocarbon dating and archaeology: history, progress and present status 153 de Vries, H. (1958) ‘Variation in the concentration of radiocarbon with time and location on Earth’, Koninklijke Nederlandse Akademie van Wetenschappen (Series B), 61: 94–102. Dean, J. S. (1978) ‘Independent dating in archaeological analysis’, in M. B. Schiffer (ed.) Advances in Archaeological Method and Theory 1, New York: Academic Press, 223–255. Dee, M., Wengrow, D., Shortland, A., Stevenson, A., Brock, F., Girdland Flink, L. and Bronk Ramsey, C. (2013) ‘An absolute chronology for early Egypt using radiocarbon dating and Bayesian statistical modeling’, Proceedings of the Royal Society A, 469: 20130395. Dellinger, F., Kutschera, W., Nicolussi, K., Schießling, P., Steier, P and Wild, E. M. (2004) ‘A 14C calibration with AMS from 3500 to 3000 BC, derived from a new high-elevation Stone-pine tree-ring chronology’, Radiocarbon, 46(2): 969–978. Doumas, C. (1983) Thera: Pompeii of the Ancient Aegean, London: Thames and Hudson. Fantalkin, A., Finkelstein, I. and Piasetzky, E. (2011) ‘Iron Age Mediterranean chronology: a rejoinder’, Radiocarbon, 53(1): 179–198. Ferguson, C. W., Gimbutas, M. and Suess, H. E. (1976) ‘Historical dates for Neolithic sites of Southeast Europe’, Science, 191: 1,170–1,172. Ferguson, C. W., Huber, B. and Suess, H. E. (1966) ‘Determination of the age of Swiss lake dwellings as an example of dendrochronologically-calibrated radiocarbon dating’, Zeitschrift für Naturforschung, 21A: 1,173–1,177. Finkelstein, I. (2011) ‘The Iron Age chronology debate: is the gap narrowing?’, Near Eastern Archaeology, 74(1): 50–54. Finkelstein, I. and Piasetzky, E. (2010) ‘Radiocarbon dating the Iron Age in the Levant: a Bayesian model for six ceramic phases and six transitions’, Antiquity, 84: 374–385. Friedrich, W. L. (2009) Santorini: Volcano, Natural History, Mythology, trans. A. R. McBirney, Aarhus: Aarhus University Press. Galimberti M., Bronk Ramsey, C. and Manning, S. W. (2004) ‘Wiggle-match dating of tree ring sequences’, Radiocarbon, 46(2): 917–924. Gillespie, R. (2002) ‘Dating the first Australians’, Radiocarbon, 44(2): 455–472. Gove, H. E. (1992) ‘The history of AMS, its advantages over decay counting: applications and prospects’, in R. E. Taylor, A. Long and R. S. Kra (eds) Radiocarbon after Four Decades: An Interdisciplinary Perspective, New York: Springer-Verlag, 214–229. Hassan, F. A (1986) ‘The restoration of time: radiocarbon dating 1965–1985/6’, Quarterly Review of Archaeology, 7(2): 5–6. Helms, M. W. (1988) Ulysses’ Sail: An Ethnographic Odyssey of Power, Knowledge, and Geographical Distance, Princeton, NJ: Princeton University Press. Higham, C. and Higham, T. (2009) ‘A new chronological framework for prehistoric Southeast Asia, based on a Bayesian model from Ban Non Wat’, Antiquity, 83: 125–144. Hodder, I. (2012) Entangled: An Archaeology of the Relationships between Humans and Things, Malden: Wiley-Blackwell. Höflmayer, F. (2012a) ‘The date of the Minoan Santorini eruption: quantifying the “Offset”’, Radiocarbon, 54(3–4): 435–448. Höflmayer, F. (2012b) Die Synchronisierung der minoischen Alt- und Neupalastzeit mit der ägyptischen Chronologie, Wien: Verlag der Österreichischen Akademie der Wissenschaften. Hogg, A. G., Hua, Q., Blackwell, P. G., Niu, M., Buck, C. E., Guilderson, T. P., Heaton, T. J., Palmer, J. G., Reimer, P. J., Reimer, R. W., Turbey, C. S. M. and Zimmerman, S. R. H. (2013) ‘SHCal13 Southern hemisphere calibration, 0–50,000 years cal BP’, Radiocarbon, 55(4): 1,889–1,903. Kohler, E. L. and Ralph, E. K. (1961) ‘C-14 dates for sites in the Mediterranean area’, American Journal of Archaeology, 65: 357–367. Kromer, B. (2009) ‘Radiocarbon and dendrochronology’, Dendrochronologia, 27: 15–19. 154 Sturt W. Manning

Kutschera, W., Bietak, M., Wild, E. M., Bronk Ramsey, C., Dee, M., Golser, R., Kopetzky, K., Stadler, P., Steier, P., Thanheiser, U and Weninger, F. (2012) ‘The chronology of Tell el-Daba: a crucial meeting point of 14C dating, archaeology, and Egyptology in the 2nd millennium BC,” Radiocarbon, 54(3–4): 407–422. Lee, S., Bronk Ramsey, C. and Mazar, A. (2013) ‘Iron Age chronology in Israel: results from modeling with a trapezoidal Bayesian framework’, Radiocarbon, 55(3–4): 731–740. Libby, W. F. (1951) ‘Radiocarbon dates II’, Science, 114: 291–296. Libby, W. F. (1961) ‘Radiocarbon dating’, Science, 133: 621–629. Libby, W. F. (1963) ‘Accuracy of radiocarbon dates’, Science, 140: 278–280. Libby, W. F., Anderson, E. C. and Arnold, J. R. (1949) ‘Age determination by radiocarbon content: world-wide assay of natural radiocarbon’, Science, 109: 227–228. Lorentzen, B., Manning, S. W., Cvikel, D. and Kahanov, Y. (2014) ‘High-precision dating the Akko 1 shipwreck, Israel: wiggle-matching the life and death of a ship into the historical record’, Journal of Archaeological Science, 41: 772–783. Manning, S. W. (1988) ‘The Bronze Age eruption of Thera: absolute dating, Aegean chronology and Mediterranean cultural interrelations’, Journal of Mediterranean Archaeology, 1(1): 17–82. Manning, S. W. (1995) The Absolute Chronology of the Aegean Early Bronze Age: Archaeology, History and Radiocarbon, Sheffield: Sheffield Academic Press. Manning, S. W. (2013a) ‘Cyprus at 2200 BC: rethinking the chronology of the Cypriot Early Bronze Age’, in A. B. Knapp, J. M. Webb and A. McCarthy (eds) J. R. B. Stewart – An Archaeological Legacy, SIMA CXXXIX, Uppsala: Åströms Förlag, 1–21. Manning, S. W. (2013b) ‘The Roman world and climate: context, relevance of climate change, and some issues’, in W. V. Harris (ed.) The Ancient Mediterranean Environment between Sciences and History. Leiden: Brill, 103–170. Manning, S. W. (2014) A Test of Time and A Test of Time Revisited: The Volcano of Thera and the Chronology and History of the Aegean and East Mediterranean in the Mid-Second Millennium BC, Oxford: Oxbow Books. Manning, S. W. and Kromer, B. (2011) ‘Radiocarbon dating archaeological samples in the eastern Mediterranean 1730–1480 BC: further exploring the atmospheric radiocarbon calibration record and the archaeological implications’, Archaeometry, 53(2): 413–439. Manning, S. W. and Swiny, S. (1994) ‘Sotira Kaminoudhia and the chronology of the Early Bronze Age in Cyprus’, Oxford Journal of Archaeology, 13: 149–172. Manning, S. W. and Weninger, B. (1992) ‘A light in the dark: archaeological wiggle matching and the absolute chronology of the close of the Aegean Late Bronze Age’, Antiquity, 66: 636–663. Manning, S. W., Bronk Ramsey, C., Kutschera, W., Higham, T., Kromer, B., Steier, P. and Wild, E. (2006) ‘Chronology for the Aegean Late Bronze Age’, Science, 312: 565–569. Manning, S. W., Höflmayer, F., Moeller, N., Dee, M. W., Bronk Ramsey, C., Fleitmann, D., Higham, T., Kutschera, W. and Wild, E. M. (2014) ‘Dating the Thera (Santorini) eruption: coherent archaeological and scientific evidence supporting a high chronology’, Antiquity 88 (in press). Manning, S. W., Kromer, B., Dee, M. W., Friedrich, M. Higham, T. F. G. and Bronk Ramsey, C. (2013) ‘Radiocarbon calibration in the mid to later 14th century BC and radiocarbon dating Tell el-Amarna, Egypt’, in A. J. Shortland and C. Bronk Ramsey (eds) Radiocarbon and the Chronologies of Ancient Egypt. Oxford: Oxbow Books, 121–145. Manning, S. W., Pulak, C., Kromer, B., Talamo, S., Bronk Ramsey, C. and Dee, M. (2009) ‘Absolute age of the Uluburun shipwreck: a key Late Bronze Age time-capsule for the east Mediterranean’, in S. W. Manning and M. J. Bruce (eds) Tree-Rings, Kings and Old Radiocarbon dating and archaeology: history, progress and present status 155

World Archaeology and Environment: Papers Presented in Honor of Peter Ian Kuniholm. Oxford: Oxbow Books, 163–187. Manning, S. W., Weninger, B., South, A. K., Kling, B., Kuniholm, P. I., Muhly, J. D., Hadjisavvas, S., Sewell, D. A. and Cadogan, G. (2001) ‘Absolute age range of the Late Cypriot IIC period on Cyprus’, Antiquity, 75: 328–340. Mazar, A. (2011) ‘The Iron Age chronology debate: is the gap narrowing? Another viewpoint’, Near Eastern Archaeology, 74(2): 105–111. Mazar, A. and Bronk Ramsey, C. (2008) ‘14C dates and the Iron Age chronology of Israel: a response’, Radiocarbon, 50(2): 159–180. McCormick, M., Büntgen, U., Cane, M. A., Cook, E. R., Harper, K., Huybers, P., Litt, T., Manning, S. W., Mayewski, P. A., More, A. F. M., Nicolussi, K. and Tegel, W. (2012) ‘Climate change during and after the Roman Empire: reconstructing the past from scientific and historical evidence’, Journal of Interdisciplinary History, 43(2): 169–220. McCullagh J., Marom, A. and Hedges, R. (2010) ‘Radiocarbon dating of individual amino acids from archaeological bone collagen’, Radiocarbon, 52(2): 620–634. Mellaart, J. (1979) ‘Egyptian and Near Eastern chronology: a dilemma?’, Antiquity, 53: 6–18. Mellars, P. (2006a) ‘A new radiocarbon revolution and the dispersal of modern humans in Eurasia’, Nature, 439: 931–935. Mellars, P. (2006b) ‘Archaeology: progress and pitfalls in radiocarbon dating – reply’, Nature, 443: E4. Milojcˇic´, V. (1949) Chronologie der jüngeren Steinzeit Mittel- und Südosteuropas, Berlin: Gebr. Mann Verlag. Milojcˇic´, V. (1957) ‘Zur Anwendbarkeit der C14-Datierung in der Vorgeschichtsforschung’, Germania, 35: 102–110. Moeller, N. and Marouard, G. (with a contribution by Ayers, N.) (2011) ‘Discussion of late Middle Kingdom and early Second Intermediate Period history and chronology in relation to the Khayan Sealings from Tell Edfu’, Ägypten und Levante, 21: 87–121. Nalawade-Chavan, S., McCullagh, J., Hedges, R., Bonsall, C., Boroneant, A., Bronk Ramsey, C. and Higham, T. (2012) ‘Compound specific radiocarbon dating of essential and non-essential amino acids: towards determination of dietary reservoir effects in humans’, Radiocarbon, 55(2–3): 709–719. Ogburn, D. E. (2012) ‘Reconceiving the chronology of Inca imperial expansion’, Radiocarbon, 54(2): 219–237. Panagiotakopulu, E. and Buckland, P. C. (1991) ‘Insect pests of stored products from Late Bronze Age Santorini, Greece’, Journal of Stored Products Research, 27: 179–184. Pearson, G. W. and Stuiver, M. (1986) ‘High-precision calibration of the radiocarbon time scale, 500–2500 BC’, Radiocarbon, 28(2B): 839–862. Pearson, G. W., Pilcher, J. R. and Baillie, M. G. L. (1983) ‘High precision 14C measurements of Irish oaks to show the natural 14C variations from 200 BC to 4000 BC’, Radiocarbon, 25: 179–186. Pearson, G. W., Pilcher, J. R., Baillie, M. G. L., Corbet, D. M. and Qua, F. (1986) ‘High precision C-14 measurements of Irish oaks to show the natural C-14 variations from AD 1840 – 5210 BC,’ Radiocarbon, 28(2B): 911–934. Pearson, G. W., Pilcher, J. R., Baillie, M. G. L. and Hillam, J. (1977) ‘Absolute radiocarbon dating using a low altitude European treering calibration’, Nature, 270: 25–28. Piggott, S. (1954) Neolithic Cultures of the British Isles: A Study of the Stone-Using Agricultural Communities of Britain in the Second Millennium BC, Cambridge: Cambridge University Press. Piggott, S. (1959) ‘The radio-carbon date from Durrington Walls’, Antiquity, 33: 289–290. Raczky, P. and Siklósi, Z. (2013) ‘Reconsideration of the Copper Age chronology of the eastern Carpathian Basin: a Bayesian approach’, Antiquity, 87: 555–573. 156 Sturt W. Manning

Reimer, P. J., Baillie, M. G. L., Bard, E., Bayliss, A., Beck, J. W., Bertrand, C. J. H., Blackwell, P. G., Buck, C. E., Burr, G. S., Cutler, K. B., Damon, P. E., Edwards, R. L., Fairbanks, R. G., Friedrich, M., Guilderson, T. P., Hogg, A. G., Hughen, K. A., Kromer, B., McCormac, G., Manning, S., Ramsey, C. B., Reimer, R. W., Remmele, S., Southon, J. R., Stuiver, M., Talamo, S., Taylor, F. W., van der Plicht, J. and Weyhenmeyer, C. E. (2004) ‘INTCAL04 terrestrial radiocarbon age calibration, 0–26 cal kyr BP’, Radiocarbon, 46(3): 1,029–1,058. Reimer, P.J., Baillie, M. G. L., Bard, E., Bayliss, A., Beck, J. W., Blackwell, P. G., Bronk Ramsey, C., Buck, C. E., Burr, G. S., Edwards, R. L., Friedrich, M., Grootes, P. M., Guilderson, T. P., Hajdas, I., Heaton, T. J., Hogg, A. G., Hughen, K. A., Kaiser, K. F., Kromer, B., McCormac, F. G., Manning, S. W., Reimer, R. W., Richards, D. A., Southon, J. R., Talamo, S., Turney, C. S. M., van der Plicht, J. and Weyhenmeyer, C. E. (2009) ‘IntCal09 and marine09 radiocarbon age calibration curves, 0–50,000 years cal BP’, Radiocarbon, 51(4): 1,111–1,150. Reimer, P. J., Bard, E., Bayliss, A., Beck, J. W., Blackwell, P. G., Bronk Ramsey, C., Buck, C. E., Cheng, H., Edwards, R. L., Friedrich, M., Grootes, P. M., Guilderson, T. P., Haflidason, H., Hajdas, I., Hatté, C., Heaton, T. J., Hoffmann, D. L., Hogg, A. G., Hughen, K. A., Kaiser, K. F., Kromer, B., Manning, S. W., Niu, M., Reimer, R. W., Richards, D. A., Scott, E. M., Southon, J. R., Staff, R. A., Turney, C. S. M. and van der Plicht, J. (2013) ‘IntCal13 and marine13 radiocarbon age calibration curves 0–50,000 years cal BP’, Radiocarbon, 55(4): 1,869–1,887. Renfrew, C. (1969) ‘The autonomy of the south-east European Copper Age’, Proceedings of the Prehistoric Society, 35: 12–47. Renfrew, C. (1970) ‘The tree-ring calibration of radiocarbon: an archaeological evaluation’, Proceedings of the Prehistoric Society, 36: 280–311. Renfrew, C. (1971) ‘Carbon 14 and the prehistory of Europe’, Scientific American, 225: 63–72. Renfrew, C. (1972) The Emergence of Civilization: The Cyclades and the Aegean in the Third Millennium BC, London: Methuen. Renfrew, C. (1973) Before Civilisation: The Radiocarbon Revolution and Prehistoric Europe, London: Jonathan Cape. Renfrew, C. (1980) ‘The great tradition versus the great divide: archaeology as anthropology?’, American Journal of Archaeology, 84: 287–298. Renfrew, C., Boyd, M. and Bronk Ramsey, C. (2012) ‘The oldest maritime sanctuary? Dating the sanctuary at Keros and the Cycladic Early Bronze Age’, Antiquity, 86: 144–160. Rutter, J. B. (1983) ‘Some observations on the Cyclades in the later third and early second millennia’, American Journal of Archaeology, 87: 69–76. Rutter, J. B. (2011) ‘Late Minoan IB at Kommos: a sequence of at least three distinct stages’, in T. M. Brogan and E. Hallager (eds) LM IB Pottery: Relative Chronology and Regional Differences: Acts of a Workshop Held at the Danish Institute at Athens in Collaboration with the INSTAP Study Center for East Crete, 27–29 June 2007, Monographs of the Danish Institute at Athens Volume 11, Athens: The Danish Institute at Athens, 307–343. S¸ahogˇlu, V. (2005) ‘The Anatolian trade network and the Izmir region during the Early Bronze Age’, Oxford Journal of Archaeology, 24: 339–361. S¸ahogˇlu, V. (2011) ‘Trade and interconnections between Anatolia and the Cyclades during the 3rd millennium BC’, in V. S¸ahogˇlu and P. Sotirakopoulou. (eds) Across: The Cyclades and Western Anatolia During the 3rd Millennium BC. Istanbul: Sakip Sabanci Muzesi, 172–177. Salzer, M. W., Bunn, A. G., Graham, N. E. and Hughes, M. K. (2014) ‘Five millennia of paleotemperature from tree-rings in the Great Basin, USA’, Climate Dynamics, 42: 1517–1526. DOI: 10.1007/s00382-013-1911-9. Radiocarbon dating and archaeology: history, progress and present status 157

Sarpaki, A. (2001) ‘Processed cereals and pulses from the Late Bronze Age site of Akrotiri, Thera: preparations prior to consumption, a preliminary approach to their study’, Annual of the British School at Athens, 96: 27–41. Schiffer, M. B. (1986) ‘Radiocarbon dating and the “old wood” problem: the case of the Hohokam chronology’, Journal of Archaeological Science, 13: 13–30. Soles, J. S. (2004) ‘Radiocarbon results’, in K. A. Barnard and T. M. Brogan (eds) Mochlos IC. Period III. Neopalatial Settlement on the Coast: The Artisan’s Quarter and the Farmhouse at Chalinomouri. The Small Finds, Philadelphia, PA: INSTAP Academic Press, 145–149. Snodgrass, A. M. (1985) ‘The new archaeology and the classical archaeologist’, American Journal of Archaeology, 89: 31–37. Staubwasser, M. and Weiss, H. (2006) ‘Holocene climate and cultural evolution in late prehistoric–early historic West Asia’, Quaternary Research, 66: 372–387. Stuiver, M. (1961) ‘Variations in radiocarbon concentration and sunspot activity’, Journal of Geophysical Research, 66: 273–276. Stuiver, M. (1982) ‘A high-precision calibration of the AD radiocarbon time scale’, Radiocarbon, 24: 1–26. 14 Stuiver, M. and Braziunas, T. F. (1993) ‘Sun, ocean, climate and atmospheric CO2: an evaluation of causal and spectral relationships’, The Holocene, 3: 289–305. Stuiver, M. and Quay, P. D. (1980) ‘Changes in atmospheric carbon-14 attributed to a variable sun’, Science, 207: 11–19. Stuiver, M., Braziunas, T. F., Becker, B. and Kromer, B. (1991) ‘Climatic, solar, oceanic, and geomagnetic influences on late-glacial and Holocene atmospheric 14C/12C change’, Quaternary Research, 35: 1–24. Stuiver, M., Reimer, P. J., Bard, E., Beck, J. W., Burr, G. S., Hughen, K. A., Kromer, B., McCormac, G., van der Plicht, J. and Spurk, M. (1998) ‘IntCal98 radiocarbon age calibration, 24,000–0 cal BP’, Radiocarbon, 40(3): 1,041–1,083. Suess, H. E. (1965) ‘Secular variations of cosmic ray produced carbon-14 in the atmosphere’, Journal of Geophysical Research, 70: 5,937–5,952. Suess, H. E. (1970) ‘Bristlecone pine calibration of the radiocarbon time-scale 5200 BC to the present’, in I. U. Olsson (ed.) Nobel Symposium 12, Radiocarbon Variations and Absolute Chronology, 303–311. New York: John Wiley & Sons. Taylor, R. E. (1987) Radiocarbon Dating: An Archaeological Perspective, Orlando: Academic Press. Taylor, R. E. (1997) ‘Radiocarbon dating’, in R. E. Taylor and M. J. Aitken (eds) Chronometric Dating in Archaeology. New York: Plenum Press, 65–96. Taylor, R. E. (2000) ‘Fifty years of radiocarbon dating’, American Scientist, 88: 60–67. Taylor, R. E., Stuiver, M. and Reimer, P. J. (1996) ‘Development and extension of the calibration of the radiocarbon time scale: archaeological applications’, Quaternary Science Reviews, 15: 655–668. Thomas, H. L. and Ehrich, R. W. (1969) ‘Some problems in chronology’, World Archaeology, 1: 143–156. Toffolo, M. B., Fantalkin, A., Lemos, I. S., Felsch, R. C. S., Niemeier, W.-D., Sanders, G. D. R., Finkelstein, I. and Boaretto, E. (2013) ‘Towards an absolute chronology for the Aegean Iron Age: new radiocarbon dates from Lefkandi, Kalapodi and Corinth’, PLoS ONE 8(12): e83117; doi: 10.1371/journal.pone.0083117 Turney, C. S. M., Roberts, R. G. and Jacobs, Z. (2006) ‘Archaeology: progress and pitfalls in radiocarbon dating’, Nature, 443(7,108): E3. van Andel, T. H. (2005) ‘The ownership of time: approved C-14 calibration or freedom of choice?’, Antiquity, 79: 944–948. van der Plicht, J., Beck, J. W., Bard, E., Baillie, M. G. L., Blackwell, P. G., Buck, C. E., Friedrich, M., Guilderson, T. P., Hughen, K. A., Kromer, B., McCormac, F. G., Bronk 158 Sturt W. Manning

Ramsey, C., Reimer, P. J., Reimer, R. W., Remmele, S., Richards, D. A., Southon, J. R., Stuiver, M. and Weyhenmeyer, C. E. (2004) ‘NotCal04 – comparison/calibration C-14 records 26–50 cal kyr BP’, Radiocarbon, 46(3): 1,225–1,238. van der Plicht, J., Bruins, H. J. and Nijboer, A. J. (2009) ‘The Iron Age around the Mediterranean: a high chronology perspective from the Groningen radiocarbon database’, Radiocarbon, 51(1): 213–242. Vieira, L. E. A., Solanki, S. K., Krovova, N. A. and Usoskin, I. (2011) ‘Evolution of the solar irradiance during the Holocene’, Astronomy & Astrophysics, 531(A6): doi: 10.1051/0004- 6361/201015843. Wardle, K., Higham, T. and Kromer, B. (2014) ‘Dating the end of the Greek Bronze Age: a robust radiocarbon-based chronology from Assiros Toumba’, PLoS ONE 9(9): e106672. doi: 10.1371/journal.pone.0106672. Warren, P. and Hankey, V. (1989) Aegean Bronze Age Chronology, Bristol: Bristol Classical Press. Waterbolk, H. T. (1971) ‘Working with radiocarbon dates’, Proceedings of the Prehistoric Society, 37: 15–33. Weiss, H. (2014) ‘Altered trajectories: the Intermediate Bronze Age’, in M. Steiner and A. E. Killebrew (eds) The Oxford Handbook of the Archaeology of the Levant c.8000–332 BCE, 367–387. Oxford: Oxford University Press. Weiss, H., Courty, M.-A., Wetterstrom, W., Guichard, F., Senior, L., Meadow, R. and Curnow, A. (1993) ‘The genesis and collapse of third millennium North Mesopotamian civilization’, Science, 261: 995–1,004. Weiss, H., Manning, S. W., Ristvet, L., Mori, L., Besonen, M., McCarthy, A., Quenet, P., Smith, A. and Bahrani, Z. (2012) ‘Tell Leilan Akkadian Imperialization, collapse and short-lived reoccupation defined by high-resolution radiocarbon dating’, in H. Weiss (ed.) Seven Generations Since the Fall of Akkad, Studia Chaburensia 3. Weisbaden: Harrassowitz Verlag, 163–192. Whittle, A. (1988) Problems in Neolithic Archaeology, Cambridge: Cambridge University Press. Whittle, A., Healy, F. and Bayliss, A. (2011) Gathering Time: Dating the Early Neolithic Enclosures of Southern Britain and Ireland, Oxford: Oxbow Books. Wiener, M. H. (2013) ‘“Minding the gap”: gaps, destructions, and migrations in the Early Bronze Age Aegean: causes and consequences’, American Journal of Archaeology, 117: 581–592. Wild, E. M., Gauß, W., Forstenpointner, G., Lindblom, M., Smetana, R., Steier, P., Thanheiser, U. and Weninger, F. (2010) ‘14C dating of the Early to Late Bronze Age stratigraphic sequence of Aegina Kolonna, Greece’, Nuclear Instruments and Methods in Physics Research B, 268(7–8): 1,013–1,021. Willis, E. H., Tauber, H., and Munnich, K. O. (1960) ‘Variations in the atmospheric radiocarbon concentration over the past 1,300 years’, Radiocarbon, 2: 1–4. 9 USING EVIDENCE FROM NATURAL SCIENCES IN ARCHAEOLOGY

David Killick

Archaeology has always had a close working relationship with the natural sciences. The origins of archaeology were closely entwined with the emergence of geology and palaeontology from the mid-nineteenth century (Trigger 1988), and the chemical analysis of archaeological objects goes back further still, to the late eighteenth century (Pollard and Bray, Chapter 7, this volume). Today, almost all archaeologists make some use of evidence generated by scientific methods. Kristiansen (2011) has suggested that archaeology has swung between periods in which new questions are stimulated mostly by clusters of new methods, and periods when reactions against method-driven archaeology lead some prominent archaeologists to favor the interpretive approaches of the humanities. He goes on to argue that we are now in a period in which new scientific methods are increas- ingly driving research, displacing the prior dominance of the humanist post- processual archaeology from the mid-1980s to 2000. While I agree with him on this last point, I am less convinced by his argument that archaeology over the past two hundred years has swung, pendulum-like, between visions of archaeology as applied science and of archaeology as a branch of the humanities. In my view, this swing really only applies to a subset of British and Scandinavian archaeologists, and to an even smaller subset of American archaeologists. Even in these regions humanistic archaeologies – classical archaeology, biblical archaeology, etc. – have always coexisted with science-assisted empiricism. Humanism and science-assisted archaeology have always coexisted in German, Eastern European, Indian, Chinese, Japanese, Near Eastern, and Egyptian archaeologies. In regions with few or no written records, such as sub-Saharan Africa, Australia, and South America, archaeology has always been empirical and method-driven. Kristiansen’s statement that we are currently seeing a surge in applications of science to archaeology is supported by examination of archaeological publishing patterns, by trends in the hiring of academic faculty and by trends in research awards 160 David Killick in Europe and North America. I have followed these trends over the last twenty years in the USA and in Britain, and have paid particular attention to the institutional and national research cultures that have encouraged or retarded the adoption of archaeological science (Killick and Young 1997; Killick 2008; Killick and Goldberg 2009). What I conclude from this is that British universities have moved much more swiftly than their American counterparts to take advantage of new scientific techniques and scientific expertise. I think that there are two reasons for this difference. The first is that the combined pool of British and European Community funding for archaeological science is much larger than in the USA or Canada. The second reason is that British archaeology departments are freestanding units, whereas almost all archaeologists in the USA are embedded within departments of anthropology, classics, or Near Eastern languages. British archaeology departments can, and increasingly do, hire faculty for their particular skills. It is increasingly common to find British archae- ology departments hiring archaeological scientists whose doctoral degrees are in other disciplines, such as geology or molecular biology. But, it is very unusual indeed for North American departments of anthropology, classics, or Near Eastern studies to hire anyone who has a doctoral degree in the natural sciences. Thus North American archaeologists, for the most part, have had to look elsewhere for scientists with whom to collaborate, but the supply of willing and suitably experi- enced archaeological scientists in departments of natural science and engineering is not nearly equal to the demand for their services. Some forms of expertise that are widely available to archaeologists in Europe are quite difficult to obtain in North America. These include luminescence dating, heavy isotope analysis, organic residue analysis, ancient DNA analysis, geophysical remote sensing, optical petrog- raphy and expertise in the reconstruction of ancient pyrotechnologies (Killick 2008). North American archaeologists are however in a much better position in this respect than are archaeologists in India, Africa, and South America, who often find it very difficult to obtain access to even the most essential forms of archaeo- logical sciences, such as radiocarbon dating or chemical analysis. But in Europe, at least, archaeology is about fifteen years into a major surge in applications of the natural sciences to archaeology. This has brought many benefits to archaeology, particularly in allowing archaeologists to ask new questions – for example, was this deceased person born in this area? (Evidence for region of birth can be found in strontium and oxygen isotopic ratios in teeth.) New methods have also revived old questions that had been set aside for lack of conclusive evidence – for example, where were cattle first domesticated? (Genomic studies of modern cattle show unequivocally that there were three separate domestications of cattle in West Eurasia, India, and North Africa.) The recent flood of publications on the “cognitive revolution” between 100,000 and 60,000 years ago owe their existence to the recent development of reliable methods of dating deposits in this age range. But in the rush to put these new scientific methods to work on the archaeological record, there has been little attention paid to the question of how best to integrate scientific evidence into archaeology. What is the role of scientific evidence in Using evidence from natural sciences in archaeology 161 archaeology today? Given that most archaeologists have little training in basic sciences, how can misunderstandings be avoided? What are the most productive and trustworthy forms of cooperation between archaeologists and archaeological scientists? And how should scientific evidence be combined with other archaeo- logical evidence to make convincing interpretations?

The role of scientific evidence in archaeology Telling stories about the past is an essential part of the craft of archaeology. Without narratives archaeological reports would simply be catalogues of finds and their classifications. Narratives convey our more or less expert interpretations of our findings to other archaeologists, to descendant communities, and to the wider audience whose taxes support most archaeological research. They animate the past and stimulate discussion among experts and interest in archaeology among the wider public, whose taxes support our work. Our narratives about the lives of people long deceased usually employ the techniques of the humanities, not those of the sciences, and are most similar to the narratives of historians. (The major exception is in evolutionary archaeology, whose narratives are adapted from those of neo-Darwinian evolutionary biology (Kuhn 2004; Shennan 2008).) Where we differ sharply among ourselves is in how tightly we feel constrained by the actual archaeological evidence in telling our stories. Narratives influenced by phenomenology (e.g. Tilley 2004; van Dyke 2008) are by definition subjective and are relatively loosely constrained by archaeological evidence. At the other end of the spectrum, evolutionary archaeologists attempt to strip away any trace of subjectivity. Their models are also stories, but are tightly constrained by evolutionary theory, and are explicitly constructed for testing against data. If these models fail to explain patterning in the data, so much the better: their failures allow the building of more adequate models (Kuhn 2004). This approach has definite benefits, but remains a minority position within archaeology because most archaeologists prefer to talk and write about people rather than about evolu- tion. Evolutionary archaeologists often accuse other archaeologists of lacking rigor, but evolutionary archaeologists stand accused in turn of being excessively narrow. Most archaeologists operate somewhere in between these end points. They accept that the stories they tell should be consistent with archaeological and other evidence, but they have more relaxed standards for explanation than evolutionary archaeologists do, and thus address a much wider range of issues. Most archaeolo- gists are pragmatists, and thus evaluate the adequacy of interpretations/explanations by examining how well they account for the archaeological evidence, without resorting to formal models. But archaeological evidence is woefully incomplete, and this widely acknowledged deficiency opened the door for relativists like Shanks and Tilley (1987) to charge that the stories archaeologists tell are no more than politically charged fables imposed from the present onto the past. Their position has been widely rejected as too extreme, but it did serve the useful function of making archaeologists – and their allies within philosophy – think 162 David Killick about how to build narratives that are more tightly constrained by evidence. The philosopher who has thought most deeply about explanation in archaeology, Alison Wylie, has argued that building a robust archaeological interpretation requires the weaving together (“cabling”) of truly independent lines of evidence (Wylie 1989, 2000). This is where archaeological science is of crucial importance, for the physical phenomena that it exploits are truly independent of cultural influences, past or present, or of location. The half-life of radiocarbon, for example, is the same whether one is studying hunter-gatherer archaeology in the Kalahari or the archaeology of complex societies in Cambodia, and applies equally to processualists, postprocessualists, and evolutionary archaeologists. Cabling together independent lines of evidence can reduce the number of possible alternative archaeological narratives, and ideally one continues to add strands to the cable until only one narrative endures. This process of elimination is somewhat reminiscent of Karl Popper’s argument that scientists should try to refute hypotheses rather than confirm them (e.g. Popper 1963). Though Popper’s philosophy of science is now of only historical interest, it is certainly true that many of the most compelling applications of science to archaeology are those that have refuted long-standing interpretations of the past. Examples of these are given below, and Mark Pollard has long argued that scientific techniques for provenance of inorganic materials should be employed to exclude potential sources rather than to claim that a given artifact derived from a particular source (Wilson and Pollard 2001; Pollard and Bray, Chapter 7, this volume). His point is that one can rarely be sure that the database of chemical or isotopic “fingerprints” of potential sources is complete. An artifact may have a perfect match to the “fingerprint” of a particular source, but how can we be sure that it might not also match the “fingerprint” of a source yet undiscovered or uncharacterized?

Some examples One way in which scientific evidence constrains interpretations of the past is by disrupting long-established narratives. My first two examples are of this kind.

Ceramic petrography and the tyranny of the ethnographic record In the mid-1930s, Alfred V. Kidder, one of the leading archaeologists in the southwestern USA, encouraged one of his young research associates, Anna Shepard, to apply optical petrography to prehistoric pottery in the Rio Grande river valley and adjacent regions. Her studies (collected in Shepard 1942) showed that during the Pueblo IV period (1350–1600 CE) the production of decorated pottery had been concentrated at just a few pueblos, with other pueblos importing most of their decorated wares. Her findings were rejected by many archaeologists working in that region because (1) they could not evaluate the petrographic evidence, and (2) there was no specialization in pottery among historic and contemporary pueblos Using evidence from natural sciences in archaeology 163 along the Rio Grande. The latter point must be emphasized. In no other region of the world – with the possible exception of the Southern African Kalahari (Wilmsen 1989) – has archaeological interpretation been so thoroughly subordinated to the ethnographic record as in the Southwestern USA (see Lekson 2009). Because there is direct historical continuity from prehistoric pueblos to historic pueblos, interpretations of the archaeology of the prehistoric pueblos have been largely conducted by analogy to historic pueblos. Since historic pueblos did not specialize in ceramic production, most archaeologists in the 1940s refused to believe that prehistoric pueblos did so. But Shepard’s findings have stood the test of time, and ceramic petrography continues to expand the evidence for specialization in ceramic production during the Pueblo IV period (e.g. Cordell and Habicht-Mauche 2012). Shepard also undertook petrographic analysis of pottery from the excavations directed by Neil Judd in Pueblo Bonito, Chaco Canyon, which was inhabited from c.830 to c.1130 CE. She found that almost all the pottery that she examined had been made at least 80 kilometers away. Even though the Great Houses of Chaco Canyon obviously differed in important respects from historic pueblos, Judd was so wedded to the idea of self-sufficiency that he refused to accept her conclusion. He published her report in his monograph for the site, but added a dissent, which Shepard followed with a rebuttal (Shepard 1954). But her conclusion still stands, and it paved the way for other applications of science to Chacoan materials. Some fifty years later, strontium isotope ratio analysis of a sample of the estimated 200,000 roof beams in Pueblo Bonito and other great houses in Chaco Canyon – with an estimated average mass of 275 kilograms each – shows that all the specimens tested grew in mountain ranges 75–100 kilometers distant from Chaco Canyon (English et al. 2001). Studies of strontium, oxygen, and carbon isotopic ratios in a sample of deer bones from Chaco Canyon also show that most of them were imported from at least this far away (Grimstead 2011). Analysis of organic residues from an unusual type of pottery, most known examples of which come from a single room in Pueblo Bonito, has shown that these residues contain biomarkers for cocoa, the nearest possible source of which is in central Mexico, about 2400 kilometers south of Chaco Canyon (Crown and Hurst 2009). Osteological analyses have also produced evidence of markedly skewed sex ratios in burials and of violence against some women in the Chaco region. These data have been interpreted as evidence of captive women obtained by raiding (Cameron 2013: 224–225). Each of these contributions from archaeological science provides an independent strand in an emerging narrative about the Great Houses. In the first decade of this century, the dominant interpretation of Chaco was that it was a peaceful center of pilgrimage for religious rituals, with a very small resident population at other times of the year (Mills 2002). It now appears more likely that the fully developed great houses were inhabited by a powerful elite who dominated a large area and had at least indirect contact with Mesoamerica. Southwestern archaeologists, strongly influenced by the egalitarian social structure of historic pueblos, have gone to extraordinary lengths to avoid calling Chaco a state – but it increasingly looks like one (Lekson 2009). 164 David Killick

Maize and mounds: the stable carbon isotope revolution

The period dating from c.500 cal BCE to c.400 cal CE in the central and eastern USA is called the Middle Woodland, and is divided into two partly contemporary complexes called Adena and Hopewell. The most striking features of Adena and Hopewell archaeological sites are large earthen mounds. Some of these cover burials, which often contain large quantities of burial goods, including stone tools, figurative art skillfully cut from sheets of copper and mica, and elaborately carved objects in pipestone and other soft rocks. Other mound groups consist of geometric platforms, plazas, and processional ways demarcated by earth walls, and a few are large-scale landscape sculptures, like the Serpent Mound in Ohio. Until the 1980s all textbooks of American archaeology stated that later Adena and all Hopewell groups were maize agriculturalists (e.g. Willey and Sabloff 1974). Maize cobs, albeit rather small ones, had been recovered from some Adena and Hopewell sites, and it was an article of faith in the cultural evolutionary mindset – then dominant in American anthropological archaeology – that foraging societies do not invest collective labor in building monuments. In the mid-1970s two scholars, working in far-off South Africa, shook this comfortable consensus to its core. Joseph Vogel was the director of the radiocarbon laboratory in Pretoria, and Nikolaas van der Merwe was professor of archaeology at the University of Cape Town. Van der Merwe was one of the first individuals to be trained in both archaeology and science. His BSc degree was in physics, and in 1966 he was awarded his PhD in anthropology from Yale University for his research on direct radiocarbon dating of iron-carbon alloys. Because he was intimately familiar with the radiocarbon literature, he knew that by 1970 radiocarbon scientists had identified an anomaly while dating some archaeological sites in North America. This was that radiocarbon dates on carbonized corn cobs were consistently younger than dates on wood charcoal from the same contexts. This anomaly arises because trees – and all other plants in eastern North America – have a different photosynthetic chemistry from maize and many other tropical grasses. The former type is called C3 photosynthesis because the product is a three- carbon molecule; the latter is called C4 photosynthesis because the product is a four-carbon molecule. (In both cases, the products are polymerized to form 12- carbon sugars.) The carbon comes from carbon dioxide in air, but both of these photosynthetic pathways slightly discriminate against (fractionate) the heavier carbon isotopes (13C and 14C), so that the three-carbon and four-carbon molecules contain more of the lightest isotope (12C) than does atmospheric carbon dioxide. Radiocarbon scientists used this knowledge to correct radiocarbon dates on maize for fractionation of 14C, but van der Merwe realized that fractionation also offered a way of tracking the adoption of maize (a C4 plant) in eastern North America, where all plants except maize utilized C3 photosynthesis. He and Vogel were able to obtain samples of human bone from Woodland and later contexts, and measured the 13C/12C ratios in the bone protein (collagen) extracted from them. Their initial results implied that Middle Woodland populations essentially Using evidence from natural sciences in archaeology 165 ate no maize, and that maize did not become a major part of human diet in this region until the Early Mississippian period (800–1000 CE). Their conclusion proved hugely controversial, and negative reviews by archaeologists delayed publication by three years (van der Merwe, personal communication). But, the data themselves were unimpeachable, and eventually overcame resistance from archaeologists (Vogel and van der Merwe 1977; van der Merwe and Vogel 1978). Since then stable carbon isotopes have become a tool of fundamental importance in studies of prehistoric diet and of the spread of agri- culture (Pollard et al. 2007: 180–188). Why were so many archaeologists initially so opposed to the conclusions drawn from this carbon isotope study? It was because it appeared to contradict one of the most basic generalizations derived from the archaeological record: that foraging societies did not cooperate to build large monuments. (A generalization since toppled from its pedestal by the discovery of pre-agricultural monuments at Göbekli Tepe in Anatolia (Dietrich et al. 2012)). But there was another possible interpretation, which is that Adena and Hopewell societies were agriculturalists, but not maize agriculturalists. This thought sent a group of paleobotanists scurrying back to the archaeological record. There they discovered evidence of Middle Woodland agriculture based on domesticated plants, including marsh elder, goosefoot, lamb’s quarters, maygrass, knotweed, squash, and sunflower. All but squash and sunflower were abandoned in North American agriculture when maize was adopted (Smith 1989). Stable carbon and nitrogen isotope analyses have since become standard tools in the reconstruction of past diet (Pollard et al. 2007). Their measurement is now largely automated and cheap, and there are probably close to one hundred thousand such measurements in the archaeological literature. Along the way, however, their status as evidence has changed. All of the easy isotopic challenges to archaeological orthodoxies about the adoption of crops have already been made, and thus light stable isotopes have made the transition from a starring to a supporting role in reconstructions of ancient diet. Isotopic analysis of collagen from humans provides just two numbers: one the ratio of stable carbon isotopes; the other a ratio of stable nitrogen isotopes. What these numbers provide is evidence about the average diet over a period of five to seven years. Except in the easiest cases, like the adoption of maize agriculture in North America, they do not prove that any particular food was consumed. The major challenge to reconstructing diet at any particular archaeological site is that the types of evidence represent different intervals of time. Food residues in coprolites are evidence of a single meal; organic residues in pots of one or more meals; macrobotanical and faunal remains represent ranges of time from a season to multiple years; and stable isotopes provide the dietary average over five to seven years. Mismatches of scale can be found in many other archaeological reconstructions. In reconstructing migrations, for example, strontium and oxygen isotopes in teeth provide evidence for the region of residence of an individual in middle childhood, and stylistic evidence (from pottery, metalwork, etc.) provides evidence for the migration of individuals or groups to a particular site; while 166 David Killick genomic studies of modern populations may be unable to distinguish between single or multiple migrations from a particular region of origin (as, for example, the several Mongol incursions from the Eastern Steppes into Western Asia).

Accuracy and precision in scientific evidence I have worked at the interface between archaeology and the laboratory sciences for thirty-five years, and one of the conclusions that I have drawn from my experience is that many archaeologists do not really understand the concepts of accuracy and precision in scientific measurement. Laboratory scientists need to be fanatical about accuracy and precision. A laboratory that publishes inaccurate data will eventually be discredited – as the British Museum radiocarbon laboratory was some two decades ago – and techniques that offer better precision of measurement constantly displace techniques that are less precise. For example, accelerator mass spectrometry (AMS) has largely displaced beta counting in radiocarbon dating over the last twenty years, and mass spectrometry has replaced alpha counting in uranium-series dating. The most accurate and most precise scientific dating technique in archaeology is dendrochronology. It was developed in Southwestern USA in the 1920s and 1930s (e.g. Haury 1935) and the techniques employed have changed little since then. Master dendrochronologies have since been constructed for dozens of locations across the Colorado Plateau and Mogollon Rim. The single-year pre- cision that dendrochronology provides allows archaeologists to reconstruct chronology with a degree of detail that is the envy of archaeologists who work in other parts of the world. The growth of large structures with hundreds of rooms can be reconstructed by dating each room (e.g. Dean 1978), and the growth of settlements in one region can be correlated exactly with abandonments elsewhere. Variation in the width of tree rings in this region is controlled by rainfall. Archaeologists have matched historical crop yields under rain-fed agriculture to the corresponding tree-ring widths to obtain mathematical functions that can then be used to model prehistoric agricultural yields, and thus to infer whether abandon- ment of regions was a response to drought (e.g. Axtell et al. 2002). The chrono- logical span of pottery types can be tightly defined through their occurrence in contexts dated by tree rings, which allows archaeological sites in the Southern Desert (where trees are scarce) to be cross-dated by pottery imported from the Colorado Plateau. Fired-clay hearths in rooms dated by tree rings provide the samples used to build a master archaeomagnetic curve for the Southwest, the geographic range of which extends far beyond that in which dendrochronology can be employed. Dendrochronology sets the gold standard for accuracy and precision in archaeological science, as it is absolutely reproducible between observers and laboratories. None of the radiometric dating techniques used in archaeology is nearly so precise (though all may be as accurate). Unlike dendrochronology, radiometric dating is also subject to random and systematic inter-laboratory errors. Using evidence from natural sciences in archaeology 167

These were common in radiocarbon dating before the 1980s but systematic errors have been greatly reduced since then by periodic calibration exercises, in which samples of known radiocarbon content are sent out repeatedly for “blind” dating to many laboratories, and the cumulative results subjected to statistical analysis. But because of the stochastic nature of radioactive decay it is statistically quite unlikely that two consecutive dates run on the same sample by any radiometric technique will return exactly the same result. The precision of radiometric dates is vitally important in archaeological practice because it limits what kind of questions one can ask about the past. This is a difficulty that many archaeological theorists routinely overlook. For example, Dobres (2000) argues that similarity in the styles of bone needles in Magdalenian sites from the French Pyrenees is evidence of a shared habitus between women in these sites. Unfortunately for her argument, the precision of radiocarbon dating does not even permit the presumption that any two of the levels from which the needles came date to the same century.

Building on the work of others: the importance of accuracy in archaeological science Precision and accuracy are not just important in dating. For a rather different example, let us consider the role of chemical analysis in inferring the provenance of pottery. In most parts of the world earthenware pottery is the most abundant class of artifact preserved in archaeological sites younger than 10,000 years, and archaeolo- gists devote a significant amount of their time to studying it. Description and classification of shape and decoration occupies most of this effort, and from these data archaeologists develop ceramic chronologies, infer the methods of manu- facture, and attempt to distinguish pots that were made elsewhere from those made locally. In some regions the cumulative volume of stylistic evidence is huge, truly rising to the level of “big data.” A recent network analysis of changing connections between sites in Southwestern USA between 1200 CE and 1450 CE was based on some 800,000 pieces of decorated pottery (Mills et al. 2013). But often the evidence from style and shape is suggestive but inconclusive; exotic styles can be copied by local potters. Several scientific methods have therefore been applied to the problem of distinguishing local from imported pottery. The two main methods used in attempts to answer the question “Where was this pot made?” are optical petrography and chemical analysis. The most widely used approach to provenance of pottery is “bulk” chemical analysis (actually analysis of 100–200 milligram samples of each sherd) for about forty elements by neutron activation analysis (NAA), X-ray fluorescence (XRF) or inductively coupled plasma mass spectrometry (ICP-MS). Graphical techniques and multivariate statistics are then used to search for structure in the data. But if statistically valid chemical groups are found in the data, what meaning should be assigned to them? Since potsherds are physical mixtures of clay, rock, and mineral 168 David Killick fragments, and sometimes other materials (e.g. ground potsherds), it is not usually possible to map chemical compositions on to regional geology. Comparison of the chemistry of raw clay samples and potsherds is sometimes attempted but is rarely satisfactory because (1) comprehensive sampling of clay resources is rarely possible, and (2) any addition of temper other than pure quartz makes it very difficult to match sherds to clays. The approach usually employed is to try to create chemical reference groups for pottery types that are designated a priori as “local” for each site. If ceramic kilns are present, as in the Ancient Near East, these reference groups are usually created on failed pots (“wasters”) from kilns. But where open firing was the norm, as in most of the New World, more tenuous a priori decisions must be made about which potsherds should be designated as “local.” In practice, the most abundant type of pottery on site is often designated as “local.” But as we have already seen – in Anna Shepard’s work in New Mexico – this is not always a safe assumption. This assump- tion can be checked by petrographic analysis. Stylistic, chemical, and petrographic evidence provide three independent strands for weaving a narrative about connec- tions between places, and three strands make a stronger cable than one or two. Chemical analysis is expensive, and it would be wasteful indeed if the chemical evidence could not be combined with that from other provenance studies of pottery in the same region. Fortunately the major NAA, XRF, and ICP-MS laboratories in the USA that do chemical analysis of pottery and obsidian use the same external chemical standards, and also have regularly exchanged samples. The data produced by each laboratory are comparable in terms of precision and accuracy, and the results obtained can be combined with samples run by the other laboratories to build databases that can be used with confidence in subsequent studies. Unfortunately this model of how to undertake provenance analysis in archae- ology is now threatened by the spread of relatively cheap portable XRF (p-XRF) spectrometers. We are now seeing a flood of “do-it-yourself ” provenance studies by archaeologists using p-XRF. These spectrometers measure only a small subset of the elements measurable by the specialist laboratories mentioned above, and the accuracy and precision of the spectrometers differ by manufacturer and model. Many of these units were designed for identifying metal alloys in warehouses and recycling plants, and the onboard software is tuned accordingly. When used on non-metals – stone artifacts, pottery, soils, etc. – they will not give accurate results. Some archaeologists are aware of this (e.g. Frahm 2012), but argue that it is insignificant. If the p-XRF results are internally consistent and can distinguish between sources for the purposes of a given project, then (they argue) the results are “good enough” for archaeology. The “good enough” attitude towards scientific evidence is unfortunately quite common among archaeologists, but should be strongly opposed (Speakman and Shackley 2013). Provenance studies whose data are of unknown accuracy and precision cannot be combined with other studies unless the latter are carried out with the same instrument. Science is supposed to be a cumulative enterprise, in Using evidence from natural sciences in archaeology 169 which the evidence from individual studies can be combined with other studies of comparable quality to create larger data sets. The research value of large electronic databases of good data has been appreciated for many years in science, engineering, and medicine. Enthusiasm for “big data” has spread more recently to the financial sector and to social media corporations, and is starting to migrate into archaeology (Axtell et al. 2002; Knappett 2011; Bray and Pollard 2012; Mills et al. 2013). But the “good enough” attitude towards scientific evidence in archaeology is absolutely incompatible with the cumulative growth of knowledge.

Discussion Scientific evidence is particularly valuable in the construction of durable archae- ological narratives because it is often independent of other sorts of archaeological evidence, and because the underlying principles of physics, chemistry, and biology have not changed through time. This allows us to combine archaeological science with experimental archaeology to generate and characterize products that we can compare with those from the archaeological record. This dual strategy has perhaps been most extensively employed in reconstructing extinct metallurgical tech- nologies (e.g. Dungworth and Doonan 2013). The last fifteen years have seen a huge surge in applications of science to archae- ology. This surge has been strongly concentrated in Europe, but is now increasingly apparent in China. It has not been nearly as marked in the USA and Canada, where there is much less funding than in Europe for archaeology in general and archaeological science in particular. The growth of science-assisted archaeology in North America has also been slowed by an acute shortage of specialists in archaeological science, which is a consequence of the reluctance of anthropology and classics departments to hire faculty with degrees in other fields. (Archaeologists in many anthropology departments are also outnumbered by cultural and social anthropologists, many of whom are not fond of science in any form.) European archaeology departments have much more flexibility, and have been hiring archaeological scientists in droves, in part (I suspect) because of their potential to bring in large amounts of external funding. Archaeological scientists are much more familiar with archaeological theory than they were twenty years ago (Martinón-Torres and Killick, in press), but many archaeologists still refuse to learn even the basic principles of the scientific tech- niques upon which they are increasingly dependent (for explicit refusals by prominent archaeologists, see Jones 2002: 33; and Shott 2010). Since many training programs are still run by senior archaeologists who think that one can never have enough theory, many doctoral students still graduate without having taken even a basic survey course in archaeological science. This is the case even at the University of Arizona, which has been a world leader in archaeological science for more than seventy years! This might not matter if there were an adequate supply of experi- enced archaeological scientists with whom these archaeologists could collaborate, but only Western Europe has been training enough of these. 170 David Killick

The explosive growth of new scientific methods for archaeology over the last twenty years is changing archaeological practice. The old model of site excavation resembled a wagon wheel, with the principal archaeologist (site director) at the hub, a ring of specialists around the rim, and the spokes of the wheel representing the separate two-way communications between the site director and each specialist. Today the specialists often need to communicate directly with each other to do their jobs – for example, the archaeometallurgist needs the geologist to identify ore samples, the botanist to identify charcoals, the lead isotope specialist to distinguish local from exotic metals. Synthesis and interpretation today tends to be made by teams rather than by the site director. Trends in publication reflect this tendency; scientific studies now tend to be published separately rather than included as dry appendices in the site monograph, as used to be the case. Perhaps the greatest impact of the new scientific methods is yet to come, in that they can potentially provide compelling justifications for the extremely expensive habit of storing and conserving indefinitely the finds from past excavations. Recent identifications of bits of Neanderthal genome in modern humans were made by genetic analysis of bones excavated, in some cases, a century or more ago. Chapter 7 in this volume, by Mark Pollard and Peter Bray, demonstrates a new approach to provenance of copper alloy objects that makes use of archival chemical data, which can be improved by new and more precise reanalysis of objects in stored collections. Radiocarbon dating is now so sensitive – requiring less than a milligram of carbon – that it is possible to return to old stored collections and to date materials without causing significant damage to them. Redating old collections significantly enhances their value to present scholarship. And it is perhaps not too impudent to suggest that the future of prehistoric archaeology in North America lies largely in working with stored collections, given that new excavations are increasingly limited to salvage of sites threatened with destruction. New scientific techniques allow archaeologists to bring significant new questions to old collections.

References Axtell, R. L., Epstein, J. M., Dean, J. S., Gummerman, G. J., Swedlund, A. C., Harburger, J., Chakravarty, S., Hammond, R., Parker, J. and Parker, M. (2002) “Population growth and collapse in a multiagent model of Kayenta Anasazi in Long House,” Proceedings of the National Academy of Sciences, 99: 7275–7279. Bray, P. J. and Pollard, A. M. (2012) “A new interpretative approach to the chemistry of copper-alloy objects: source, recycling and technology,” Antiquity, 86: 853–867. Cameron, C. M. (2013) “How people moved among ancient societies: broadening the view,” American Anthropologist, 115: 218–231. Cordell, L. S. and Habicht-Mauche, J. A. (eds) (2012) Potters and Communities of Practice: Glaze Paint and Polychrome Pottery in the American Southwest, AD 1250–1700, Anthropological Papers of the University of Arizona 75. Tucson: University of Arizona Press. Crown, P. L. and Hurst, J. W. (2009) “Evidence of cacao use in the Prehispanic American Southwest,” Proceedings of the National Academy of Sciences, 106: 2110–2113. Using evidence from natural sciences in archaeology 171

Dean, J. S. (1978) “Independent dating in archaeological analysis,” Advances in Archaeological Method and Theory, 1: 223–255. Dietrich, O., Heun, M., Notroff, J., Schmidt, K. and Zarnkow, M. (2012) “The role of cult and feasting in the emergence of Neolithic communities. New evidence from Göbekli Tepe, south-eastern Turkey,” Antiquity, 86: 674–695. Dobres, M. A. (2000) Technology and Social Agency: Outlining a Practice Theory for Archaeology, Malden: Blackwell Publishers. Dungworth, D. and Doonan, R. (eds) (2013) Accidental and Experimental Archaeometallurgy, London: Historical Metallurgy Society. English, N., Betancourt, J. L., Dean, J. S. and Quade, J. (2001) “Strontium isotopes reveal distant sources of architectural timber in Chaco Canyon, New Mexico,” Proceedings of the National Academy of Sciences, 98: 11,981–11,986. Frahm, E. (2012) “Validity of ‘off-the-shelf’ handheld portable XRF for sourcing Near Eastern obsidian chip debris,” Journal of Archaeological Science, 40: 1080–1092. Grimstead, D. (2011) “Applications of evolutionary theory and stable isotope geochemistry shed light on North American prehistoric human behavior and regional procurement systems,” PhD dissertation, University of Arizona. Haury, E. (1935) “Tree rings: the archaeologists’s time-piece,” American Antiquity, 1(2): 98–108. Hawkes, J. (1968) “The proper study of mankind,” Antiquity, 42: 255–261. Hodder, I. (ed.) (1982) Structural and Symbolic Archaeology, Cambridge: Cambridge University Press. Hodder, I. (1995) Interpreting Archaeology: Finding Meaning in the Past, New York: Routledge. Jones, A. (2002) Archaeological Theory and Scientific Practice, Cambridge: Cambridge University Press. Killick, D. J. (2004) “Review essay. What do we know about African iron working?,” Journal of African Archaeology, 2(1): 97–112. Killick, D. J. (2008) “Archaeological science in the USA and in Britain,” in A. Sullivan (ed.) Archaeological Concepts for the Study of the Cultural Past, Salt Lake City, UT: University of Utah Press, 40–64. Killick, D. J. and Goldberg, P. (2009) “A quiet crisis in American archaeology,” SAA Archaeological Record, 29(1): 6–10. Killick, D. J. and Young, S. M. M. (1997) “Archaeology and archaeometry: from casual dating to a meaningful relationship?,” Antiquity, 71: 518–524. Knappett, C. (2011) An Archaeology of Interaction: Network Perspectives on Material Culture, New York: Oxford University Press. Kristiansen, K. (2011) “Theory does not die – it changes direction,” in M. Pearce and J. Bintliff (eds) The Death of Archaeological Theory?, Oxford: Oxbow Books, 72–79. Kuhn, S. L. (2004) “Evolutionary perspectives on technology and technological change,” World Archaeology, 36(4): 561–570. Lekson, S. H. (2009) A History of the Ancient Southwest, Santa Fe, NM: School of American Research. Martinón-Torres, M. and Killick, D. J. (in press) “Archaeological theories and archaeological sciences,” in A. Gardner, M. Lake and U. Sommer (eds) The Oxford Handbook of Archaeological Theory, Oxford: Oxford University Press. Mills, B. M. (2002) “Recent research on Chaco: changing views on economy, ritual and society,” Journal of Archaeological Research, 10: 45–117. Mills, B. M., Clark, J. J., Peeples, M. A., Haas, W. R., Roberts, J. M., Hill, J. B., Huntley, D. L., Borck, L., Breiger, R. L., Clauset, A. and Shackley, M. S. (2013) “Transformation 172 David Killick

of social networks in the late pre-Hispanic US Southwest,” Proceedings of the National Academy of Sciences, 110: 5785–5790. Pollard, A. M., Batt, C. M., Stern, B. and Young, S. M. M. (2007) Analytical Chemistry in Archaeology, Cambridge: Cambridge University Press. Popper, K. (1963) Conjectures and Refutations: The Growth of Scientific Knowledge, New York: Basic Books. Shanks, M. and Tilley, C. (1987) Social Theory and Archaeology, Cambridge: Polity Press in association with Blackwell Publishers. Shennan, S. (2008) “Evolution in archaeology,” Annual Review of Anthropology, 37: 75–91. Shepard, A. O. (1942) Rio Grande Glaze Paint Ware: A Study Illustrating the Place of Technological Analysis in Archaeological Research, Washington, DC: Carnegie Institution of Washington Publication 526, Contribution 39. Shepard, A. O. (1954) “Rebuttal to Judd,” in N. M. Judd (ed.) Material Culture of Pueblo Bonito, Washington, DC: Smithsonian Miscellaneous Collections 124, 236–238. Shott, M. J. (2010) “Crises and solutions in American archaeology,” SAA Archaeological Record, 10(1): 37–38. Smith, B. (1989) “Origins of agriculture in eastern North America,” Science, 246: 1566–1570. Speakman, R. J. and Shackley, M. S. (2013) “Silo science and portable XRF in archaeology: a response to Frahm,” Journal of Archaeological Science, 40: 1435–1443. Tilley, C. (2004) The Materiality of Stone, Oxford: Berg Publishers. Trigger, B. G. (1988) “Archaeology’s relations with the physical and biological sciences: a review,” Proceedings of the 26th International Archaeometry Symposium, Department of Physics, University of Toronto, 1–9. van der Merwe, N. J. and Vogel, J. C. (1978) “13C content of human collagen as a measure of prehistoric diet in Woodland North America,” Nature, 276: 815–816. Van Dyke, R. M. (2008) The Chaco Experience: Landscape and Ideology at the Center Place, Santa Fe, NM: SAR Press. Vogel, J. C. and van der Merwe, N. J. (1977) “Isotopic evidence for early maize cultivation in New York State,” American Antiquity, 42: 238–242. Willey, G. R, and Sabloff, J. A. (1974) A History of American Archaeology, San Francisco, CA: W. H. Freeman. Wilmsen, E. N. (1989) Land Filled with Flies: A Political History of the Kalahari, Chicago, IL: University of Chicago Press. Wilson, L. and Pollard, A. M. (2001) “The provenance hypothesis,” in D. R. Brothwell and A. M. Pollard (eds) Handbook of Archaeological Sciences, London: John Wiley, 507–517. Wylie, A. (1989) “Archaeological cables and tacking: the implications of practice for Bernstein’s Options beyond Objectivism and Relativism,” Philosophy of the Social Sciences, 19: 1–18. Wylie, A. (2000) “Questions of evidence, legitimacy, and the (dis)unity of science,” American Antiquity, 65(2): 227–237. 10 WORKING THE DIGITAL

Some thoughts from landscape archaeology

Marcos Llobera

Digital information is nowadays pervasive to all aspects of archeology. Despite its ubiquity, many archaeologists still regard studies that rely on the manipulation of digital representations with some (or much!) suspicion, questioning whether they have epistemic value and/or are appropriate within certain theoretical frameworks. In this chapter, I hope to show how the use of digital information can extend our understanding of the archaeological record. This may be accomplished in part through the creation of tools that derive new information from the existing record, and the use of (simple) simulations aimed at further clarifying and “contextualizing” archaeological patterns. I have previously referred to these types of model (if we may think of it as a category) as scaffolding models (see Llobera 2012) – that is, models that are meant to be exploratory by nature, and limited in scope. However, as the (ever-)increasing volume of digital information exposes archaeologists to new possibilities, forms of inquiry, and evidence, this also raises new challenges and difficulties. To illustrate both the possibilities and challenges, I shall make use of a landscape archaeology example. This example is based on an earlier study by Criado-Boado and Vázquez-Villoch (1998, 2000) that centered around a set of Neolithic barrows or mamoas and their relationship with patterns of movement on the Sierra de Barbanza in Galicia (northwestern Spain). My intention is to show the usefulness of digital data and tools in terms of a very simple simulation. Indeed, nothing I present here is meant to be highly original or sophisticated – quite the contrary. However, it is hoped that this will serve to demonstrate that such models are relevant and easily attainable when guided by clearly articulated archaeological goals. As we embrace the new research opportunities that emerge from the increased adoption of digital information (data and techniques), we need to temper these with careful consideration of their suitability for certain specific purposes. Is the information we are handling adequate for the type of investigations we are 174 Marcos Llobera interested in? Is its resolution or the way the information is stored appropriate? These are all basic questions we need to confront. Some of these issues have been discussed extensively, particularly with regards to GIS (geographic information systems), in fields such as geography (for different representations of space, see Miller and Wentz 2003; Peuquet 2002) but their implications do not always resonate amongst archaeologists. For the most part, we happily accept existing forms of representation without much thought or consideration. However, the growing prevalence of digital information makes it very easy to lose sight of the impact these might have on our analyses (see Lucas (2001: 148–170) for a similar point on the impact of the Harris Matrix). This realization becomes particularly poignant when, for instance, GIS techniques are used to investigate questions aimed at understanding the experiential properties of past landscapes. Given their cartographic origin, GIS are undoubtedly well suited to addressing questions at certain spatial (regional) scales for which the use of two-dimensional representations, as those provided by the spatial primitives (point, line, polygon), are appropriate. GIS are very successful at tasks such as: exploring the relation between archaeological site locations and ecological zones; obtaining contextual information, such as shortest distances from features like rivers, coasts, or quarries; and rendering the distribution of surface scatters on a landscape, to mention a few examples. However, as we move away from these scales of analysis and consider questions that require finer resolutions and more complicated (i.e. three- dimensional) forms of landscape representation, we soon confront the limitation inherent with these types of spatial representations, and with GIS in general. Hence, despite great advances in high-resolution terrain models, our ability to integrate these with more realistic representations of the built environment remains very limited. These limitations have a direct impact on our ability to address certain questions (e.g. using GIS to learn how the appearance of features changes through a landscape). As an example, the simulations presented here use a 5 x 5 meter-squared cell size digital elevation model (DEM) derived from LiDAR data. At the time, this model represented a sensible compromise between the size of the study area and the size of the Neolithic round barrows or mamoas (min.: 86 m2, max.: 694 m2, average: 302 m2) that I was interested in studying. However, this choice of data resolution (and data structure) did force some reflection on the nature and scale of the questions I was interested in exploring (as it should have!). For instance, to what extent does the choice of a certain vector feature to represent a barrow, and the use of a 5 x 5 m2 DEM model, constrain my ability to resolve questions regarding the appearance of these on the landscape? In the context of this study, a set of points evenly distributed throughout the entire surface of each monument was used to represent each mamoa. This stands opposed to the standard practice of representing an entire monument with a single point feature (see Figure 10.1a). The number of points oscillated between 86 and 694, and these changes had dramatic conse- quences. In addition to significantly extending the visible area associated with each monument, this measure allowed us to capture changes in visibility throughout the Working the digital: some thoughts from landscape archaeology 175

(a) (b)

0

50

100

FIGURE 10.1 (a) Area of visibility for one mamoa calculated from single point. (b) Area of visibility for the same mamoa after using a set of points to represent it (these do not appear). In the latter case, this area is effectively a cumulative viewshed that can be reclassified to indicate the percentage of mamoa visible from each location. This allows us to explore subtle changes in visibility depending on the direction of movement around the mamoa (see Llobera 2003: 39–44). landscape that would have been lost before (see Figure 10.1b). Admittedly, the use of several point features covering the extension of each barrow is not an ideal solution, but awareness of this limitation is a step in the right direction. When it comes to dealing with information systems, archaeological evidence cannot be divorced from its representation, and all representations imply a choice (conscious or not) about what constitutes meaningful information (in fact this is true for any archaeological evidence). Part of this choice requires a simplification of sorts, which is what ultimately enables us, in the case of information systems, to provide easy and quick manipulation. Representations allow new information to be generated within these systems. As archaeologists, we need to be cognizant of the nature of these representations and the impact they have in shaping our questions. Once information is represented in digital form it is potentially subject to different transformations and manipulations. One of these is their use within a simulation. Simulations are useful when trying to understand certain phenomena and processes, particularly in the presence of uncertainty or scarce information. However, they are not as widely used in archaeology as one might expect. There are many potential reasons for this – not least, lack of training within the discipline, and the association (for some archaeologists) of simulations with particular top- down approaches and/or normative views, among others. Recently, this situation has been changing and the use of simulations for exploratory purposes is picking up some momentum (e.g. Bevan and Wilson 2013; Bevan and Lake 2013; White and Surface-Evans 2012; Knappett 2013; Crema 2012; Eve and Crema n.d.). Within landscape archaeology studies, it is common to find references to simu- lations whenever GIS are applied. The term “simulation” is commonly (ab)used to 176 Marcos Llobera refer loosely to an ordered sequence of GIS procedures. When a sequence is executed in such a fashion, the result is often prescriptive and monolithic. That is, in spite of the complexity of some of the operations, the results are somewhat predictable (especially when the algorithms involved are known). In these cases, little additional knowledge is gained that was not available from the onset, and the outcome resembles a kind of (sophisticated) rendering rather than a probing into the data. Furthermore, the production of a single monolithic outcome provides a false sense of precision and removes our ability to contextualize the result, which would have enriched our interpretations (Gillings 1998). Within this chapter, I shall use the term in a more conventional way – i.e. to refer to a procedure that through repetition attempts to incorporate some aspect of variability.

Background The following examples focus on a series of mamoas found on the Serra da Barbanza, a wide mountain ridge located on the Barbanza peninsula in Galicia (see Map 10.1). Based on typological similarities with other Galician barrows, these mamoas are likely to have been constructed between roughly 4000 and 3500 BC, and would have remained in use until 2800 BC (Criado-Boado and Vázquez-Villoch 1998: 70), coinciding with the climatic transition between the Atlantic and Subboreal climatic periods. This group of barrows has been the subject of two earlier publications, an extensive study by Criado-Boado et al. (1986) and a later revision by Criado-Boado and Vázquez-Villoch (1998, 2000). In spite of these studies (which included some paleoenvironmental analyses and limited excavations), little is known about the archaeology of the area beyond the location of these barrows. According to Criado- Boado et al., the barrows were constructed using material that could be obtained

04Km

A L U S N I N E P SPAIN A Z N A B R A B

0150Km

MAP 10.1 Location of the Barbanza peninsula in Galicia (NW Spain). Working the digital: some thoughts from landscape archaeology 177 locally. No clear evidence has been found of nearby habitational structures, though Criado et al. have postulated that a relationship must have existed between these and the burials (1986). Their proposition rests on the observation that there is a short distance, and a visual connection, between the mamoas and the most productive agricultural fields in the area (some of which are currently occupied by cultivated fields and villages). Environmentally, the evidence retrieved from under some of the mamoas points towards an open landscape, very similar to the current one, with perhaps slightly more presence of oak (Quercus) during the Atlantic climatic period. While both studies have made reference to the relationship between the mamoas, visibility, and movement, it is in the latter (Criado-Boado and Vázquez-Villoch 1998, 2000) that this connection is central to the interpretation. According to the authors, the main purpose for the location of the mamoas would have been to guide people across the landscape and to (re-)produce particular scenes for travelers. The purpose of the movement and the social interpretation behind the pattern of scenes were never fully explored by the authors. Because of the relatively exposed nature of the Barbanza ridge, it is unlikely that the area would have sustained continuous occupation. Instead the ridge is likely to have operated mainly as a conduit to and from nearby locations. In light of this, the authors maintain that mamoas are to be related to movement on three different scales. On a large scale, the mamoas would have been situated on the central throughway crossing the main axis of the Barbanza ridge. On a medium scale, the barrows are related to the parts of the landscape that are most likely to have been local paths. Finally, on the smallest scale, the monuments would not only have provided a general sense of orientation, by marking roughly in what direction the traveler was meant to move, but would have indicated how best to negotiate the local terrain at any given location. In addition, the authors maintain that the location of the mamoas follows a well- established sequence of altering visual scenes (scene 1: low-lying areas with a reduced panorama; scene 2: higher areas with a wider panorama) that share essentially the same visual structure (Criado and Villoch 2000: 205):

1 They have a circular vista, visually enclosed and defined by topographic divisions within either a low-lying area or a basin. This is particularly apparent and logical in the barrows found close to watercourses, although it also occurs with those found on hills; in this case, the panorama is framed by lines of distant hills. 2 The border of this panorama is marked by natural features and artificial monuments: barrows can always be seen on the horizon, at times outlined against the horizon. 3 The scene tends to have a well-defined central point from which its totality may be observed, and which is identified by the presence of a barrow, situated either on the top of an elevated feature or in the axis of a depression. 178 Marcos Llobera

Simulations Prompted by these studies, I created a set of simple geospatial simulations to explore, and possibly extend, some of the interpretations put forward by Criado- Boado and Vázquez-Villoch (1998, 2000). The first simulation was designed to provide some context to observations made by the authors regarding the connection between the mamoas and movement. Given the lack of known paths for this period, I was interested in retrieving what would have been the most likely routes within and across the peninsula. Where were people going? Would have they been moving along the peninsula or across? What was the role of the mamoas? Where they key when determining these routes? The idea was to use a simulation to help me obtain some sense of the movement possibilities within the area, with the intention of later relating these to the barrows. More specifically, the simulation aimed at investigating movement on different scales – i.e. movement patterns with different destinations in mind – and to determine to which of these scales the distribution of mamoas related best. With these goals in mind, I constructed a deceptively simple simulation, as described here. A location representing the origin of a path was situated randomly anywhere within the peninsula. Three random destination locations were generated within the following distance intervals from the origin: 0–5 km; 5–12 km; and 12–19 km. The scale of these ranges is meaningful as they roughly correspond to the average width of the coastal plains surrounding the base of the Barbanza peninsula (short), the average Euclidean distance separating both sides of the penin- sula (connecting the local rias, or coastal inlets, located on each side) (medium), and the distance needed to cross the peninsula along its mountain ridge (long). An optimal path connecting the origin to each of the destinations was then created. Each path was calculated using anisotropic (varying dependent on the direction of movement) costs, based on metabolic energy consumption (see Llobera and Sluckin 2007) when moving along a gradient. Thus minimizing cost of movement across the terrain was determined in terms of tacking slopes at an angle (i.e. walking slope) rather than by the terrain slope alone. Though this does not consider other possible factors (e.g. vegetation, cultural landmarks) that may have affected the direction of movement, it serves my goal of describing the potential for different scales of movement within the Barbanza landscape. It is worth mentioning the resemblance between this simulation and the model employed by Murrieta Flores (2012) to explore movement patterns within Sierra Morena Occidental in southwestern Spain (for other similar examples, also see Cerrillo Cuenca 2007). Both simulations are designed to model the role that topography is likely to have played in shaping the choice of routes within the landscape. There are some notable differences, however, between the two approaches. Within the earlier approach, the pool of candidate origin and destination locations is limited to passes. These are further restricted to locations at the periphery of the study area (delineated arbitrarily around the Sierra Morena occidental region), i.e. entry points. It is only in relation to these entry points that Working the digital: some thoughts from landscape archaeology 179 least-cost paths are calculated and accumulated. The simulation presented here is more generic in nature, and allows for better control over issues of scale. With the calculation of each path, the simulation recorded various pieces of information (e.g. distance along the path, frequency and number of mamoas seen along the path, minimum distance from path to each mamoa, and so on). In addition, paths within each of the three distance ranges were added together. The simulation generated in excess of 16,000 travels for each scale of movement. As a result, locations that were traversed more frequently ended up being assigned higher values. The outcome obtained after adding all iterations (one raster image per scale of movement) was further transformed into a density surface describing the frequency of path intersections within 100m of each location. At first glance (see Figure 10.2), the mamoas are closely related to every scale of movement. However, their association with short-distance movement is definitely not as strong as it is with medium and long-distance movements. These patterns become a bit clearer when we consider only those locations with a high number of path intersections within 100 meters (i.e. those representing percentage values

Local range (< 5 km) Medium range (5–12 km) Long range (12–19 km)

Mamoas <10%

50%

>90%

35% 100% 70%

20% 60% 40% Terrain Mamoas 5% 20% 10%

<10% >90% <10% >90% <10% >90% 40–50% 40–50% 40–50%

FIGURE 10.2 Density of path intersections for different scales of movement. Gray scales are related to deciles. Darker shades correspond to locations where the density of interceptions within 100m was higher. Histograms at the bottom compare the distribution (percentage) of mamoas across the different deciles (in black) against the distribution (percentage) of locations for the entire Barbanza peninsula (which should be roughly similar in each decile). This comparison clearly shows that mamoas are situated on locations with a high density of path intersections. 180 Marcos Llobera in the upper twenties) (see Figure 10.3). We can view these locations as forming corridors across the landscape. For long distances, the mamoas are located on the longest continuous stretch crossing the peninsula, and on or near the only corridor along the Barbanza ridge. For medium distances, the barrows are definitely related to corridors that connect both sides of the upper (northern) part of the peninsula (see Figure 10.3). While not formally tested (due in no small part to the formidable difficulties of extracting information from a 1941 1:50,000 map with a different datum), it is worth remarking the apparent correspondence between these corridors and many known historical paths. Given that these corridors were generated independently from the mamoas, it is clear that the mamoas did not determine the direction of travel but rather reinforced it or exploited the presence of pre-existing travel corridors in some way. This outcome is quite remarkable, and while it does not contradict Criado-Boado and Vázquez-Villoch’s overall claims about the close relationship between these monuments and movement, it does further clarify the nature of this relationship. The simulation provides us with some additional observations. First of all, there are several (possibly up to five) long-distance corridors crossing the peninsula north to south, but the only one that climbs up to the Barbanza ridge is the one on which the mamoas can be found (Figure 10.3b). The others primarily cross the coastal lowlands bordering both sides of the ridge. With regards to movement across the Barbanza peninsula (rather than along it), there are several, possibly up to four, well-defined corridors (Figure 10.3a). As mentioned, mamoas are found on the corridor that cuts across the northern part of the peninsula. If we combined both types of movement, along and across the peninsula, it is clear that mamoas are located on a privileged crossroads, an area that offered the most choices in travel

(a) (b)

FIGURE 10.3 Corridors of movement generated for medium and long-range movement. These corridors were generated after selecting locations with high density of path intersections within 100m (upper 20 per cent). Note (a) various corridors crossing the Barbanza ridge from side to side and (b) the location of the mamoas on the only corridor that crosses along the Barbanza ridge. Working the digital: some thoughts from landscape archaeology 181 directions. An observation that could not be corroborated with the simulation, however, is the alleged role that nearby mamoas would have played in directing people where to move next when negotiating the local terrain. The simulation produced corridors that were well defined and continuous rather than intermittent, so it would have been unnecessary for mamoas to mark the way, which was pre- established. It is nevertheless possible that this might be a shortcoming of the simulation due to the scale at which it was generated – i.e. the use of a 5x5m2 cell- size DEM and the 100m radius used to generate the corridors – though a cursory change in the radius used to generate corridors (not shown here) down to 10 meters did not change this finding. Based on the fact that mamoas appear to have been built on pre-existing corridors of movement, I decided to explore the role that these may have played in shaping people’s experience while moving along them. Are the viewing conditions of the mamoas when moving along these corridors particular? More specifically, what is the mamoas’ relationship with local horizons and how do they appear to the viewer (what is their angle of incidence and elevation)?1 To investigate these questions, a set of simple simulations was devised. In these simulations, the north–south travel corridor crossing the Barbanza ridge that had resulted from the previous simulation was isolated, and seeded with 5000 viewpoints located at random. From each of the viewpoints, the horizon, distance from horizon, and angle of elevation and incidence were calculated (see below). Horizon locations (i.e. in this study, locations within 5m of where a sight line has been interrupted) were added together in order to generate a horizon index, which indicates how often a location has been part of a horizon. The values for all of these properties were recorded at each of the mamoas (as mentioned earlier, each mamoa was represented by a set of points). The relationship between horizons and monuments, which has been frequently noted by archaeologists (e.g. Bradley 2000; Tilley 2004; Bourgeois 2013), has been interpreted in various ways. For instance, when barrows appear on top of a horizon (i.e. “cresting” a horizon), they have been interpreted as being related to long- distance viewing (Bourgeois 2013: 107), marking the edges of scenes and/or connecting distant points on the landscape. In these locations, barrows are less likely to be related to concrete, local viewing areas. When used as viewing platforms, they have been associated with broad vistas. Overall, when barrows form part of a horizon they have frequently been interpreted as symbolic (e.g. they are seen as constructing a sacred landscape and/or operating as a nexus between the earth and sky). When barrows are off horizons it seems likely that they are naturally meant to be seen from concrete locations, and in particular directions within the local surroundings. This suggests that their role was closely related to celebrations of ancestry, lineage and/or individuals. While these connotations derive partially from ways in which the barrows can frame perception given their location, they are not either/or choices. The results of the second simulation break down what would otherwise have been an integral experience. These are presented through a series of normalized histograms (see Figures 10.4 and 10.5). In the first histogram (Figure 10.5a), the 182 Marcos Llobera distribution of horizon indexes has been divided into deciles. It shows the percentage of locations (for the entire Barbanza peninsula and the area within the corridor) and of mamoas that fall within each decile. Mamoas appear to be associated primarily with locations with a high horizon index; that is, they are situated on locations that would most frequently have been horizons (representing the upper 30 percent of the distribution of horizon indexes values obtained), though they also appear on locations with low horizon indexes (2nd–4th percentile). In either case, the frequency found at each horizon index is far higher than found (expected) within the Barbanza peninsula and within the area forming the corridor. The total absence (or drop) in the presence of mamoas in locations with medium horizon indexes (5th–7th percentiles) is also noticeable. All this indicates that while the majority of mamoas (around 60 percent) were preferably sited on locations that were more frequently part of horizons (or on immediately neighboring locations), there were some instances (around 25 percent) that would have appeared on a horizon only occasionally. Figure 10.5b shows the percentage of times that mamoas have been associated with certain distances from the horizon. A sizable percentage of mamoas (around 18 percent) would have frequently appeared on horizon locations or nearby. A slightly larger percentage (20 percent) of times, however, mamoas would have been built on locations that avoided horizons by relatively short distances. This points towards an intention to associate the viewing of the barrows with particular viewing locations and directions (as remarked by Criado-Boado and Vázquez-Villoch 2000: 206). Based on the results obtained for the angle of elevation (see Figure 10.5a), we see that the majority of mamoas are concentrated in the part of the chart that represents angles of elevation between –4° to 2° when visible. This means that while walking along this corridor, individuals would have been able to view (engage with) the monuments with relative ease most of the time. The small angular values would have required people to lift their heads a few degrees above the most comfortable (–10°) viewing angle. However, people would have been able to continue moving while viewing the mamoas. Indeed, this supports the interpretation that within the corridor mamoas were perhaps not meant to be understood as fixed destinations (though they could have operated in this way) but rather as beacons along the way. The values obtained for the angle of incidence (see Figure 10.5b) provide some additional information regarding the appearance of the mamoas. We observe several distinct settings. The lower values correspond to the number of times when mamoas were situated on horizon-like locations (<5°) – around 14 percent; and on locations that would have appeared to be longitudinal planes (<15°), i.e., their flat appearance would have made them look as being easily accessible or approachable – around 20 percent. The higher values (>15°) highlight the large number of times mamoas appear as frontal planes (around 66 percent). Typically, this occurs when mamoas are located on the side of the valleys. It is important to address the slight mismatch between the percentage of mamoas that are located on horizon-like locations according to their angle of incidence and the percentage obtained from the horizon index. This lack of concordance may be (a) (b) Horizon index Distance to horizon 35 Barbanza corridor mamoas 30 35

25 20

oas 20 m

oas 15 a m m 15 a m % 10 10 % 5 5

0 0 1st 2nd 3rd 4th 5th 6th 7th 8th 9th 10th 0–5 >200 Percentile 10–15 20–25 30–35 40–45 50–55 60–65 70–75 80–85 90–95 100–105 110–115 120–125 130–135 140–145 150–155 160–165 170–175 180–185 190–195 Distance (m)

FIGURE 10.4 (a) Histogram showing the distribution (percentage) of times that locations within the Barbanza peninsula (white), within the Barbanza ridge corridor (light gray) and where mamoas are positioned (dark gray) are associated with certain horizon indexes. Higher horizon indexes represent locations that have been part of a horizon more frequently. (b) Histogram showing the distribution (percentage) of times when mamoas have been associated with certain distances from horizon locations. 184 Marcos Llobera attributed to several factors: first, the results were obtained using different simula- tions, though with the same parameters (ideally these are obtained from a single procedure); second, the effect of terrain resolution (in theory an angle of incidence of 0° should correspond to a horizon, but in reality values found on horizons oscillate due to cell resolution); third (also related to terrain resolution), the rapidity with which the character of the planes near horizons can change from frontal to longitudinal (very rapidly, given our current resolution). Nevertheless, if we take the values associated with the angle of incidence as indicative of the appearance of the mamoas as viewed from within the Barbanza ridge corridor, we can say that they primarily appear to be bordering the corridor (when located outside) or “breaking”/“marking” the traverse into distinct stages (when located within). The impression that we can piece together from this information is not very different from that described by Criado-Boado and Vásquez-Villoch (2000). Along the Barbanza ridge corridor, the mamoas would have weaved in and out of local horizon points. Viewing the mamoas would not have required travelers to pause or break while moving. For the most part, mamoas would have appeared vertical and erect; they would have been encountered as marking stages along the way. It is interesting to note that whenever they have been ascertained, chamber orientations of the mamoas are east-southeast facing. Indeed, throughout Galicia, the significance of this orientation appears to be reinforced through the use of stones with different colors. Quartz or other white stones are used to fashion a sort of “shield” centered around the eastern side, while dark stones are used to mark the western side (see Criado-Boado 2012: 309; Gianotti et al. 2011). This introduces an interesting twist to our interpretation. Mamoas may have operated very differently depending on the direction of travel. In this particular case, I explored their role in relation to the longitudinal corridor that crosses along the Barbanza ridge (what I refer to as the Barbanza ridge corridor), following the lead of Criado-Boado and Vázquez-Villoch (2000). Medium-range corridors traversing from one side of the peninsula to the other (i.e. connecting the rias or coastal inlets on both sides) were not investigated. These corridors, running primarily east–west, would align with the mamoas’ main axes of orientation, and may therefore be associated with very different visual experiences. Unfortunately, due to time and space limitations this is not something I can address in this chapter, so it remains pending for future work.

Conclusions Overall, the results of the simulations discussed here are in line with the findings reported by Criado-Boado and Vázquez-Villoch. Many of these were originally based on repeated and extensive field observations (Criado-Boado, personal communication), hence, their convergence on the results obtained through the use of digital information is noteworthy. The total running time of the different simulations amounted to about four days (with the benefit of hindsight, this could be shortened to a day and a half). In addition to providing a new line of evidence that supports existing claims, the use of digital information has also helped to (a) (b) Angle of elevation Angle of incidence 35 20 18 30 16 25 14 12 oas 20 oas m m a a 10 m m 15 8 % % 10 6 4 5 2 0 0 0–1 1–2 2–3 3–4 0–5 <–7 –1–0 5–10 10–15 15–20 20–25 25–30 30–35 35–40 40–45 45–50 50–55 55–60 60–65 –7––6 –6––5 –5––4 –4––3 –3––2 –2––1 Degrees Degrees

FIGURE 10.5 (a) Histogram showing the distribution (percentage) of times mamoas were viewed within certain angles of elevation. Angle of elevation is the vertical angle at the viewer location of the sight line, measured positively above the horizontal (0°) and negatively below. (b) Histogram showing the distribution (percentage) of mamoas as viewed within certain angles of incidence. Angle of incidence is the vertical angle between the terrain and the sight line measured at the target location. 186 Marcos Llobera

“contextualize” and extend previous observations. The delineation of other sig- nificant corridors that may have been used to travel across the peninsula is a good example. As shown in this case, when used properly, digital information and methods can play an important role when constructing arguments that lean on field observations. The simulations used here were designed to explore very basic properties of the study area. Understanding these properties, how they vary and are distributed throughout the landscape, constitutes one of the building blocks on which more nuanced interpretations can safely be built. One of the benefits of addressing these elemental aspects is that it allows us to build models that we can easily comprehend. In this paper I purposely used very simple exploratory models in order to keep the volume of output manageable (simulations can generate huge amounts of data), and to facilitate the interpretation of results. It is important to clarify that this study does not show how mamoas changed the previously existing (visual) experience of moving through the Barbanza ridge corridor. Instead, I have characterized some of the visual properties that are associated with mamoas as they are today. While my original intention was to focus on how the mamoas altered previous experience, our ability to explore this question, at this point in time, is limited by the possibilities offered by the GIS application used (in this case, ArcGIS, though the same limitations are likely to apply to other systems). This serves as a reminder that in spite of technological advances, we are still short of realizing the full potential that digital information has to offer. The importance of representation was brought up at the beginning of this paper. It seems almost trivial – almost comical – to remark on the importance of form and representation when discussing digital information. After all, they go hand in hand. However, the impact of predominant forms of spatial representation, through the use of GIS, has gone largely unquestioned in many landscape studies that use such applications. For example, there has been little discussion about, or even acknow- ledgement of, the effect of using a single-point feature to represent monuments on visibility studies (such as the investigation of visibility and intervisibility patterns) or on establishing (usually unsuccessful) comparisons between field observations and those obtained through digital means (e.g. Cummings and Whittle 2004). In this study, I used a cloud of points to overcome this limitation to some extent, but this is less than ideal. In a perfect world, we want to be able to integrate the built environment with landscape models and to develop tools that will allow us to move entire architectures (i.e. not just points) and/or alter landscapes at will (something that is within technical reach but has yet to be realized). Our approach to digital information requires a similar orientation, a similar care, as that observed when building a model where questions about scale, form, range of variation, and so on are paramount. As I have argued elsewhere (Llobera 2011), the advent of digital information paired with the capabilities of computing can, at the very least, extend our analytical capabilities and provide new evidence, regardless of the theoretical framework. Viewed as a wider and thoughtful effort, rather than the mechanical application of Working the digital: some thoughts from landscape archaeology 187 software applications, these possibilities have the potential to open new research themes and venues.

Acknowledgements Any contribution I may have made through this chapter would have not been possible without the generous help of several people. First, I would like to thank Jake Deppen (a doctoral student at the University of Washington), who kindly assisted me in the running and collating of the results of the simulation described above. I would also like to thank Cesar Parcero Oubiña, Pastor Fábregas Álvarez, and Felipe Criado-Boado (Heritage Institute, CSIC, Spain) who, in addition to providing me with the data I needed to do this study, were gracious enough to entertain my questions and comments on this set of monuments. I would also like to thank Alison Wylie and Bob Chapman for kindly inviting me to contribute to this volume and for providing many insightful comments on an earlier draft. Of course, any mistakes found in this article are mine.

Note 1 The angle of elevation is the angle between the horizontal line at the viewer’s eye level and the sight line between a viewer and an object. When negative, this angle may be related to whether an object is sensed as being near or far (see angle of depression, Higuchi 1989: 36–49). When positive, it is associated with ease of viewing. Objects that are situated above the horizontal line require more effort to view. As the angle gets larger, the viewer’s mobility and (generally) capacity to engage with certain actions while viewing become more and more limited (Higuchi 1989). The angle of incidence is the angle that a viewer’s sight line makes with the terrain (here we assume that the terrain is made of many small planes). When this angle is large (>15°) the plane is said to be frontal and its appearance to be that of a wall, barrier, or obstacle. If the angle is smaller (<15°), the plane is said to be longitudinal, and it appears to afford walking on it.

References Bevan, A. and Lake, M. (eds) (2013) Computational Approaches to Archaeological Spaces, Walnut Creek, CA: Left Coast Press. Bevan, A. and Wilson, A. (2013) “Models of settlement hierarchy based on partial evidence,” Journal of Archaeological Science, 40: 2415–2427. Bourgeois, Q. P. J. (2013) Monuments on the Horizon: The Formation of the Barrow Landscape throughout the 3rd and 2nd Millennium BC, Leiden: Sidestone Press. Bradley, R. (2000) The Archaeology of Natural Places, London: Routledge. Cerrillo Cuenca, E. (2007) “Recorriendo un territorio desaparecido: restitución fotogramétrica y análisis del paisaje de la necrópolis prehistórica del Vado de Alconétar,” in V. Mayoral Herrera and S. Celestino Pérez (eds) Técnologia de información geográfica y análisis arqueológico del territorio: actas del V simposio internacional de arqueología de Mérida, Mérida, Spain: Consejo Superior de Investigaciones Científicas, 145–159. Crema, E. R. (2012) “Modelling temporal uncertainty in archaeological analysis,” Journal of Archaeological Method and Theory, 19(3): 440–461. Criado-Boado, F. (2012) Arqueológicas: la razón perdida, Barcelona: Bellaterra. 188 Marcos Llobera

Criado-Boado, F. and Vázquez-Villoch V. (1998) “La monumentalización del Paisaje: percepción actual y sentido original del megalitismo de la sierra de Barbanza (Galicia),” Trabajos de Prehistoria, 55(1): 63–80. Criado-Boado, F. and Vázquez-Villoch V. (2000) “Monumentalizing landscape: from present perception to the past meaning of Galician megalithism (northwest Iberian peninsula),” European Journal of Archaeology, 3(2): 188–216. Criado-Boado, F., Aira Rodríguez, M. J. and Díaz-Fierros Viqueira, F. (1986) La Construcción del Paisaje: megalitismo y ecología. sierra de Barbanza, Santiago de Compostela, Spain: Xunta de Galicia. Cummings, V. and Whittle, A. (2004) Places of Special Virtue: Megaliths in the Neolithic Landscapes of Wales, Cardiff Studies in Archaeology, Oxford: Oxbow Books. Eve, S. J. and Crema, E. R. (n.d.) “A house with a view? Multi-model inference, visibility fields, and point process analysis of a Bronze Age settlement on Leskernick Hill (Cornwall, UK),” Journal of Archaeological Science, 49(1): 267–277, doi: 10.1016/j.jas. 2013.12.019. Gianotti, C., Mañana Borrazás, P., Criado-Boado, F. and López Romero, E. (2011) “Deconstructing Neolithic monumental space: the Montenegro enclosure in Galicia (northwest Iberia),” Cambridge Archaeological Journal, 21(3): 391–406. Gillings, M. (1998) “Embracing uncertainty and challenging dualism in the GIS-based study of a palaeo-flood Plain,” European Journal of Archaeology, 1: 117–144. Higuchi, T. (1989) The Visual and Spatial Structure of Landscapes, Cambridge, MA: MIT Press. Knappett, C. (ed.) (2013) Network Analysis in Archaeology: New Approaches to Regional Interaction, Oxford: Oxford University Press. Murrieta Flores, P. (2012) “Understanding human movement through spatial technologies: the role of natural areas of transit in the Late Prehistory of south-western Iberia,” Trabajos de Prehistoria, 69(1): 103–122. Llobera, M. (2003) “Extending GIS-based visual analysis: the concept of visualscapes,” International Journal of Geographic Information Science, 17(1): 25–48. Llobera, M. (2011) “Archaeological visualization: towards an archaeological information science (AISc),” Journal of Archaeological Method and Theory, 18(3): 193–223. Llobera, M. (2012) “Life on a pixel: challenges in the development of digital methods within an ‘interpretive’ landscape archaeology framework,” Journal of Archaeological Methods and Theory, 19(4): 495–509. Llobera, M. and Sluckin T. J. (2007) “Zigzagging: theoretical insights on climbing strategies,” Journal of Theoretical Biology, 249: 206–217. Lucas, G. (2001) Critical Approaches to Fieldwork: Contemporary and Historical Archaeological Practice, London: Routledge. Miller, H. J. and Wentz, E. A. (2003) “Representation and spatial analysis in geographic information systems,” Annals of the Association of American Geographers, 93: 574–594. Peuquet, D. J. (2002) Representations of Space and Time, New York: Guilford Press. Tilley, C. (2004) The Materiality of Stone: Explorations in Landscape Phenomenology, Oxford: Berg. White, D. A. and Surface-Evans S. L. (eds) (2012) Least Cost Analysis of Social Landscapes: Archaeological Case Studies, Salt Lake City, UT: University of Utah Press. 11 CRAFTING KNOWLEDGE WITH (DIGITAL) VISUAL MEDIA IN ARCHAEOLOGY

Sara Perry

Visual producers have a deep and inseparable relationship with the institutionalisa- tion and development of archaeological practice. Their role in articulating con- cepts, circulating knowledge, refining interpretations, and publicising sites, finds and features – indeed demarcating those sites/finds/features in the first instance – is hardly a point for contention today. That role is increasingly attested to, not only by varied scholarly and professional investigations (e.g. Earl 2013; Llobera 2011; Moser 2014) but also by visualisers themselves, who chronicle their work and process both in print and online forums. Such chronicling is far from new, as evidenced by the reflective publications of, for example, the early twentieth- century American reconstruction artist Charles Robert Knight (e.g. Knight 1946), mid-twentieth-century British archaeological illustrator Alan Sorrell (e.g., Sorrell 1973), or current ‘palaeo-artist’ John Gurche (e.g. Gurche 2013), among others. It is complemented by biographical meditations on the productivity of scientist- artist partnerships, such as the First World War-era anthropologist Aimé Rutot’s positive assessment of his working relationship with the Belgian sculptor of prehistoric ‘portraits’ Louis Mascré (Rutot 1919; see also Knight’s autobiographical reflections, cited in Cain 2010, on his ‘mutually helpful’ collaborations with Henry Fairfield Osborn, curator at the American Museum of Natural History at the turn of the twentieth century). Similarly, it is extended by a growing number of weblog and social media-based archives, wherein contemporary producers are circulating their often in-progress visual outputs for comment (e.g. Swogger 2014; Watterson 2014), thereby opening themselves and their practice up to the scrutiny and critique of myriad audiences. These archives are not necessarily artistic portfolios or online repositories of final imagery, but spaces for argument and conceptual refinement, helping to expose the intellectual work at the heart of archaeological visual media. 190 Sara Perry

Despite such a proliferating body of knowledge about their skillsets, epistemic contributions, and on-the-ground impact, however, visualisers working in archaeo- logical circles still sit somewhat uncomfortably within the discipline’s architecture. Often consigned to archaeology’s sidelines, it is not uncommon to see their credibility, relevance, and financial worth challenged, their subject areas eliminated from university curricula and their work outsourced to unpaid or severely under- paid labourers. Recent surveys of the archaeological illustration sector in the UK by Hodgson (2008; also Aitchison 2011; Gibbons 2011) suggest that opportunities for career progression are poor, more than 50 per cent of practitioners are pressured to undercharge for their services, and the majority of employees earn under £20,000 per year – with a surprising proportion of freelancers earning less than £5,000 per annum. Aitchison (2011) goes further, demonstrating that illustrators and photographers charge less for their services than other archaeological specialists (e.g. conservators, surveyors, finds experts) and appear to be experiencing the most acute reductions in their workload. At the same time, these speciality areas seem to be at particular risk of loss of skills owing to retirement and related intentions to leave the sector. The predicament appears to be made worse by a lack of formal training programmes, including the recent closure of the UK’s only postgraduate course in archaeological illustration (at Swindon College), as well as by the seeming competitiveness, oversensitivity, and misinformation circulating within the archaeological visual community itself (see below). James’s (in prep) manifesto on visual competence (or, more accurately, lack thereof) within the archaeology sector highlights further challenges, such as the almost total absence of concern for visualisation skills within standard-setting pedagogical frameworks like the UK Quality Assurance Agency’s benchmarking document for university courses in archaeology (QAA 2007). Whilst that document is now due for revision, it is notable that in spite of archaeology’s long-standing stake in honing visual expertise, the only apparent model of good practice available is, as James points out, a recent set of American interdisciplinary guidelines for visual literacy competency in higher education (Hattwig et al. 2011). These findings are important, particularly in light of the fact that, as Morgan (2012: 77) writes, studies about (visual) media made by archaeologists are rare. So whilst enquiries into archaeologists’ engagements with different media outlets are becoming more and more commonplace (e.g. Clack and Brittain 2007; Schablitsky and Hetherington 2012), as are analyses of artists’/visualisers’ varying contributions to the discipline (e.g. Russell and Cochrane 2014), the same does not seem to be true of research into archaeologists themselves as creative makers (Morgan’s work, however, is one exception, alongside the emerging projects of, for example, ‘punk archaeology’ practitioners, e.g. Caraher et al. (in press)). One might infer from this gap that there are institutional difficulties in simultaneously and equivalently equipping practitioners with high-level skills in both archaeology and media production. In other words, the development of competency in the former seems perhaps necessarily to demand that focus be partially turned away from the latter (and vice versa), if disciplinary expertise is to be honed fully. While there is room Crafting knowledge with (digital) visual media in archaeology 191 to discuss the value of nurturing programmes that explicitly groom archaeologist- artists, my concern here is to examine the situation that presents itself when such dual experts are scarce, and our visualising of the archaeological record thus depends upon specialist technical producers who may or may not have discipline-specific training. In all cases, my argument is that this relationship between visualiser and archaeologist is a critical one with profound implications for knowledge making and the tracing of epistemic genealogies in archaeology. Below I review the significant literature on the productivity of visualisation, moving from there into an examination of the on-the-ground work of multiple practitioners contributing to the visual representation of the Neolithic site of Çatalhöyük in Turkey. These examinations attend, in particular, to digital visual production, but with the explicit recognition that digital media are merely an extension of analogue media, and indeed that all visualisers necessarily work back and forth between the two. Whilst our tools might be changing, many of the fundamental practices, processes, and assumptions behind their application have stayed consistent across at least a half-millennium. To appreciate such continuity, I begin by contextualising graphic work in archaeology historically, and against a growing scholarship on the relationship between ‘craft’ skills and scientific development more generally. Visual producers have a literal hand in pushing forward paradigm change across disciplines – a distinction that is notable from pre-Renaissance times onwards. I consider this in relation to the present and future of archaeology, arguing that awareness of the affordances of visual outputs and proficiency in their crafting and circulation have deep consequences for the ongoing elaboration and basic sustenance of our profession.

The visualiser in archaeology A substantial literature now exists on the role of the visual producer in the development of archaeological practice. Arguably, the earliest truly critical eye applied to such production was that of Stuart Piggott, who outlines, across multiple publications (e.g. 1965, 1978), the constructed nature of archaeological draughts- manship. Piggott’s penetrative analyses emanate from his own engagements with the illustrative outputs of antiquarian scholars dating back to the 1600s, but also with the outputs of his contemporaries, who were implicated in the codification and professionalisation of the discipline (e.g. Wheeler 1954). Piggott might be viewed as setting the stage for a growing body of deconstructive studies that appear from the 1980s onwards (e.g. Moser 1992; Shanks and Tilley 1987: Chapter 4). These tend to centre upon the role of different forms of visualisation in propagating ideologically loaded notions of past peoples and places – for example, overtly gendered, or ethnically or politically skewed interpretations (among many others; see Piccini 1996; Wiber 1997). But Piggott might also be seen as spurring on enquiry into the historical and intellectual development of conventional visual styles and methods in archaeology, helping to focus our gaze on the earliest production of educational manuals (e.g. 192 Sara Perry

Petrie 1904), forums for honing practice (e.g. O. G. S. Crawford’s short-lived column in Antiquity on archaeological photography; see Crawford 1936), and whole ways of engaging with and ‘performing’ the archaeological record (e.g. Shanks 2012). What is increasingly clear from such scholarship is that, at least since the seventeenth century, illustrators have been recruited into the articulation of a ‘science of material culture’ (Moser 2014), turning antiquities into sources of data, and in so doing both shifting the types of questions asked about the past, and extending the nature and robustness of expertise on the subject. This research is complemented by philosophical reflections on the reasoning and informativeness afforded by current and historical practices of visualisation in archaeology (e.g. Lopes 2009), and the place of those practices in crafting a rigorous professional vision and knowledge base (e.g. Goodwin 1994). Running in tandem, then, are trajectories of study concerned with the ‘honest epistemic work’ (Lopes 2009: 13, 16) of archaeological imaging, and at once, its fabricated, potentially misleading and thus dishonest representations. The conflict here arguably comes to a head in recent digital visualisation research where, in an effort to both establish its epistemic productivity and shield itself from accusa- tions of dishonesty, codes of practice (e.g., the London Charter) and major methodological compendia are now being outputted which self-describe as ‘solely dedicated to the issue of intellectual transparency of visualization-based research’ (Bentkowska-Kafel and Denard 2012: 4). In what might be perceived as an overreaction to critiques of the persuasive rhetoric of the visual, specialists appear increasingly to be investing in efforts to define, inventory, and quantify visualisation practice, designing systems literally to spell out the illustrative process, or otherwise to open it up to peer review (e.g. Opitz et al. 2013). While the latter initiative has potential to foster meaningful dialogue and to intersect with the weblog-based discussion forums noted above, the former seems naïve and short-sighted. For example, Beacham, in his chapter “Defining our Terms in Heritage Visualization” (which launches Bentkowska-Kafel et al.’s 2012 keystone text on archaeological visual paradata), writes of those producing ‘popular’ virtual reality models:

Although scholarship of this dubious sort may draw attention (and even vital funding) to those creating the visualization, ultimately it carries the risk of discrediting the reputation of visualization-based research which must be protected if such visualizations are to be perceived and taken seriously by scholars as the extraordinarily valuable ‘publications’ they undoubtedly have the potential to be. (Beacham 2012: 10)

Such argumentation ignores the fact that visualisers (whether ‘popular’ or not) have long been concerned with producing evidence-based images (e.g. Perry and Johnson 2014), and with articulating terms by/on which archaeological visualisa- tions should be constructed and judged (e.g. Hodgson 2001). Moreover, these practitioners (as below) tend to follow personalised, but still highly systematic Crafting knowledge with (digital) visual media in archaeology 193 courses of action in crafting their images, grounded in intensive research and drawing upon a range of data points to inform their composition. Yet they also have a concern for artistry and flourish that pulls audiences into the visual narrative and engages the imagination. That skill is arguably impossible to inventory because, borrowing from the archaeological illustration instructor Grahame Smith (in Perry 2010), at its core is a kind of ‘creative ambiguity’ and ‘imaginative dissonance’ that is effective precisely because of its unpredictable ability to resonate. What is important here is the notion that archaeological visualisation has an inspiration and soul to its practice that renders it more than mere ‘publication’, and that thus extends its relevance beyond researchers alone. Archaeologists have long struggled to reconcile themselves with this volatility of the image (which, in fact, is characteristic of the archaeological process overall), and the typical response has been, as now, to strive for some ultimate certification of transparency – of the accuracy of the intellectual endeavour. I would suggest that this continues to be a doomed effort: impossibly onerous and reductionistic, presupposing that all aspects of the representational process can be isolated and captured in an archive. Decades ago, the editors of Current Archaeology lamented that ‘Archaeologists have no soul’ (Selkirk and Selkirk 1973: 163). I understand them to mean that practitioners (prompted, arguably, by the uncertainty of their datasets) often collapse their work into uninspired documentary reports, beating all the soulful and expressive detail out of the archaeological record as a result of their risk aversion. This confused act of collapse is seemingly meant to mask the craftwork of archaeology (Shanks and McGuire 1996) behind thoroughly catalogued and defined evidence bases – as if the two are mutually opposed. Critically, ‘soul’ resides in the skill of the practitioner: a set of competencies built up through forms of apprenticeship and performance that are not always easily describable. As Smith (2013: 175) puts it in her commentary on ‘making things’: ‘we still do not entirely know what to make of handwork, for much of it involves tacit knowledge, which is hard to codify in writing because it requires acute observation and attention to the circumstances of the ephemeral moment’. Archaeologists and archaeological visualisers alike are usually attuned to such work, particularly in the field where, for example, the subtleties of soil colour and texture demand a nuanced eye. Jordanova hints at the complexity, then, of attempting to catalogue these nuanced corporeal practices: ‘some skills are so ingrained in the body that practitioners would be hard pressed to put them into words. As a result, many processes remain unarticulated, best understood either by direct observation or first-hand experience’ (Jordanova 2012: 66). However, the archaeological sector seems ever prepared to undermine itself by devaluing, misunderstanding, or entirely ignoring such skills. This is perhaps most obvious amongst the illustrative community, where practitioners not uncommonly disparage their own colleagues, especially those who engage in digital archaeologi- cal visualisation. The argument here appears to be that digital outputs can never be as expert and impactful as those produced by hand. Some go so far as to proclaim 194 Sara Perry that the ‘tyranny of computer graphics’ might make specialist archaeological visualisers obsolete altogether (Read and Smith 2009), presumably replacing them with vapid automated processing. While digital visualisers have seemingly only worsened the predicament by investing so heavily in attempts to define accounting and compliance measures, the evidence used by their critics is dubious. Data from a comprehensive survey of British archaeological graphics teams suggests that 83 per cent of graphics staff are 30 years of age or older, 58 per cent have at least 11 years of experience (34 per cent with more than 20 years), all of their offices self- describe as primarily digital working environments, and yet 63 per cent still use a combination of digital and ‘traditional’ visual methods or ‘traditional’ methods exclusively (Gibbons 2011). In their lack of awareness of this range of practice – and in their arguably irreconcilable attempts to, on the one hand, make transparent and, on the other, respect the ineffable soul of the image – archaeological visual producers seem to be their own worst enemies: they sideline themselves within the discipline, focusing on establishing the specialness of their visual outputs, therein removing themselves from general, everyday archaeological discourse. What is critical is that the history of visual representation in the sciences testifies to the importance of visual producers in that discourse. Their skilful and seamless participation in the scientific process over multiple centuries has meant their implication in critical intellectual revolu- tions and major paradigm change. Understanding their role here, then, offers an opportunity not only to tease out the actual relationship between visualisation and knowledge making (in archaeology and beyond), but to suggest more productive engagements between visualisers and the larger academic/professional sector in the future.

Visual skill and scientific ‘revolution’ A growing body of literature is accumulating on the skills and expertise of visual producers and their relationships to epistemological change. In these conceptualisa- tions, artists may be likened to scientists themselves, deploying close observational methods, fieldwork, collecting practices, imitation, experimentation, efforts at replication, site visits, independent research, and recording in the name of knowledge making. Smith’s (2004) pioneering volume on The Body of the Artisan speaks to such an ‘artisanal epistemology’, wherein bodily engagement through creative acts (from illustration to metallurgy to pottery production and alchemy, etc.) can be understood as literally creative: generative of new ways of thinking and doing, and of otherwise engaging with and interpreting the natural world. By this reckoning, artistic practice is inseparably bound up in, and constitutive of, the arguments, artefacts, and authority that comprise scientific knowledge. Fundamental to such practice, however, is a kind of ‘artisanal literacy’, or under- standing through actual labour, which provides hands-on comprehension of materials, their specificities and malleability, that is distinct from text-based or verbal forms of understanding (Smith 2004: 8). Crafting knowledge with (digital) visual media in archaeology 195

Smith’s research builds on multiple streams of study into the work of art and craft in science, including its shaping and driving forward of the Scientific Revolution. Here, visual producers have been appreciated as intimate collaborators in reconceptualising energy, space, maths, mechanics, etc. – or as pivotal inventors/scientists themselves (e.g. Edgerton 1991; Field 1997). Representational modes have been recognised as being strategically harnessed in the enactment of experimental science. The visual is thus seen, in certain circumstances, as ‘positively kaleidoscopic in its generativity’ (Hunter 2013: 25). Indeed, as Hunter grandiosely describes the impact of Robert Hooke’s seventeenth-century construction of a paper model micrometer, ‘it began life as a picture, then matured as an object at a nexus of technological competition, artistic skill, and frankly wild speculation among leading French and English experimentalists before giving birth to varieties of conceptual shape-shifting . . .’ (Hunter 2013: 25). However, what is problematic about many of these studies, especially the earliest, is their tendency to overlook what Smith (2004: 21) calls the ‘epistemologi- cal status of craft operations’; that is, artisanal practice as cognition itself. In her interpretation, making, doing, performance, and reworking – the mistakes and messiness and embodied learning at the core of creative practice – are crucial to knowing and generating new knowledge of the world. Elsewhere, Smith (2013: 180) refers to the ‘maker’s “philosophy” or “vernacular science”’, which can be likened, I think, to Bentley’s (in Auslander et al. 2009: 1,386) commentary on insiders’ knowledge, wherein understanding ‘accrued from intimate, lived experi- ence can allow for a richness of meaning that is qualitatively different from the perspective of those without an experiential connection’. Smith and Bentley are among various scholars to draw attention to such ‘visual intelligence’, characterised by Jordanova (2012: 54) as ‘the constellation of attributes that makers possess’. By Jordanova’s logic, making is a process of problem solving dependent upon specialised ways of thinking. Sibum (in Auslander et al. 2009: 1,384, 1401) labels it ‘gestural knowledge’, and reflects on the sixteenth and seven- teenth centuries as a convergence point ‘in which things experienced by humans became the new mode of natural inquiry, and testimony and authority were resorted to only when individual experiential access was not possible’. Prak (2013) speaks of an ‘artisan revolution’ in Late Medieval/Early Modern Europe, wherein craftspeople were lauded as ‘the engineers of the past’. While much of this reflection is centred upon teasing out the contributing factors to intellectual revolutions of the past, it also seems increasingly to be applied to anthropological enquiries in the present. These suggest that it is via doing that we know the world around us; it is via participation that we come to think differently; it is via understanding in practice that we come to learn (e.g. Ingold 2013; Lave 1997). To borrow from Sennett (2008), such doing confronts us with resistances and ambiguities that we must work around, improvising, inventing in the moment, engaging sympathetically and intuitively with materials, yielding to some forces and pressing back against others. As per Ingold (2013: 7), doing both catches us up in ‘generative currents’ and cultivates within us forms of ‘sensory 196 Sara Perry awareness’. And Ingold goes further, to suggest that it is arguably among artists that we can best see this ‘art of enquiry’ at play. As he notes elsewhere (Ingold 2011: 6), ‘To read creativity “forwards” is to find it in the moment-by-moment inventiveness of practice – that is, in its improvisatory quality – as it carries on, in the midst of things, always responsive to what is going on in the surroundings.’ These scholars advocate for craft production and direct participation in craft making, not only because it situates us in the middle of – and necessarily responsive to – the flow of action around us (Smith 2013: 181), but also because it cultivates a moral ethic linked to good work, good citizenship, democracy, and a better understanding of ourselves and our productive relationships with other people and things (Sennett 2008). Even the basic act of copying, as evidenced in the training of scientific illustrators for centuries, testifies to such epistemic value, for copying recognises expertise, draws on and values the wisdom of preceding practitioners, reduces risk of error, introduces concepts of conventionalisation and economy of design, and so hones an artist’s manual skills and ‘visual judgment’ whilst still allowing space for invention and modification (Nickelsen 2006). Such skilful practice has led visualisers themselves to be recognised as intellectual authorities. Knight’s reconstruction work saw him appreciated as an expert on the subject of palaeontology by the popular press, and as a peer by some in the scientific community (Cain 2010: 295). Speaking of exhibition designers more broadly at the American Museum of Natural History in the early twentieth century, Cain (2011: 216) highlights how some were granted a kind of ‘scientific status’ by curators ‘by virtue of their field observations, their work with natural objects, and their passion for the natural world’. Today, there is insinuation that visual producers are again at the core of a paradigm shift – namely, the digital revolution (e.g. Sapsed and Tschang 2014). Herein, experiments with graphic creation, remixing, and analysis are arguably driving forward fundamental changes to the nature of media, aesthetics, innovation, and the form and dimensions of technology overall (e.g. see Manovich 2013). Within the sciences, digital visualisation is now understood as deeply implicated in the development of new disciplinary knowledge, with some also acknowledging the skill and effort invested in this imaging work (e.g. Kolijn 2013). Whole fields of practice have emerged that centre upon such visualisation (e.g. virtual anthropology; see Weber and Bookstein 2011), and some are suggesting ways in which extant disciplines might be reconfigured through comparable engagement (e.g. art history; see Bentkowska-Kafel 2013). However, the craft behind this digital visual production – the handiwork that simultaneously provides it with rigour and artistry, with the means to think about and think through the materials and subjects – is less obvious. Perhaps this is unsurprising given historical trends that suggest the practice of visualisation has long gone unmapped. As Jordanova notes,

Fully documented processes of making are relatively rare. What incentives were there for makers to record their production processes, even assuming this was possible? In any case, many wanted to keep the tricks of the trade to Crafting knowledge with (digital) visual media in archaeology 197

themselves . . . As a result, many processes remain unarticulated . . . hence the interest in re-enacting now obsolete forms of production. ( Jordanova 2012: 66)

Sibum (in Auslander et al. 2009: 1,359) further outlines the tense relationship between hand-work and head-work over time, arguing that ‘engagement with physical objects as a means to create knowledge has always challenged the identity of the scholar’. The challenge comes via having to negotiate rationalism and text- based intellectualism with embodied engagement and empathetic understanding. But the situation is arguably exacerbated in the digital world, where the physicality of the visualisation process is often masked by computational auto- mation. Skill can thus seemingly go unnoticed, leading some to devalue digital imaging for its lack of soul, and others to invest in transparency accounting on the chance that accusations of soulful flights of fancy are ever thrust at it. What appears to have been stripped from computer-based visual production is, to borrow from Cain (2012), the ‘craftsmanship aesthetic’: the literal sight of skilled laborers – good citizens – doing the hard, identifiable, generative work that Sennett (2008) positions as being at the core of democratic society (as in Figure 11.1). In others words, the actual craftsperson and his or her individual expertise are typically made invisible in digital visualisations, arguably through the rendering process itself. Yet the act of rendering – characterised by long, frustrating, and often

FIGURE 11.1 The craftsmanship aesthetic: Members of the Çatalhöyük visualisation team and site staff watch graphic designer Ian Kirkpatrick prepare material for ultimate display in the Çatalhöyük Visitor’s Centre. (Photograph: author.) 198 Sara Perry unsuccessful processing periods, sometimes lasting for days, as computers transform the artist’s working file into a polished product – is itself an echo of older trial-and- error routines of craft production (e.g. Smith’s experiments with metal casting where the ‘mind is held in suspense and fear regarding the outcomes’; Smith 2013: 193). Other unseen and lengthy activities – for example, participating in chat forums dedicated to highly specialist programs or techniques that seek to offer communal advice and critique – also have precursors in the craft guilds of the Early Modern era, which acted as essential spaces of knowledge transmission, training, and networking. Thus, despite their seeming automation, the effort, negotiation, experimentation, and craft underpinning many digital visualisations today retain a distinctly human origin. Our inability or unwillingness to recognise this, however, arguably makes us ill equipped to appreciate the genuine epistemic productivity of digital visual production.

Crafting knowledge at Çatalhöyük The Neolithic site of Çatalhöyük in central Turkey provides a perfect base from which to assess such productivity in detail. Here, a series of visualisers in the recent and distant past (especially since 1993, when the site began to see continuous excavation) have played crucial roles in conceptualising Çatalhöyük’s prehistory, creating imagery that has directly driven forward or otherwise meaningfully contributed to academic interpretation and re-interpretation of the archaeological record. Just how evident such contributions have been to our understandings of the site, however, is debatable, and I would suggest that this relates to the precarious position of visual producers both at the site itself and in the discipline at large. While one might make the argument that experimentation with visualisation extends back to the first excavations of Çatalhöyük by James Mellaart and collaborators in the early 1960s, it was in the 1990s, with the reopening of the site and the arrival of an international team headed by British archaeologist Ian Hodder, that its immediate relevance to the archaeological endeavour was explicitly spelled out. At this time, ‘presentation’ was itemised as one of three key components of the larger research project (Hodder 1996). Video, illustration, photography, physical reconstruction, museological display, site signage, and site tours, among other sensorially engaging tools, quickly became inseparably embedded into the methodological routine at the site. Although the potential for visualisation to prove truly transformative to the archaeological process was obviously there, to some extent ‘presentation’ seems primarily to have been construed – and then actualised – as a typical top-down showing of exhibitionary material to public audiences, as one-off (and often quite bizarre) performances, or as mere perfunctory recording of the archaeology. Moreover, many of the early visual efforts (e.g. video diaries) were mostly abandoned, and the teams working on different visual products and projects regularly saw high turnover and short lifespans from the 1990s through to the end of the 2000s (see Perry 2013a for further reflection on the dynamics at play here). Nevertheless, several of the site’s visualisers pressed forward with pictorial Crafting knowledge with (digital) visual media in archaeology 199 innovation, often in maverick fashion, playing around with different forms of visual interpretation and publishing short, reflective articles about their work. Often these outputs were relegated to the concluding chapters of their respective publications, or to separate books altogether (distinct from the primary data-reporting volumes), thereby reinforcing the well-worn tendency to appreciate visualisation as a terminal and disparate exercise, disconnected from active archaeological practice and interpretation (e.g. Hodder 2000). Even so, in these publications we still see experts, like Çatalhöyük’s first illustrator John Swogger, explicitly articulating the epistemological promise of engaged forms of visual production:

. . . there are hundreds of . . . interpretative ‘stories’ floating around the site – in the labs, on the veranda, in the dining room, or around the campfire. All archaeologists are familiar with such stories, but they are interpretations so fleeting that it is difficult to record them. Illustration can act as an interpreter for these stories, translating the ephemeral narrative of verbal interpretation and capturing it in the form of an image . . . So just as the fragmented archaeological conversations in labs and around the bar are proving-grounds for interpretations, so too are these reconstructions. They are photographs of suggestions – snapshots of ideas . . . their purpose lying in their ability to record the process of creating archaeological knowledge. (Swogger 2000a: 131)

Elsewhere Swogger goes further, advocating for the formulation of a real methodology for, and repeated field-testing of, experiments in archaeological illustration. These would include quick, temporary, and frequent sketches, or ‘instant reconstructions’, that visualize fleeting ideas, off-the-cuff remarks, and common assumptions on site which, in their immediacy and tangible realisation, necessarily compel ‘a much closer examination of the interpretation’ (Swogger 2000b: 147). He also champions the use of illustration as an everyday tool to facilitate communication among different specialists (especially lab-based workers and excavators who might otherwise speak at cross-purposes or not speak at all). He suggests that, indeed, everyone (no matter their expertise) should be encouraged to produce visual material. Underlying Swogger’s recommendations is the rationale that participation in the illustrative process demonstrates its complexities (e.g. the associated research, detail, composition, emphasis, etc.), which in turn forces attention towards ‘the entire web of logic which creates the style, content and presentation of a reconstruction’ (Swogger 2000b: 149). As I understand it, then, active involvement in illustrative practice makes palpable the very nature of interpretation itself. Unfortunately, there is no evidence that Swogger’s method has ever been seriously engaged, despite the fact that the Çatalhöyük project team clearly appreciates the intellectual potential of visualisation tools (e.g. Hodder 1997). Digital media – e.g. virtual reality (VR) – are also recognised for their obvious 200 Sara Perry promise to create ‘deeper understanding’ of the phenomenology of the site (Hodder 1997: 698). Yet, in typical fashion, various VR producers at the site have under- mined their own efforts by insinuating that virtual reality

will merely reinforce our own preconceived ideas about the world as it was then. One or two archaeologists have meanwhile complained that the computer reconstruction has become so lodged in their minds that they no longer actually see what is in front of them when excavating the site. Computer animation, if permanently in use, leads to the disappearance of historical time and space. The sensation proper to archaeology, the historical dimension, atrophies. (Emele 2000: 225; emphasis mine)

My concern is that these claims are myopic in that they ignore the fact that all forms of interpretation (whether digital or analogue, based in pictures, words, or otherwise) might be subject to the same critique. Moreover, the process of visualisation is arguably one of the few means available of genuinely actualising some of the ‘sensation’ of archaeology. It entails thinking, making, meaning- creation, through the strokes of the pencil, brush, digital stylus, etc. In more recent years, teams well versed in the application of digital media to archaeology have joined – and then left – Çatalhöyük (e.g. BACH - Berkeley Archaeologists at Çatalhöyük). The extent to which their contributions have facili- tated wider epistemic change across the site and amongst its many disparate specialists is still to be fully debated and untangled (although see Tringham and Stevanovi´c 2012). Outside the excavations, however, BACH team member Colleen Morgan (2009) has published a now cornerstone text on the epistemo- logical productivity of VR as realised through the online virtual world Second Life. Here she demonstrates creativity, technical skill, and automation intersecting through digital modelling, each contributing to a rethinking of the archaeological record and our understandings of everyday life in the past. To give one example, modelling in this programmed environment required Morgan to think about what times of day certain activities (e.g. wall plastering) could transpire, both in the online and offline Çatalhöyük worlds (i.e. in Second Life, where the sun sets promptly according to seasonal hours, and outside it, at Çatalhöyük during its actual occupation), or how houses in the settlement would be identifiable from one another. Perhaps more interestingly, as I see it, Morgan speaks directly of the craft involved in the digital reconstruction process and the manner by which working through the medium led to reconsiderations of the archaeological record. For instance, she describes the rudimentary visual functionality of Second Life, which demanded use of hard lines to represent features (e.g. ovens) that, in reality, would have been rounded. What is important is that the limitations of the technology arguably drove forward critical thinking about processes of making and the craft (or constituent dimensions) of the features themselves: Crafting knowledge with (digital) visual media in archaeology 201

This requires the archaeologist to approach artifacts, architecture, and the landscape from a different perspective; one that requires an additive, accretive process, breaking down the object into component parts instead of viewing excavated materials as a whole. (Morgan 2009: 476)

As I read it, the resistances of media compelled Morgan to reflect on the actual archaeology – both its virtual and material dynamics. It also led her to seek out new and more robust visual outputs to help structure her modelling work. In other words, her work exposed, for example, the lack of adequate extant photographs to enable close looking at the site. Despite an estimated 50,000 photos available to her, many failed to capture the texture, detail, and angle of features necessary for their full digital visualisation. She then found her own field practice altered as she began to take further photographs, better tailored to the aims of archaeological reconstruction. What strikes me as most profound is the possibility that the rote and hollow nature of some ubiquitous methods in archaeology (e.g. aspects of photography) is made obvious when one begins to engage in the craft of illustration. Here, following traditional modes of practice, the illustrator begins to replicate existing imagery to build new pictures out of it, and in so doing, discovers that the extant material was lacking sufficient detail to enable actual careful observation of the site. The implication, I suggest, is that if we better engage with digital and hand-based illustration, the expert skillsets and datasets involved in field practice itself have the potential to be pushed forward, made more rigorous, rich, and intellectually meaningful. So the craft of (digital) visual practice can directly fold in upon the craft of excavation practice, and vice versa:

making these interpretive decisions while recreating the room interior challenged my perceptions of the site, and made me truly engage with some of the questions that as an excavator I had pondered only in passing while filling out my data sheets. (Morgan 2009: 476)

Current work at Çatalhöyük, by the illustrator Kathryn Killackey and 3D graphic artist Grant Cox, further testify not only to the skill and epistemic promise of visual reconstruction (particularly in digital form), but also to its collaborative dimensions and inherent collegiality. Here the craft, embodiment, and intuitiveness at the core of the process of visualisation – and its relationship to understanding the actual creation of the archaeological record itself – are overt. As Cox notes in discussing his 3D modelling of Çatalhöyük’s experimental house (Figure 11.2),

I just got this feeling that the material they used had been applied quite liberally across the surfaces in a way that it finished everything and the walls kind of merged into the floor, the floor merged into the ceiling, the platforms 202 Sara Perry

merged into the walls and the floor and so the core of the house was kind of one entity almost because they covered everything with the same sort of finish. And that was something I picked up on, I thought that was the first thing I wanted to try and get in the model. (Grant Cox, interview with the author and team, 2012; emphases mine)

Echoing what I see as the nature of general archaeological interpretation overall, he goes on to explain the digital composition of the image as a reflexive dialogue between ambience, performance, and evidence, and he hints at the detail that close looking at different datasets facilitated for his visualisation (Figures 11.2 and 11.3):

there is a lot of contrast because you have the light from the outside world, but it’s cutting you off from knowing what’s out there. But then you also have the darkness of the corner of the room, you don’t really know what’s going on over there, but it’s also, I can imagine for example going up the ladder and hearing noise, while going down the ladder it is more quiet and reserved. That was kind of the general feel that I have of the space . . . [Look at] the wear on the corners of the platforms, here you can see one on the corner and also around the oven. There is a lot of wicker material scattered throughout the scene, so if you look up close you can see a lot of small bits of wicker just everywhere, all over the floor, just shards of it lying around, some clay balls, also the smoke from the oven on the back wall, the discoloration, and also up here where they plastered in the wooden ceiling beams . . . (Grant Cox, interview with the author and team, 2012)

FIGURE 11.2 3D model merging the existing experimental house with the ‘hunting shrine’ (uncovered by James Mellaart’s team) at Çatalhöyük, view looking south-east. (Image by Grant Cox; courtesy of Artas Media.) Crafting knowledge with (digital) visual media in archaeology 203

FIGURE 11.3 3D model merging the existing experimental house with the ‘hunting shrine’ (uncovered by James Mellaart’s team) at Çatalhöyük, view looking north. (Image by Grant Cox; courtesy of Artas Media.)

Cox is clear about the artistry behind his ‘physically accurate’ rendering work, further spotlighting the seeming futility of accounting exercises like the London Charter:

you can build something in a rendering engine but then apply post production to it and theoretically and practically it has been produced in a physically accurate rendering engine but you can make two images that are completely different . . . two physically accurate models can use different mathematics and produce different images and retracing algorithms and the materials are going to be different . . . and however you develop the materials is going to be different . . . I don’t really see how you can use physical accuracy on its own as a way to justify what you’ve done without explaining the artistic elements. So it feels when I read through people’s work who have done that, that they are doing it because they want it to come across as a scientific experiment, when most of the decisions about a model are subjective. (Grant Cox, interview with the author and team, 2012)

Perhaps most interestingly, he emphasises the labour and commitment behind his visual work – a kind of labour that would equally resonate with most site excavators:

I mean this is my house. You know, it is their house, but it is my house too. I built this house. (Grant Cox, interview with the author and team, 2012) 204 Sara Perry

Cox’s reflections are complemented by those of Çatalhöyük illustrator Killackey, who expounds on the necessarily participatory and inclusive character of her visual production (Figure 11.4):

I started with this plan that Camilla [Mazzucato] made based on two buildings from the lower excavations, the Hunting Shrine and the Vulture Shrine, and these layouts are a little bit iffy because they are based on Mellaart’s very little drawings in Anatolian Studies and we were also looking at photos that Mellaart had taken and going back and forward a bit to decide some of the features and I think she [Camilla] also did a little bit of adjustment down here . . . And then Building 77 is from the recent excavations so it is very well planned and then this building, Ian [Hodder] kind of decided he wanted a composite of the different elements . . . (Kathryn Killackey, in interview with the author and team, 2012)

The illustrator plays the intermediary at multiple levels, and is directly implicated in the probative intellectual enterprise of archaeology. As Killackey describes it, ‘some people just have not thought through things specifically, so I come back

Deneysel Evler ia Düzen Planı (experimental houses cutaway)

Tipik Çatalhöyük Evi (composite building)

77 No’lu Ev (Building 77)

Av Sahneli Kutsal Yapı (hunting shrine)

Akbabalı Kutsal Yapısı (vulture shrine)

012m

FIGURE 11.4 Plan of a proposed new experimental house at Çatalhöyük constructed from the visual outputs of multiple practitioners – past and present – from the site. (Image courtesy of Katy Killackey.) Crafting knowledge with (digital) visual media in archaeology 205 with questions and they don’t know or they haven’t gotten to that place in their head’. She goes on to explain the extensive academic debate that ensued when Çatalhöyük’s multiple team leads were involved in an exercise of producing a generalised landscape view of the site – an exercise that literally actualised the many complexities of ‘assembling Çatalhöyük’. What is partially at stake here again is the ‘good citizenship’ that Sennett (2008) describes – for visualisation stands at the nexus of negotiation and interaction with many people and materials over time and space, all coming together in a form of creative compromise. Killackey’s skill – her style and ability to capture a mood, her adeptness with and attention to the archaeological record, and with cooperating with varied specialists – means that she has a unique insight into the nature of archaeology itself: its soul, its data, its practitioners, and its weaving of these components into disciplinary knowledge. This is the power of visual production (whether digital or not).

The archaeological eye Nordbladh (2007: 111) has said about imagery in the history of archaeology that it can ‘tell us primarily what was sought and what good it could give rise to, for individuals or society’. I would like, however, to stretch this argument further and suggest that the interrogation of the visual – and active participation in its creation – might illuminate not only what has been sought in archaeology, and with what consequences for the field, but the very means by which knowledge is negotiated, constructed, and aligned with existing structures of expertise. Complemented by other non-representational examinations of archaeological practice (see Perry 2013b, for example, and other contributions in Alberti et al. 2013), such work can (literally) draw us into critical discussion about normative architectures of power and discipline in archaeology: where and why archaeology as an institutionalised pursuit has emerged; which bodies, markets, and materials have been implicated and combined in the meaning-making process; what has been born of or made possible by the mobilisation of visualisation; and how, in an archaeological sense, (visual) artefacts can be understood to transform the world. Thinking through Mitchell (2008), then, I see visualisation (and visual media) as engaged social practice, rather than some kind of solid, deconstructible object or one-way communicative device. As such, it is caught up in the flows and fabrics of all forms of everyday action, and thus is not reducible to simplistic categories of ‘public’ versus ‘academic’, or to singular items detached from modes of seeing and economies of circulation and consumption. Shanks and Webmoor (2013) are similarly concerned to move past traditional semiotic and communicational models of visual representation, towards the broader ‘work’ of graphic media in archaeology. In so doing, however, they advocate for a more loosely defined concept of ‘mediation’, which does not privilege optics as much as varied, multi-sensory, multi-modal articulations and negotiations of people and things – that is, all of those forms of labour and administration (the artefacts, relationships, ideologies, systems, and instrumentation) that make archaeology 206 Sara Perry possible as a disciplinary pursuit. While I appreciate the fluidity of these negotiat- ions, I am concerned that they may be rendered meaningless (and that the notion of ‘mediation’ is potentially destined to become a shapeless, catchall term) if we are to overlook the particularities of human-material practices. I think there is a need, then, to continue to interrogate specific visual engagements. This is not to give vision an undue primacy, but to make clear, as Grasseni (2007) also aims to demonstrate, that it makes a difference, both personally and professionally, to be skilled to see (and produce) in one manner as opposed to another. Willerslev (2007) contends that vision and visual forms have effectively – and unreflectively – been rejected as legitimate topics of study owing to caustic critiques of the often objectifying, detached, dehumanising impacts of the anthropological gaze (in the vein of Fabian 1983, for example). However, following Willerslev, I argue that in attending to the micro-scale, ground-level movements and realisations of the visual, we have an opportunity both to expose the very taken-for-granted operations that allow ocularcentrism to manifest itself in the first place, and hence to sensitively rethink the knowledge and objects that vision can enable and assist/resist. As Herzfeld (2007: 214) has said in regard to anthropology, disciplinary training ‘is often inchoate; explicit instruction raises suspicions of betrayal of craft’. Within archaeology, such training is seemingly very explicit and rigorous, and yet tends to be achieved through unspoken, embedded apprenticeship in certain ways of seeing and articulating sights/sites. These processes are described by Grasseni (2007) as ‘skilled visions’, and, as she explains it, to ‘exercise skilled vision means to belong socially in communities and networks that share aesthetic sensibilities, principles of good practice, rituals of participation, processes of apprenticeship, ideological stances and political interests’. So too, I think, does the command of such skilled vision in fact make possible these very communities and networks. As evidenced at Çatalhöyük, this is the potent role that (digital and analogue) visualisation plays in archaeology. It enables forms of thinking and practice that are potentially revolutionary for the discipline. But, more broadly, it has a stake in transforming the world at large – in building a more democratic, inclusive, critically engaged, and truly reflexive network of media and people.

Acknowledgements Many thanks go to Stephanie Moser, Graeme Earl, Colleen Morgan, John Swogger, Katy Killackey, Grant Cox (Artas Media), Alice Watterson, Jennie Anderson, Ian Kirkpatrick, and the various members of the Çatalhöyük visualisa- tion team, with whom I have collaborated since 2009. They have variously read my words, critiqued my ideas, contributed invaluable information to my research, and continue to inspire me. Crafting knowledge with (digital) visual media in archaeology 207

References Aitchison, K. (2011) ‘Survey of archaeological specialists 2010–11’, prepared for the Higher Education Academy and English Heritage, report reference HEH06, 12 April. Alberti, B., Jones, A. M. and Pollard, J. (2013) Archaeology After Interpretation: Returning Materials to Archaeological Theory, Walnut Creek, CA: Left Coast Press. Auslander, L., Bentley, A., Halevi, L., Sibum, H. O. and Witmore, C. (2009) ‘AHR conversation: historians and the study of material culture’, American Historical Review, 114: 1,355–1,404. Beacham, R. C. (2012) ‘Defining our terms in heritage visualization’, in A. Bentkowska- Kafel, H. Denard, and D. Baker (eds) Paradata and Transparency in Virtual Heritage, Farnham: Ashgate, 7–11. Bentkowska-Kafel, A. (2013) ‘“I bought a piece of Roman furniture on the Internet. It’s quite good but low on polygons.” Digital visualization of cultural heritage and its scholarly value in art history’, Visual Resources, 29: 38–46. Bentkowska-Kafel, A. and Denard, H. (2012) ‘Introduction’, in A. Bentkowska-Kafel, H. Denard and D. Baker (eds) Paradata and Transparency in Virtual Heritage, Farnham: Ashgate, 1–4. Bentkowska-Kafel, A., Denard, H. and Baker, D. (eds) (2012) Paradata and Transparency in Virtual Heritage, Farnham: Ashgate. Cain, V. E. M. (2010) ‘“The direct medium of the vision”: visual education, virtual witnessing and the prehistoric past at the American Museum of Natural History, 1890–1923’, Journal of Visual Culture, 9: 284–303. Cain, V. E. M. (2011) ‘The art of authority: exhibits, exhibit-makers, and the contest for scientific status in the American Museum of Natural History, 1920–1940’, Science in Context, 24: 215–238. Cain, V. E. M. (2012) ‘The craftsmanship aesthetic: showing making at the American Museum of Natural History, 1910–45’, Journal of Modern Craft, 5: 25–50. Caraher, W., Kourelis, K. and Reinhard, A. (eds) (in press) Punk Archaeology, Grand Forks, ND: University of North Dakota Digital Press. Clack, T. and Brittain, M. (eds) (2007) Archaeology and the Media, Walnut Creek, CA: Left Coast Press. Crawford, O. G. S. (1936) ‘Archaeological photography’, Antiquity, 10: 351–352. Earl, G. (2013) ‘Modeling in archaeology: computer graphic and other digital pasts’, Perspectives on Science, 21: 226–244. Edgerton, S. (1991) The Heritage of Giotto’s Geometry: Art and Science on the Eve of the Scientific Revolution, Ithaca, NY: Cornell University Press. Emele, M. (2000) ‘Virtual spaces, atomic pig-bones and miscellaneous goddesses’, in I. Hodder (ed.) Towards Reflexive Method in Archaeology: The Example at Çatalhöyük, Cambridge: McDonald Institute for Archaeological Research, 219–227. Fabian, J. (1983) Time and the Other: How Anthropology Makes its Object, New York: Columbia University Press. Field, J. V. (1997) The Invention of Infinity: Mathematics and Art in the Renaissance, Oxford: Oxford University Press. Gibbons, G. (2011) ‘2010 VIA Snapshot Survey summary results’, report to English Heritage (HEEP Project No. 5172MAIN), Southampton: Visualisation in Archaeology. Goodwin, C. (1994) ‘Professional vision’, American Anthropologist, 96: 606–633. Grasseni, C. (2007) ‘Introduction’, in C. Grasseni (ed.) Skilled Visions: Between Apprenticeship and Standards, New York: Berghahn, 1–20. Gurche, J. (2013) Shaping Humanity: How Science, Art, and Imagination Help Us Understand our Origins, New Haven, CT: Yale University Press. 208 Sara Perry

Hattwig, D., Burgess, J., Bussert, K. and Medaille, A. (2011) ‘ACRL Visual Literacy Competency Standards for Higher Education’. Available at: www.ala.org/acrl/standards/ visualliteracy; accessed 13 March 2014. Herzfeld, M. (2007) ‘Epilogue: envisioning skills: insight, hindsight, and second sight’, in C. Grasseni (ed.) Skilled Visions: Between Apprenticeship and Standards, New York: Berghahn, 207–218. Hodder, I. (1996) ‘Re-opening Çatalhöyük’, in I. Hodder (ed.) On the Surface: Çatalhöyük 1993–95, Cambridge: McDonald Institute for Archaeological Research, 1–18. Hodder, I. (1997) ‘“Always momentary, fluid and flexible”: towards a reflexive excavation methodology’, Antiquity, 71: 691–700. Hodder, I. (ed.) (2000) Towards Reflexive Method in Archaeology: The Example at Çatalhöyük, Cambridge: McDonald Institute for Archaeological Research. Hodgson, J. (2001) ‘Archaeological reconstruction: illustrating the past’, IFA Paper No. 5, Reading: Association of Archaeological Illustrators and Surveyors, Institute of Field Archaeologists. Hodgson, J. (2008) Special Interest Group for Illustration and Survey: Survey 2008 Report, Reading: Institute for Archaeologists. Hogarth, P. (1986) The Artist as Reporter, London: Gordon Fraser. Hunter, M. C. (2013) Wicked Intelligence: Visual Art and the Science of Experiment in Restoration London, Chicago, IL: University of Chicago Press. Ingold, T. (2011) ‘Introduction’, in T. Ingold (ed.) Redrawing Anthropology: Materials, Movements, Lines, Farnham: Ashgate, 1–20. Ingold, T. (2013) Making: Anthropology, Archaeology, Art and Architecture, London: Routledge. James, S. (in prep) ‘Visually challenged: stunted “visual competence” in archaeology, a problem hiding in plain sight’. Jordanova, L. (2012) The Look of the Past: Visual and Material Evidence in Historical Practice, Cambridge: Cambridge University Press. Knight, C. (1946) Life through the Ages, Bloomington: Indiana University Press. Kolijn, E. (2013) ‘Observation and visualization: reflections on the relationship between science, visual arts, and the evolution of the scientific image’, Antonie van Leeuwenhoek, 104: 597–608. Lave, J. (1997) ‘The culture of acquisition and the practice of understanding’, in D. Kirshner and J. A. Whitson (eds) Situated Cognition: Social, Semiotic, and Psychological Perspectives, Mahwah, NJ: Lawrence Erlbaum Associates, 17–35. Llobera, M. (2011) ‘Archaeological visualization: towards an archaeological information science (AISc)’, Journal of Archaeological Method and Theory, 18: 193–223. Lopes, D. M. (2009) ‘Drawing in a social science: lithic illustration’, Perspectives on Science, 17: 5–25. Manovich, L. (2013) Software Takes Command, New York: Bloomsbury. Mitchell, W. J. T. (2008) ‘Addressing media’, MediaTropes, 1: 1–18. Morgan, C. L. (2009) ‘(Re)Building Çatalhöyük: changing virtual reality in archaeology’, Archaeologies: Journal of the World Archaeological Congress, 5: 468–487. Morgan, C. L. (2012) ‘Emancipatory digital archaeology’, unpublished PhD dissertation, University of California, Berkeley, CA. Moser, S. (1992) ‘The visual language of archaeology: a case-study of the Neanderthals’, Antiquity, 66: 831–844. Moser, S. (2014) ‘Making expert knowledge through the image: connections between antiquarian and early modern scientific illustration’, Isis, 105(1): 58–99. Nickelsen, K. (2006) ‘Draughtsmen, botanists and nature: constructing eighteenth-century botanical illustrations’, Studies in History and Philosophy of Biological and Biomedical Sciences, 37: 1–25. Crafting knowledge with (digital) visual media in archaeology 209

Nordbladh, J. (2007) ‘Observing the past and making it visible: on the use of images in archaeology’, in L. Johannesson, U. Eliasson, P. Hallberg and B. Karlsson (eds) Images in Arts and Sciences, selected papers from a conference held by the Royal Society of Arts and Sciences in Göteborg, 13–14 October 2004, Göteborg: Acta Regiae Societatis Scientiarum et Litterarum Gothoburgensis. Interdisciplinaria. Kungl, Vetenskapsoch vitterhets- samhället, Göteborg, 9: 111–126. Opitz, R., Terrenato, N., Meliconi, I. and Frischer, B. (2013) ‘21st c. data, 21st c. publications. 3D model publication and building the peer reviewer community’, workshop held at the Digital Heritage Congress, Marseille, France, 28 October – 1 November. Perry, S. (2010) ‘Reflections on the second VIA workshop at the University of Southampton, 22–23 October 2009’, report to English Heritage (HEEP Project No. 5172MAIN), Southampton: Visualisation in Archaeology. Perry, S. (2013a) ‘Debating the legacy of postprocessualism: visual reflexivity at Çatalhöyük, Turkey’, paper presented at the Institute for Archaeologists Annual Conference, Birmingham, UK, 17–19 April. Perry, S. (2013b) ‘Archaeological visualization and the manifestation of the discipline: model-making at the Institute of Archaeology, London’, in B. Alberti, A. M. Jones and J. Pollard (eds) Archaeology After Interpretation: Returning Materials to Archaeological Theory, Walnut Creek, CA: Left Coast Press, 281–303. Perry, S. and Johnson, M. (2014) ‘Reconstruction art and disciplinary practice: Alan Sorrell and the negotiation of the archaeological record’, Antiquaries Journal, 94: in press. Petrie, W. M. F. (1904) Methods and Aims in Archaeology, London: Macmillan and Co. Piccini, A. (1996) ‘Filming through the mists of time – Celtic constructions and the documentary’, Current Anthropology, 37: S87–S111. Piggott, S. (1965) ‘Archaeological draughtsmanship: principles and practice, part I: principles and retrospect’, Antiquity, 39: 165–176. Piggott, S. (1978) Antiquity Depicted: Aspects of Archaeological Illustration, London: Thames and Hudson. Prak, M. (2013) ‘An artisan “revolution” in late medieval and early modern Europe?’. Available at: http://dutchworld.columbia.edu/files/2013/08/ArtisansRevolution_2013- 1-1.pdf; accessed 16 March 2014. QAA (2007) ‘Subject benchmark statement: archaeology’, Quality Assurance Agency. Available at: www.qaa.ac.uk/Publications/InformationAndGuidance/Documents/ Archaeology.pdf; accessed 13 March 2014. Read, R. and Smith, G. (2009) ‘Training the undervalued and unacknowledged: specialist training provision for archaeological illustrators in the UK’, paper presented at the second Visualisation in Archaeology workshop, Southampton, UK, 22–23 October. Russell, I. A. and Cochrane, A. (eds) (2014) Art and Archaeology: Collaborations, Conversations, Criticisms, New York: Springer. Rutot, A. (1919) ‘Un essai de reconstitution plastique des races humaines primitives’, Mémoire de l’Académie royale de Belgique, Classe des beaux-arts, 2(1). Sapsed, J. and Tschang, F. T. (2014) ‘Art is long, innovation is short: lessons from the Renaissance and the digital age’, Technological Forecasting and Social Change, 83: 127–141. Schablitsky, J. M. and Hetherington, N. J. (2012) ‘Archaeology on the screen’, in M. Rockman and J. Flatman (eds) Archaeology in Society: Its Relevance in the Modern World, New York: Springer, 139–152. Selkirk, A. and Selkirk, W. (1973) ‘Archaeologists have no soul’, Current Archaeology, 41: 163. Sennett, R. (2008) The Craftsman, London: Penguin. Shanks, M. (2012) The Archaeological Imagination, Walnut Creek, CA: Left Coast Press. 210 Sara Perry

Shanks, M. and McGuire, R. (1996) ‘The craft of archaeology’, American Antiquity, 61: 75–88. Shanks, M. and Tilley, C. (1987) Re-Constructing Archaeology: Theory and Practice, Cambridge: Cambridge University Press. Shanks, M. and Webmoor, T. (2013) ‘A political economy of visual media in archaeology’, in S. Bonde and S. Houston (eds) Re-presenting the Past: Archaeology Through Image and Text, Providence, RI: Brown University, 85–108. Smith, P. H. (2004) The Body of the Artisan: Art and Experience in the Scientific Revolution, Chicago: University of Chicago Press. Smith, P. H. (2013) ‘Making things: techniques and books in Early Modern Europe’, in P. Findlen (ed.) Early Modern Things: Objects and their Histories, 1500–1800, London: Routledge, 173–203. Sorrell, A. (1973) ‘The artist and reconstruction’, Current Archaeology, 41: 177–181. Swogger, J. G. (2000a) ‘Archaeological illustration’, Çatalhöyük 2000 Archive Report, Cambridge: Catalhoyuk Research Project, 129–131. Swogger, J. G. (2000b) ‘Image and interpretation: the tyranny of representation?’, in I. Hodder (ed.) Towards Reflexive Method in Archaeology: The Example at Çatalhöyük, Cambridge: McDonald Institute for Archaeological Research, 143–152. Swogger, J. G. (2014) John Swogger illustration. Available at: http://johngswogger.word press.com/; accessed 13 March 2014. Tringham, R. and Stevanovi´c, M. (2012) Last House on the Hill: BACH Area Reports from Çatalhöyük, Turkey, Los Angeles, CA: Costen Institute of Archaeology Press. Watterson, A. (2014) ‘Digital dirt virtual pasts: blogging the tears and triumphs of an archaeological illustrator’. Available at: http://digitaldirtvirtualpasts.wordpress.com/; accessed 13 March 2014. Weber, G. W. and Bookstein, F. L. (2011) Virtual Anthropology: A Guide to a New Interdisciplinary Field, New York: Springer. Wheeler, R. E. M. (1954) Archaeology from the Earth, Oxford: Clarendon Press. Wiber, M. (1997) Erect Men, Undulating Women: The Visual Imagery of Gender, ‘Race’ and Progress in Reconstructive Illustrations of Human Evolution, Waterloo, Ont.: Wilfrid Laurier University Press. Willerslev, R. (2007) ‘“To have the world at a distance”: reconsidering the significance of vision for social anthropology’, in C. Grasseni (ed.) Skilled Visions: Between Apprenticeship and Standards, New York: Berghahn, 23–46. PART III Multiple working hypotheses, strategies of elimination, and triangulation This page intentionally left blank 12 UNCERTAIN ON PRINCIPLE

Combining lines of archaeological evidence to create chronologies

Alex Bayliss and Alasdair Whittle

The world of ‘absolute, true, and mathematical time’, which ‘of itself, flows equably, without reference to anything external’ (Newton 1687), was finally killed in the trenches of the First World War, whence Schwarzschild (1916) proposed the first solution to Einstein’s general relativity field equations. From now on, space and time were inextricably linked; the past, the future, and simultaneity became relative and dependent on the reference frame of the observer (Einstein 1905; Minkowski 1908); and space-time itself became dynamic, bent by gravity relative to the position of the observer, and deformed by mass and speed (Einstein 1915). Space and time are now not merely the arena in which the drama of the universe is acted out, but part of the cast. Into this cosmos, we must place the material remnants of earthlings. But the time of physics seems to play no role at all in our time experience (Dieks 2006: 167). Instead, our time appears local and is based on the proper time (τ) of the world-line between perception events, not on cosmic time (t). Our world-line is that of Planet Earth, and so our time runs at a rate determined by the strength of its gravity and the speed of its rotation and orbit. This world-line and this time are part of the shared lived experience of both us present earthlings and those past earthlings whose material traces we study. This allows us, like Finnegan, to experience happenings ‘seemaultaneously’ (Joyce 1939: 161). We inhabit an inter- subjectively experienced specious present, which (contra Olivier 2001: 70) is not a single temporality, but a multi-temporal palimpsest of its pasts and its presents. Its depth is problem-determined (Bailey 1983: 179) – micro-seconds for quantum mechanics, light years for cosmic processes, and somewhere within the range of one-twenty-fifth of a second (approximately the shortest interval of human percep- tion) and a lifetime for human consciousness. Here, again, we have something in common with the people of the past: the limitations of human biology. 214 Alex Bayliss and Alasdair Whittle

Fortunately, we also share this common, if dynamic, planetary reference frame with the materials we study. The lines of evidence we use to construct chronologies are also earth-based, sharing the planet’s gravity (stratigraphy), climate (environ- mental records such as pollen and dendrochronology), geomagnetism, and cosmic ray flux (radiocarbon, magnetic, and luminescence dating). The cosmogenic production and nuclear decay of radiocarbon, at the heart of so many of our chronologies, inescapably force us into ‘the stochastic world of capricious electrons’ (Thiher 2005: 57). This is a probabilistic world where the chance element is inherent in nature and not merely imposed by our limited grasp of the system (Heisenberg 1958). It is also a world that is relational and discontinuous. Solid, material objects dissolve into wave-like patterns of probabilities – probabilities not of things, but of observed interactions – and are inescapably modified by the act of measurement (Einstein et al. 1935). And there are long-range correlations between particles that simply cannot be visualised in a continuous space-time, and suggest that the nature of the universe is interpenetrative in a way that allows events to be linked without time or causality (Miller 1990). These intrinsic properties of the lines of evidence we utilise in the creation of our chronologies have profound implications for the properties of those chronolo- gies. Clearly, they are earth-based; we are estimating the proper time (τ) of the past events we wish to date along the world-line of our planet. We cannot pretend that there is anything ‘absolute’ about these chronologies. They are also uncertain: not only in terms of the epistemological measurement uncertainty on our date estimates, but also in terms of the ontological indeterminacy that is an inherent characteristic of our observables. Our interpretative frameworks must acknowledge these properties.

Lines of evidence How then do archaeologists routinely construct their understanding of the chronological dimensions of past human lives? We want briefly to highlight five lines of evidence: stratigraphy; chrono-typology; seriation; spatial associations; and scientific dating. People tend to repeat themselves, and they often do so in the same places. In the end, much archaeological analysis is based on human activity concentrated in particular places, in what are often called simply sites. Although the site is in fact an enormously varied concept, much of our understanding of site chronology is based on stratigraphy. Setting aside horizontal relationships, it is natural and anthropogenic deposition, acting under the force of gravity, that produces the vertical succession that enables archaeologists to infer the passage of time and the operation of change. The use of stratigraphy is not confined to what we call sites, since the vertical succession in any given spot in a peat bog, for example, randomly chosen for pollen analysis, is similarly exploited for constructing chronological understanding. Interruptions to, or deviations from, the perception of regular succession allow us to detect recuts, occasional inversion, and later cuts. Field skill Uncertain on principle: combining lines of evidence to create chronologies 215 is crucial. At its best the discipline can be expert in its dissection of complex layers and contexts, but at its worst, much confusion and conflation can, and does, result; in both scenarios, understanding comes from the chain of observation and inter- pretation necessarily involved in excavation (Hodder 1997). People and things are inextricably entwined (Hodder 2012; Jones 2012). The rich materiality of past lives provides the basis for its chronological analysis by typology and seriation, exploiting principles of both similarity and difference. Chrono-typology works by identifying stylistic similarities across a background of difference, using also association, context, and stratigraphy (Adams and Adams 1991). Seriation, often undertaken using correspondence analysis, formally orders a matrix of types and units (such as artefact types in graves or decorative motifs on pots). Such orderings may or may not be chronologically successive, although in practice they can often be shown to be so (Greenacre 2007). With both chrono- typology and seriation, the perceptions of the material by the analyst, the assumptions of the chosen methodology, and the objectives of the study are crucial. People live their lives in webs of connection and interaction with others. People move across taskscapes for all manner of needs and wants, and they live in worlds framed by others: neighbours, friends, kin, allies, enemies, the distant, and the unknown (Neustupny´ 1998; Gamble 1999). Sites can be and perhaps too often are studied in isolation, but the wider context to which any site belongs is spatial. The survival of the culture concept in European prehistory, for example, must be based on an implicit recognition of the spatiality of human culture. But just as the earth’s gravity pulls on material to produce stratigraphy, so distance across the planetary surface tugs at relationships and networks; again, assumptions of similarity and difference are central. Whether and where to draw lines in or across the ‘unbroken landscapes of variation’ (Ingold 1996: 117), where to infer cores and where peripheries, what the nature of boundaries may be, and how to interpret changes, sometimes bafflingly abrupt, in material configurations, are all highly complex questions, again centrally involving the viewpoint of the observer. Culture history has earned some of its poor reputation through its inability or reluctance to embrace a wider set of interpretations than bounded, internally homogenous groups, and invasions and migrations; while more recent studies of materiality have not been of sufficiently broad in scale to re-energise the study of spatiality. It is rare now for studies of sites, landscapes, and material not to be accompanied in some fashion by the application of scientific dating. In part, context obviously determines use of method; in European prehistory, for example, dendrochronology is most concentrated in the favourable conditions of preservation in the Alpine foreland (Menotti and O’Sullivan 2013). The varying modern context can also strongly influence the number of scientific dates affordable by a given project. Acceptance of different scales of chronological resolution is also significant. The high-resolution and often astonishingly brief site biographies of the Alpine foreland (for example, the fifteen-year history of Arbon Bleiche 3; see Jacomet et al. 2004) do not appear to have much impact on the wider interpretation of Neolithic existence; beyond the Alpine foreland, and not just in Europe, radiocarbon dating 216 Alex Bayliss and Alasdair Whittle is central to most prehistoric (and some other) chronologies. Here again, even armed with modern scientific dating methods, the archaeological interpreter is central. Disappointingly, the discipline has taken a long time to realise the absolutely central importance of short-life samples of known taphonomy, which date, without age offsets, their contexts (Waterbolk 1971). Routinely, radiocarbon dates have been interpreted by visual inspection, or through summing of calibrated dates, both of which can be shown to produce inaccurate chronologies of exaggerated duration (Bayliss et al. 2007a). Normally, prehistorians seem content with radiocarbon-backed culture history chronologies, which employ successive units of 200 years’ duration or more; the familiar chronological tables charting these units (e.g. Figure 12.1) normally also employ a lot of horizontal straight lines at the same points to denote spatial relationships. It is no surprise that slow change over the long term has been the dominant, if not the unthinking, default chronological perspective.

Central Balkan Central Bosnia Dalmatia

SCHIER 1997 BENAC 1973 a/b GIMBUTAS 1974 STERUD/STERUD PERIĆ 1995 HOFMANN BENAC 1979 cal BC OROSS 2007 BENAC 1979 1974 in Press MÜLLER 1994 cal BC BORIĆ 2009

4300 4300 Butmir 4/ 4400 Butmir IVb Donje 4400 Moštre 4500 4500 Butmir IIIc Late Butmir c Butmir IIIb Late Butmir b Butmir 3b 4600 Vinča D2 4600 Butmir III Hvar III (late) Butmir IVa Butmir IIIa Late Butmir a Butmir 3a 4700 Vinča C3/D1 4700

4800 Butmir IIc Early Butmir c Butmir 2b 4800 Vinča C2 Butmir II Butmir IIb Early Butmir b Butmir III Hvar II 4900 (classical) 4900 č Butmir 2a Vin a C1 Butmir IIa Early Butmir a 5000 5000 Vinča B2 Hvar I/ Butmir Ib Butmir 1b Butmir I Danilo IV Vinča B1b č 5100 (early) Late Star evo d Butmir II 5100 Vinča B1a Butmir Ia Butmir 1a Danilo III 5200 5200 Vinča A3 Kakanj III Kakanj II Late Starčevo c (Proto)butmir 1b Danilo II 5300 Vinča A2 Kakanj II 5300 Kakanj I Late Starčevo b (Proto)butmir 1a Danilo I Vinča A1a,b Kakanj I 5400 5400

5500 5500 Late Starčevo/ Starčevo/ Late Starčevo Late Starčevo a Impresso C Starčevo Impresso Impresso 5600 5600

5700 5700

FIGURE 12.1 An example of an up-to-date culture-historically based regional chronol- ogy (redrawn from Müller et al. 2013: Abb. 35).

Combination Alison Wylie, drawing on several sources including Peirce, Bernstein, and Kuhn, has explored the strength and advantages of the notion that ‘scientific arguments Uncertain on principle: combining lines of evidence to create chronologies 217 are more like cables than chains’ (2002: 162); in the cable metaphor, individual lines of argument that are insufficient can be woven together to make a cumu- latively persuasive case (Wylie 2002: 162–163). She notes, however, that ‘strands that make up a cable of comparative, evaluative argument may conflict with one another’, requiring dynamic judgements and revisions (Wylie 2002: 163). In order to deal with different kinds and scales of argument, she adds another metaphor, that of tacking; according to this, ‘incommensurability between theories or worldviews is mediated (when and if it is) by a concatenation of cables of arguments, each woven in these multiple dimensions’ (Wylie 2002: 165). In the construction of archaeological chronologies, Bayesian statistics provide a formal and explicit methodology for weaving together different strands of evidence to form the cable. Calibrated radiocarbon dates, or other date estimates on the calendar scale, are combined with archaeological prior information of various kinds to produce a combined chronology that is stronger than its individual components (Buck et al. 1996; Bayliss and Bronk Ramsey 2004). The resultant cable should be both stronger (more reliable) and tighter (more precise). The basic idea behind the Bayesian approach to the interpretation of data is encapsulated by Bayes’s theorem. It is fundamentally probabilistic and contextual. It simply means that we analyse the new data we have collected about a problem in the context of our existing experience and knowledge about that problem. This enables us to arrive at a new understanding of the problem that incorporates both our previously existing knowledge and our new data. We do this by the use of formal probability theory, where all three elements of our model are expressed as probability density functions. These give us a quantitative measure of our state of knowledge of each component of the model. An accessible general introduction to the principles of Bayesian statistics is provided by Lindley (1985) and to its history by Bertsch McGrayne (2011). In producing archaeological chronologies, Bayesian statistics provide an explicit, probabilistic method for estimating the dates when events happened in the past and for quantifying the uncertainties of these estimated dates. The approach combines calibrated radiocarbon dates with knowledge of the archaeological contexts from which they are derived to produce a series of date estimates; stringent demands are made of both the radiocarbon dates, which must be chosen according to strict selection criteria (Bayliss 2009; Whittle et al. 2011: Chapter 2), and archaeological knowledge of stratigraphy, associations, sample taphonomy, and context in general. Procedures must also be adopted to account for inherent scatter in a group of related radiocarbon dates (Buck et al. 1992; Bayliss et al. 2007a; Weninger et al. 2010). The power of modern computers enables millions of calculations to be undertaken, in order to reconcile the probability distributions of the individual calibrated radiocarbon dates with the other available information using Markov Chain Monte Carlo methods. Each distribution is repeatedly sampled to build up a set of solutions consistent with the model structure. Statistical checks are available to check the stability of a solution (its convergence) and the compatibility of the radiocarbon dates and the archaeological information included in the model (its 218 Alex Bayliss and Alasdair Whittle agreement). All the models presented in this paper have been calculated using the program OxCal (v.3.10–4.1), and so details of the algorithms and statistics employed can be found in Bronk Ramsey (1995, 1998, 2001, 2009a, 2009b). These diagnostic statistical tools aid us in ensuring internal consistency within our cable (see also Wylie 2002: 176–177). Our models also incorporate archaeological prior beliefs of various kinds. We need to be clear about the basis of these beliefs. Some may be comparatively unequivocal – the stratigraphic succession of a sequence of articulated animal bone samples, for example. In other cases the stratigraphic sequence may be open to alternative interpretations (see below, Figures 12.12 and 12.13), or there may be doubt about the taphonomy of the dated material. In other situations, our prior beliefs may themselves derive from other methods, which bring with them their own structures and sets of presuppositions. For example, correspondence analysis will place a grave according to the average position of each of the artefact types it contains, and so in a chronological sequence it will be located at the average time when all the artefacts in the grave assemblage were manufactured, not at the point of burial. The fundamental point is that Bayesian statistics, being a formal methodology, force archaeologists to be explicit about their strands of reasoning. In our view many archaeological opinions are held on a qualitative scale. Some may be strongly held, based on internal coherence or with reference to some sense of independent criteria, others more tentatively, but the scale of belief is not readily quantifiable. We therefore do not currently incorporate degrees of ‘prior belief ’ into our chronological models since we believe that the quantification of a qualitative belief is always on an arbitrary scale. If they satisfy the required standards of the statistical tools available, choosing between alternative models is, in the end, a matter of archaeological judgement. Sometimes alternative models are incompatible with each other. In this case, at least some of our date estimates will be erroneous. Their inherent uncertainty has, however, been explicitly estimated using formal probabilistic statistical methods. This quantified statistical uncertainty should not be confused with the possibility that a given model outcome is in error. Internal checks militate against error, though error is still possible if laboratories provide faulty measurements (which happens from time to time), or if archaeologists misread stratigraphy and relation- ships on their sites (which may happen more often), or if specialists misinterpret sample taphonomy. Some issues may be undecidable in the end. That may simply require patience until new data can be acquired, but note that, for Jacques Derrida at least, undecidability was ‘decidedly not peaceful neutrality, or some courteous, liberal tolerance of diversity of opinion’ (Hill 2007: 48). Uncertain on principle: combining lines of evidence to create chronologies 219

Chronology creation in practice

Site-based chronologies All sites need chronology; without it, we lack the means to determine context, differentiate activity, and construct history. For these reasons, simple site-based studies dominate the corpus of chronological models that are currently available (e.g. Bayliss 2009). Figure 12.2 shows a Bayesian chronological model for the Mesolithic hut at Howick on the north-east coast of England. This model combines the assemblage of 27 calibrated radiocarbon dates (Bayliss et al. 2007b: Table 6.1) with the stratigraphic sequence of deposits shown in Figure 12.3. It also allows for the fact that all these deposits, and indeed radiocarbon samples, are related because they come from the same period of past activity in a particular place. Three radiocarbon dates have been excluded from this model for a variety of reasons (denoted by a ‘?’ in Figure 12.2). Beta-153650 was a hazelnut shell from a posthole dated following the initial assessment excavation on the site but, although useful at this stage of the project in confirming the Mesolithic age of the structure, this sample does not meet the strict taphonomic criterion used for the selection of the other radiocarbon samples from this site (in this case, a putative functional association between the dated charred plant material and the burning- related features from which the samples derived). The other two omitted dates have been excluded on the basis of their very low individual indices of agreement (Bronk Ramsey 1995: 426), which suggest mismatches between these radiocarbon dates and the available stratigraphic information. Given the sandy deposits on which the site is situated and the small size of the dated samples, OxA-11829 has been interpreted as residual from earlier activity and not actually derived from the firing of hearth 293, and OxA-11805 has been interpreted as intrusive from overlying Mesolithic activity. Further discussion of the construction of this model is provided by Bayliss et al. (2007b). The model suggests that the hut at Howick was first occupied in 7970–7760 cal. BC (95% probability; start structure; Figure 12.2) and was abandoned in 7740–7560 cal. BC (95% probability; end structure; Figure 12.2). It was in use for 40–380 years (95% probability; use hut; Bayliss et al. 2007b: Figure 6.4). (In this chapter, posterior density estimates derived from Bayesian modelling are given in italics with the relevant parameter name; this derives from the model defined in figures referenced.) The strength of this model is that it is based on a closely observed stratigraphic sequence, and a series of radiocarbon dates on single-entity samples of short-lived charred plant remains. These samples may reasonably be interpreted as functionally related to the burning features from which they were recovered, and there are sufficient dates to allow outliers and misfits to be identified. Its weakness is that the dated samples were very small, and the site was shallow and on a sandy subsoil. The movement of the sampled material between contexts cannot, therefore, be ruled out. Sequence Howick {A = 69.0% (A'c = 60.0%)} Boundary end structure Sequence Howick Mesolithic hut Phase 3 Phase hearth 173 OxA-11827 121.8% OxA-11826 95.0% Phase post hole 051 Beta-153650? 11.3% Phase hearth 047 OxA-11805? 0.0% OxA-12294 119.3% AA-41788 39.8% Event rebuild 2 Phase 2 Phase layer 210 Phase burnt patch 2 OxA-12324 112.8% OxA-12347 122.4% Phase hearth 158 OxA-11853 49.9% OxA-11832 60.5% Phase hearth 109 OxA-11830 124.3% OxA-11831 124.4% Event rebuild 1 Sequence 1B Phase hearth 268 OxA-11855 47.5% OxA-11854 103.9% Phase hearth 293 OxA-11828 97.5% OxA-11829? 0.3% Phase hearth 340 OxA-11804 84.3% OxA-12325 119.6% Phase 1A Phase hearth 383 OxA-12292 116.1% OxA-12402 42.1% Sequence hearths 379/357 Phase hearth 357 OxA-11802 122.0% Phase hearth 379 OxA-11857 116.2% OxA-11856 121.6% Sequence hearths 355/291 Phase hearth 291 OxA-11803 121.3% OxA-12326 121.7% Phase hearth 355 OxA-12327 80.9% OxA-11801 95.0% Boundary start structure

9000 cal. BC 8500 cal. BC 8000 cal. BC 7500 cal. BC 7000 cal. BC Posterior density estimate

FIGURE 12.2 Probability distributions of dates from the Mesolithic use of the Howick hut (after Bayliss et al. 2007b: Figure 6.2). Each distribution represents the relative probability that an event occurred at a particular time. For each of the dates two distributions have been plotted, one in outline, which is the result produced by the scientific evidence alone, and a solid one, which is based on the chronological model used. The other distributions correspond to aspects of the model. For example, the distribution start structure is the estimated date when the hut was initially constructed. The large square brackets down the left-hand side of the graph, along with the OxCal keywords, define the overall model exactly. Uncertain on principle: combining lines of evidence to create chronologies 221

end structure

hearth 047 hearth 173 posthole 051

rebuild 2 levelling layer 049

burnt patch 2 hearth 109 burning pit 158

rebuild 1 levelling layer 210

hearth 268

hearth 293

burnt spread 340

hearth 291 hearth 357

hearth 355 hearth 379 hearth 383

start structure

FIGURE 12.3 Schematic diagram showing the archaeological sequence included in the chronological model shown in Figure 12.2 (contexts from which no samples were radiocarbon dated are shown in grey).

On their own these results are of limited utility. Clearly, they suggest that the hut was constructed and used during the first half of the eighth millennium cal. BC, and that it was in use for a number of decades, if not a century or two. This duration may have implications for the persistence of place and the social structures of the group of people who used it, but these data become far more informative when we can compare the dating of the Howick hut with that of analogous structures from Britain and Ireland (Figure 12.4). These clearly cluster strikingly in the first half of the eighth millennium cal. BC, and contribute to the construction of new narratives about the history of population movement and settlement around the edges of a shrinking North Sea basin (Waddington et al. 2007). 222 Alex Bayliss and Alasdair Whittle

Phase Mesolithic Huts

Sequence Broom Hill, Hampshire England

pit 3 (structure)

Phase Upper Mount Sandel

end main hut

start main hut

Phase Howick

end structure

rebuild 2

rebuild 1

start structure

Phase East Barns

AA-54960

AA-54961

AA-54962

9000 cal. BC 8500 cal. BC 8000 cal. BC 7500 cal. BC 7000 cal. BC

Posterior density estimate

FIGURE 12.4 Probability distributions for the dates of occupation of Mesolithic huts analogous to Howick from the British Isles. (These distributions have been obtained from the chronological models shown in Bayliss et al. 2007b: Figure 6.2; Bayliss and Woodman 2009: Figures 6–10; and Waddington et al. 2007: Figures 15.4 and 15.5.)

Calendar dating of chrono-typologies Chrono-typologies aim to provide a classification and relative sequence of artefacts, based on the progressive development of traits. In the example considered here, these traits are the decorative styles appearing on different types of Early Bronze Age pottery found in graves in Ireland. This typological analysis was carried out by Anna Brindley (2007: Chapters 11–16), prior to and independently from the statistical modelling of the radiocarbon dates. The samples were, however, dated during the production of the typology and were used to inform the selection of chronologically diagnostic traits in the analysis. The pottery is classified first into six traditions – food vessel bowls, food vessel vases, vase urns, encrusted urns, collared urns, and cordoned urns (Figure 12.5) – and then each is sub-divided into a series of typological stages based on decoration. The traditions are interpreted as independent and potentially overlapping, connected only when a particular grave contains urns of different traditions (e.g. Grave 40 at Tara, Co. Meath, which contains both a food vessel bowl and a food vessel urn). In contrast, Brindley interprets her decorative stages as successive and generally non-overlapping. She reports details of 233 radiocarbon determinations from 178 deposits, almost all of which are funerary (Brindley 2007: Chapters 4–9). Uncertain on principle: combining lines of evidence to create chronologies 223

1 23

465

7 89

10 11 12

020cm

FIGURE 12.5 Pottery traditions in Early Bronze Age Ireland: 1–3 – stages 1–3 food vessel bowls; 4–6 – stages 1–3 food vessel vases; 7–9 – stages 1–3 encrusted urns; 10–12 – stages 1–3 vase urns. (Redrawn from Brindley 2007.)

Bayliss and O’Sullivan (2013) describe a Bayesian statistical model that combines all this information to suggest calendar dating for Brindley’s chrono-typology. Figure 12.6 shows an element of this model, relating to encrusted urns. This can be used to illustrate the process we use to decide how to incorporate these data in the model. First, we consider the scientific accuracy of all 30 radiocarbon measurements associated with this ceramic style. In this case, three results on cremated bone from Strawhall, Co. Carlow, are judged to be inaccurate on scientific grounds and are excluded from the model. Next we consider the taphonomic characteristics of the dated samples, and another measurement (GrN- 16790 from Bealick, Co. Cork) is excluded from the model as it seems to have contained an element of intrusive charcoal. Then we consider the character of the Sequence Encrusted Urns {A=60.0% (A'c=60.0%)} Boundary end Encrusted Urns Phase Encrusted Urns Sequence stages Phase stage 3 Last last Encrusted Urns 3 R__Combine Corradoon, Co. Waterford 88.6% TPQ unidentified charcoal GrN-11011 102.3% Event Encrusted Urns 213 Phase stage 2 GrA-14289 101.3% GrA-21724 98.4% GrA-14772 131.7% XReference GrA-17162 95.7% Phase Ballyveelish GrA-14286 90.5% TPQ unidentified charcoal GrN-11657 111.9% TPQ unidentified charcoal GrN-10634 109.5% CxA-2660 103.6% Event Encrusted Urns 1/2 Phase stage 1 GrA-14614 124.0% GrA-14602 103.7% GrA-14680 110.1% GrA-14600 129.0% XReference GrA-17279 118.3% XReference GrA-17276 116.4% XReference GrA-17195 116.1% R__Combine Cloghskelt (grave 1) 112.1% TPQ unidentified charcoal CxA-2679 72.7% First first Encrusted Urns 1 Phase miscellaneous TPQ charcoal from long-lived species CxA-2657 69.0% Phase Killydonoghue GrA-24054 114.3% TPQ unidentified charcoal GrN-28175 83.9% TPQ unidentified charcoal CxA-2658 75.3% GrN-11030 100.0% Boundary start Encrusted Urns

3500 cal. BC 3000 cal. BC 2500 cal. BC 2000 cal. BC 1500 cal. BC Posterior density estimate

FIGURE 12.6 Probability distributions of dates associated with encrusted urns in Ireland. The format is identical to that of Figure 12.2. The large square brackets down the left-hand side, along with the OxCal keywords (Bayliss and O’Sullivan 2013: Figures 26–33) define the overall model exactly. Uncertain on principle: combining lines of evidence to create chronologies 225 material dated, and nine dates are incorporated into the model only as termini post quos because they were on samples of unidentified charcoal or charcoal from potentially long-lived species. The final step is to incorporate the sequence of decorative stages suggested by Brindley (2007: Chapter 13) for encrusted urns. In this case, the overall index of agreement for the component model that incorporates this sequence is satisfactory (Aoverall: 115.6%),1 and so we judge this interpreta- tion to be statistically plausible. In other instances, the non-overlapping sequence of stages does not seem to be in accordance with the radiocarbon dates. In these cases, the minimum adjustments necessary to produce good statistical agree- ment between the dates and the typological series are made. For example, the sequence of three stages for food vessel vases does not produce good agreement (Aoverall: 47.4%), but when only stages 1 and 3 are made successive, and stage 2 is allowed to overlap with both, does the agreement become satisfactory (Aoverall: 75.0%). This model is based on a lifetime study of the typological succession of Bronze Age urns in Ireland and on an extensive suite of radiocarbon dates. Many of these exploit the new opportunity to date calcined bone (Lanting et al. 2001) from the cremation deposits themselves. The diagnostic statistics have also been useful in determining the plausibility of different readings of the typological information available. A potential weakness is that the methodology available at the time this study was undertaken assumed that each style began and then continued at a reason- ably constant rate until it ended. There was no allowance for a gradual introduction and a gradual decline of the different decorative traditions and forms. Whether this is a drawback in practice, given the evident rapidity of change in this sequence, is a matter of archaeological judgement. The impact of this research for our understanding of the Irish Early Bronze Age is illustrated in Figure 12.7. Not only is it now possible to suggest dating to within a few decades for some vessels with particularly diagnostic decoration (e.g. stage 3 encrusted urns), but it is also possible to see interesting patterns emerge. Why, for example, do so many kinds of finds disappear so abruptly in the second quarter of the nineteenth century cal. BC? In this instance, it is not yet possible to evaluate the possible answers, but the gain has been the setting of new questions, and the creation of new research agendas. These must involve, at the least, comparison with other evidence in Ireland, including metalwork, burial rites, pollen and other environmental records, and if possible settlements, as well as comparison with other evidence in Britain and beyond, in north-west Europe, for wider social networks.

Seriation Seriation is a process of ordering closed assemblages of archaeological material, most usually artefact types in graves or pits, or decorative or other traits on objects. Nowadays it is usually undertaken using statistical techniques, often correspondence analysis. FIGURE 12.7 bronze razors Schematic miscellaneous cordoned urns summary of the stage 1 stage 2 3 chronology of selected elements faience beads of the Early bronze daggers Bronze Age of Ireland: the date battleaxes estimates are derived from the miscellaneous collared urns stage 1 stage 2 stage 3 models defined in Bayliss and plano-convex, flint and bronze knives O’Sullivan (2013: Figures 26–33, miniature vessels 39–42, 43–46, 47, miscellaneous vase urns and 48–51), 1 stage 2 3 except for those for unshaded miscellaneous encrusted urns stage 1 stage 2 3 elements, where insufficient miscellaneous vases radiocarbon dates stage 2 are available to stage 1 stage 3 allow formal stage 1 stage 2 stage 3 modelling and so miscellaneous bowls estimates are informal. cists

cremation

inhumation

wedge tombs

2100 2000 1900 1800 1700 1600 cal. BC Uncertain on principle: combining lines of evidence to create chronologies 227

Our example is a recently completed study typology of 305 furnished male Anglo-Saxon burials in England, covering the period c.570–720 AD (Bayliss et al. 2013). It was a multi-disciplinary project that was undertaken over more than a loop seriation decade. This allowed us to develop an itera- tive and reflexive research cycle (Figure 12.8), employing interdependently the three prin- cipal techniques involved. radiocarbon Artefact types were initially defined and dating an incidence matrix of their occurrence in graves compiled. Correspondence analysis was undertaken on this matrix, and the typol- Bayesian ogy refined to maximise the chronological modelling variation in the types. A preliminary series of

loop radiocarbon dates was then obtained from skeletons in graves that fell along the whole FIGURE 12.8 The research cycle arc of the seriation, and a preliminary chrono- employed in the Anglo-Saxon logical model constructed that combined the chronology project. dates with the preliminary seriation. This model allowed us to refine both the definition of artefact types and the selection of chronologically sensitive types to be included in the correspondence analysis, and informed the selection of further skeletons for radiocarbon dating. This cycle was repeated three times before the typologies were finalised and the final seriation constructed. This final seriation was built in a cumulative manner, starting with the matrix of grave assemblages with shield boss and spearhead types (the most common finds in male graves), and iteratively adding further artefact types and grave assemblages. Further interim chronological models were also constructed after each stage. During this process, it was decided to omit some artefact types, which appeared not to be chronologically diagnostic or which were too rare for robust seriation. A small number of incidences (six) were also omitted, as the correspondence analysis suggested that they were anomalously late occurrences of earlier artefact- types – in effect, heirlooms. The final seriation contained a total of 272 graves and 78 artefact types. This was then partitioned in two ways: on the basis of leading artefact types (Bayliss et al. 2013: E-figure 6.6), and on the basis of the two- dimensional plot of the grave assemblages produced by the correspondence analysis (Figure 12.9). These partitions are very similar (and are discussed fully in Bayliss et al. 2013: Chapter 6.4). By way of illustration, the final chronological model for the partition shown in Figure 12.9 is given in Figure 12.10. This model incorporates the radiocarbon dates on graves included in the final seriation, along with the sequence of phases suggested by the correspondence analysis. It also includes a number of radiocarbon dates from graves that cannot be 2 2.95% AS-Mq AS-Mr UB-4731 (MH093) AS-Mp UB-4928 (SPTip250) UB-4731 (BuD375) UB-6479 (MH040) UB-4921 (MH081) UB-4925 (SPTip068) 1 UB-4508 (EH012) UB-4736 (Ber028) AS-Ms UB-4931 (SPTip318) UB-4923 (EH007) UB-4510 (EH048) UB-6474 (BuD264) UB-6477 (BuD414) UB-4985 (WG11) UB-4422 (SutH17) UB-4926 (SPTip212) UB-4886 (MelSG077) UB-4509 (EH003) UB-4962 (SPTip196) 3.36% UB-4929 (SPTip194) 0 –3 –2.5 –2 –1.5 –1 –0.5 0 0.5 1 1.5 BM-640 (SutH01) UB-4961 (SPTip008)

UB-46478 (SPTip360) UB-4924 (SPTip113) AS-Mt –1 Eris 104 UB-4505 (Lec172/1)

UB-4976 (Fo2) UB-4683 (Lec040) UB-4727 (GaH) –2

Earlier grave assemblages

–3

FIGURE 12.9 Final correspondence analysis of male grave-assemblages and –4 artefact types from the Anglo-Saxon chronology project, partitioned into five phases according to the two- dimensional map of grave assemblages

–5 produced by correspondence analysis. Boundary end AS-Mt cal AD 660–685 (95%) cal AD 665–680 (68%) R_Combine SPTip113 [A:42] R_Date UB-4982 (Lec155) [A:104] R_Date UB-4683 (Lec040) [A:104] R_Date UB-4505 (Lec172/1) [A:112] R_Date UB-6478 (SPTip360) [A:116] R_Date UB-4920 (GaH) [A:115] R_Date UB-4074 (But2297) [A:116] R_Date UB-4976 (Fo2) [A:41] Phase AS-Mt AS-Ms/AS-Mt cal AD 610–645 (95%) cal AD 620–640 (68%) R_Date UB-4423 (SutH17) [A:100] R_Date BM-640 (SutH01) [A:120] R_Date UB-4962 (SPTip196) [A:120] R_Date UB-4929 (SPTip194) [A:63] R_Combine MelSG077 [A:75] Phase AS-Ms AS-Mr/AS-Ms cal AD 585–615 (95%) cal AD 590–610 (68%) R_Date UB-4961 (SPTip008) [A:76] R_Date UB-4928 (SPTip250) [A:112] R_Date UB-4925 (SPTip068) [A:125] R_Date UB-4508 (EH012) [A:127] R_Combine EH048 [A:108] Phase AS-Mr AS-Mq/AS-Mr cal AD 565–595 (95%) cal AD 570–585 (68%) R_Date UB-4958 (BuD375) [A:129] R_Date UB-4931 (SPTip318) [A:140] R_Date UB-4731 (MH093) [A:140] R_Date UB-4736 (Ber028) [A:110] R_Date UB-4926 (SPTip212) [A:78] Phase AS-Mq AS-Mp/AS-Mq cal AD 550–570 (95%) cal AD 555–565 (68%) R_Date UB-6475 (BuD323) [A:58] R_Date UB-4509 (EH033) [A:122] R_Date UB-6474 (BuD264) [A:119] R_Date UB-4985 (WG11) [A:119] R_Date UB-6479 (MH040) [A:104] R_Date UB-4921 (MH081) [A:99] R_Date UB-6477 (BuD414) [A:93] R_Date UB-4923 (EH007) [A:92] Phase AS-Mp Sequence seriation & diagnostic artefact-types =UB-4882 (MelSG080) =UB-4884 (MelSG079) Phase =AS-Mp/AS-Mq Sequence after AS-Mp =AS-Mr/AS-Ms R_Date UB-4730 (MH079) [A:58] R_Date UB-4981 (Lec183) [A:117] Phase =AS-Mp/AS-Mq Sequence AS-Mq or AS-Mr =AS-Ms/AS-Mt R_Combine SPTip042 [A:82] Sequence before AS-Mt =AS-Mr/AS-Ms R_Date UB-4927 (SPTip263) [A:115] Sequence before AS-Ms =AS-Mq/AS-Mr R_Date UB-4682 (WG66) [A:110] Sequence before AS-Mr Phase artefact-types Phase grave-assemblage seriation Boundary start AS-Mp cal AD 525–550 (95%) cal AD 530–545 (68%) Sequence R_Date UB-4882 (MelSG080) [A:119] R_Date UB-4884 (MelSG079) [A:111] R_Date UB-4885 (MelSG078) [A:57] =MelSG077 Sequence Melbourn Phase [Amodel:84]

400 500 600 700 800

Posterior density estimate (cal. AD)

FIGURE 12.10 Probability distributions of dates from male graves from the Anglo- Saxon chronology project, incorporating the five-phase sequence for grave assem- blages suggested by the correspondence analysis shown in Figure 12.9, with the additional radiocarbon dates from graves containing diagnostic types (Bayliss et al. 2013: Table 6.3). The format is identical to that of Figure 12.2. The large square brackets down the left-hand side along with the OxCal keywords define the overall model exactly. 230 Alex Bayliss and Alasdair Whittle included in the seriation but contain artefact types that only occur in certain phases of the seriation (e.g. Grave 66 from Westgarth Gardens, Suffolk, which contains sword type SW4, which only occurs in phases AS-Mp and AS-Mq), and a stratigraphic sequence of burials from Melbourn, Cambridgeshire. The power of this approach is apparent from the precision of the date estimates shown on Figure 12.10. Ensuring that these precise date estimates are accurate is obviously critical, and considerable attention has been given to this in the completed study (Bayliss et al. 2013: Chapters 2, 6.5, 7.5 and 9.3). This study tests the limits of the accuracy of the methods employed. The lag between the position of a grave assemblage in the correspondence analysis (based on the average dates of manufacture of the artefact types included in it) and the burial may be detectable. Radiocarbon dates on human bone may incorporate an offset for the time it takes food to be incorporated in bone collagen. They may also include an offset for a component of fish included in a dated individual’s diet. The absolute accuracy of the radiocarbon calibration curve may also be a substantive con- sideration. These issues matter in a case study where the new dating allows the archaeology to become part of the historical narrative. Among a host of questions, how does the dating of a series of ‘princely graves’ to the decades around 600 AD relate to the emergence of elites and the consolidation of the early Anglo-Saxon kingdoms? Can the abrupt ending of furnished burial in the 670s and 680s AD be related to reforms of the Anglo-Saxon Church instituted by Bishop Theodore of Tarsus (669–690 AD)?

Timescapes People live in connection with others and move across the landscapes of their own time. Sites are often studied in isolation, but they belong in a wider, spatial context. Adding the spatial dimension to Bayesian chronological models is challenging, and it must be admitted that the approaches currently available are rather primitive (this is a particular area where we hope further statistical developments will improve our archaeological chronologies). Our examples come from the first centuries of the Neolithic of southern England, where we have completed an ambitious project to date the causewayed enclosures of southern Britain and Ireland and discuss them within their context in the early Neolithic period of those islands (Bayliss et al. 2011; Whittle et al. 2011). The first step was to produce detailed site-specific models for each of the thirty- six causewayed enclosures in southern Britain for which we had radiocarbon dates (Whittle et al. 2011: Chapters 3–11). The outputs of these models were then fed into a new model, which estimated the currency of the use of causewayed enclosures in southern Britain (Whittle et al. 2011: Figures 14.7–14.10). This model was divided into regions, partially chosen to ensure that each contained a statistically viable number of sites and partially for geographical coherence, and the date of the first enclosure in each was estimated. These date estimates are shown on Map 12.1. Uncertain on principle: combining lines of evidence to create chronologies 231

N

3705-3650 cal BC (3695-3660 cal BC)

3645-3550 cal BC (3640-3580 cal BC) 3680-3615 cal BC (3660-3630 cal BC)

3700-3605 cal BC (3660-3630 cal BC) 3685-3640 cal BC (3670-3645 cal BC) 3710-3665 cal BC (3700-3670 cal BC)

3680-3640 cal BC (3665-3645 cal BC) 3705-3650 cal BC (3690-3660 cal BC) 3705-3650 cal BC (3695-3660 cal BC)

0100km

MAP 12.1 Map showing the spread of causewayed enclosures across southern Britain. Date estimates are taken from the model defined by Whittle et al. (2011: Figures 14.7–10), at 95 per cent probability (68 per cent probability in brackets). (After Bayliss et al. 2011: Figure 14.16.)

Causewayed enclosures clearly appeared across southern Britain during the thirty-seventh century cal. BC, although they emerged first, probably from the last decades of the thirty-eighth century cal. BC, in coastal regions of the south-east (in the Thames estuary at 75% probability), and a generation or two later they were found further inland. To place this pattern in context, we then gathered technical details of more than 1,400 radiocarbon dates from other early Neolithic sites in southern Britain, Ireland, and Scotland south of the Great Glen. Details of the dated material and the taphonomic and stratigraphic associations of each of these samples were then assessed in relation to a range of ‘things and practices’ that appear across the British Isles at this time. A series of models was then constructed to estimate the dates when each of these things or practices was current in different areas of the British Isles and Ireland (Whittle et al. 2011: Chapters 12 and 14). A summary diagram illustrating the results of these models is given in Figure 12.11. Clearly some Neolithic things and practices first appeared in south-east England in the decades around 4000 cal. BC. About a century later these are found in south- central England, and then in the decades around 3800 cal. BC they quickly appear across the rest of the study area. Enclosures (with the exception of an isolated 3600

3700

3800 BC

l. 3900 a c

4000

4100

4200 South-east EnglandSouth-central EnglandSouth-west EnglandSoIreland uth ScotlandNorth-east Scotland

Cereals Domesticates Plain bowl ceramics Leaf-shaped arrowheads Ground axes

Flint mines Houses Monuments Enclosures Decorated ceramics

FIGURE 12.11 Schematic diagram showing date estimates for the appearance of Neolithic things and practices across selected areas of Britain and Ireland (derived from a series of Bayesian chronological models defined and discussed by Whittle et al. 2011: Chapter 14). Uncertain on principle: combining lines of evidence to create chronologies 233 example in the west of Ireland, Magheraboy, which merits further research) and decorated pottery appear later – in a concentrated horizon in the decades around 3700 cal. BC, apparently only in any number in the southern part of the British Isles. Although these models have enabled us to weave a new narrative for the first centuries of the Neolithic in southern Britain, they have obvious strengths and weaknesses. Our models for enclosures incorporate a large suite of radiocarbon dates on short-life, single-entity samples that were specificially chosen around the emergent Bayesian models. The stratigraphic information incorporated in those models, however, is of varying quality. Some have closely observed stratigraphies recorded to modern standards, but for others we are reinterpreting archives from as far back as the 1890s. Our sample of dated enclosures is large (about a third of those known), but predicated by those that have been excavated. For the wider study, we have inherited a large dataset of varying quality and spatial distribution. We have been fortunate because Bayesian modelling has been routine professional practice in England for nearly twenty years, and so we have many sites with statistically viable numbers of radiocarbon dates on short-life material. There is still, however, a proportion of dates on bulk samples or unidentified charcoal that may incorporate an old-wood offset. All these data have to be considered individually in detail. Our spatial models for both the enclosures and the early Neolithic context may both be sensitive to the definition of each region. More sophisticated spatiotem- poral models are sorely needed. There is, however, the risk that the technically demanding and labour-intensive nature of this kind of study will make existing, provisional models fossilise to become received wisdom, as reworking them is so time-consuming.

Multiple interpretations The construction and comparison of alternative models (known as ‘sensitivity analyses’) are a fundamental part of Bayesian statistical modelling. One component of a model is changed, and the model is rerun. The outputs from the original model and its variant are then compared. When these are very similar, then the model can be regarded as insensitive to the component of the model that has been varied. When the outputs differ markedly, the model is sensitive to that component. Sensitivity analyses are useful not only in determining how far the outputs of a model are stable, but also help us to identify which components of a model are most critical to the resultant chronology. As an example, we consider the dating of the Fussell’s Lodge long barrow (Wysocki et al. 2007; see Figure 12.12). The stratigraphic sequence of this monu- ment is open to alternative interpretations. Although raising the barrow mound itself clearly closed the sequence and was preceded by the construction of the flint cairn, alternative sequences for the timber mortuary structure and deposition of human remains have been suggested (see Wysocki et al. 2007: Figures 4, 6, and 9 The Fussell’s Lodge long barrow N

Flint Cairn

Burials

05 15m

FIGURE 12.12 The Fussell’s Lodge long barrow. Uncertain on principle: combining lines of evidence to create chronologies 235 for a summary of these alternative interpretations). These three sequences have been combined with the available radiocarbon dates in three alternative chronological models (Wysocki et al. 2007: Figures 5, 7, and 10). Key parameters from these models are illustrated in Figure 12.13. It is apparent that all three models agree in placing the construction of the barrow itself in the second half of the thirty-seventh century cal. BC, probably in the 3630s or 3620s. As the first model envisages a unitary construction, this is the only estimate provided by it. The second and third models also agree in suggesting that the mortuary structure was extended in the second quarter of the thirty-seventh century cal. BC, probably in the 3660s or 3650s. The only substantive difference between the model outputs concerns the estimated date for the construction of the primary mortuary structure. Depending on archaeological and osteological readings of this material, either this structure was built in the earlier part of the thirty-eighth century cal. BC (model 2), or in the decades around 3700 cal. BC (model 3), or in the second half of the thirty-seventh century cal. BC (model 1). All three models have good overall agreement and are statistically plausible, and on the basis of the statistical methodology currently available it is not possible to determine which of the models is most probable. The choice of a preferred model, if we are not to remain undecided, must therefore be made on the basis of

Fussell’s Lodge long barrow

model 3

build__barrow

extend__box

build__box

model 2

build__barrow

extend__box

build__box

model 1

build__barrow

4000 cal. BC 3800 cal. BC 3600 cal. BC 3400 cal. BC Posterior density estimate

FIGURE 12.13 Probability distributions for key parameters for the dates of constructional events at Fussell’s Lodge long barrow, according to alternative models presented by Wysocki et al. (2007: Figures 5, 7, and 10). 236 Alex Bayliss and Alasdair Whittle archaeological evidence. In either case, we have good reasons to think that our date estimates for the raising of the barrow and the extension of the mortuary structure are robust. If we choose a preferred model on archaeological grounds (ours is model 3), greater doubt must attach to the estimated date for the initial construction of the mortuary structure (since at least two of the date estimates for this parameter from these three models must be erroneous within the quoted uncertainty).

Consequences In our examples above, we have explored the mixture of assumptions, criteria, procedures, and combinations that go into the construction of Bayesian chrono- logical models; the models illustrated here are also different in scope. With the site- based model, we assume that our knowledge of stratigraphy, context, and sample taphonomy is robust, and samples were selected according to strict criteria; the model was built step by step, allowing also for individual results to be set aside. The implications of the outcome of a single site-based model, however precise, instantly take on more significance when compared to others. In the chrono-typological study, step-by-step model creation was also important, and sample selection criteria no less strict. The seriation study showed a more iterative procedure, with suc- cessive steps combining progressive refinements of both typology and correspon- dence analysis with the selection of more radiocarbon samples. The timescape study involved the combination of multiple site-based models across a finite (if imprecisely bounded) region; further significance for results was given by the construction of wider narratives for the unfolding of the context, derived in turn from the combination of other site-based models and currency models (for example, for pottery and other monument types). Finally, to underline the central importance of the assumptions and archaeological judgements that go into all models, an example was outlined of the effect of different readings of the (in this case) stratigraphic sequence on the outcomes of alternative models. Differences were evident in the earlier part of the site sequence, but there was reassuring convergence in the estimates for ending. There is no simple solution for how to choose between alternative models when they all satisfy existing criteria of agreement and coherence (we note that model validation and comparison are active areas of research in Bayesian statistics). In the end, at present the choice of the preferred model comes down to archaeological judgement, but this leaves open the possibility of further refinement, or indeed replacement, following alternative interpretation or further dating at some future point. The consequences of this kind of model creation are profound for all archae- ologies that rely on radiocarbon chronology – and, indeed, may extend the use of the method into archaeologies that currently make little use of the technique. Greater precision affects, first, timing. With more precise timing, it becomes possible to determine ‘seemaultaneity’ and sequence much more robustly, in ways previously unattainable. A substantial Mesolithic structure at one site in northern England is now accompanied by others over a much wider area, in the seventy- Uncertain on principle: combining lines of evidence to create chronologies 237 eighth to seventy-sixth centuries cal. BC; causewayed enclosures spread rapidly from east to west across southern Britain from the late thirty-eighth to the later thirty-seventh centuries cal. BC; some long barrows were closed at the same time as the first burst of enclosure construction came to an end, in the decades around 3625 cal. BC; extensive and evidently abrupt change in ceramic styles in Ireland happened in the mid-nineteenth century cal. BC; rapidly changing male fashions, as seen in early Anglo-Saxon graves, came and went at a generational scale across the sixth and seventh centuries cal. AD in southern England and ended abruptly in the 670s or 680s cal. AD. From more precise timing and sequence open up renewed options for inter- preting relationship and causation. Did substantial Mesolithic huts mark new territories inland as the landmass of Doggerland began to shrink? Does the east–west spread of causewayed enclosures indicate not only the general importance of the proximity of the European continent from which the practice must somehow have been derived, but also the reactivation of chains of connection seen earlier, at the start of the southern British Neolithic, from the forty-first century cal. BC? Did material changes in Ireland in the nineteenth century cal. BC mark a wider shift in the power and practices of the elite across the island? How does the pace of change in Anglo-Saxon materiality, and the abrupt end of furnished burial at a particular time, illuminate our understanding of the history of Anglo-Saxon England in the later sixth and seventh centuries AD? More precise models on their own will not answer all questions. In the case of early Neolithic southern Britain, for example, we are still left questioning the gap between the initial appearance of new things and practices across the region and the later establishment of causewayed enclosures, a puzzling interval given the existence by the late fifth millennium cal. BC of Michelsberg and Chasseen antecedents for this kind of enclosure practice (Whittle et al. 2011: Chapter 15). One way to bridge this gap will be routinely to acquire dates for the settlement record or environmental proxies. Was there some kind of magic number, in terms of population or settlement density, that was required to kick off the conditions in which enclosure construction and use were deemed appropriate? From more precise timings also flow new and, for many, often surprising esti- mates of duration. Clearly, it would be absurd to argue that everything happened quickly, over the short term. Demonstrably, some prehistoric sites, including tells for example (such as Vincˇa-Belo Brdo, Serbia: Schier 1996, 2000), formed and were used on a scale of centuries. That does not exclude asking questions of timescale within such spans, but it keeps the long term in view. Nonetheless, in contrast to the default chronological perspective noted earlier, it is fair to say that it has been the demonstration of short durations for phenomena previously assumed to have been either long-lasting or slowly changing that has caused most surprise among archaeologists dependent on radiocarbon chronologies. Short durations enable us to think at scales of the lifetimes and generations of people in the past (Bayliss and Whittle 2007; Whittle et al. 2011: Chapter 11), to which prehistorians are unaccustomed but which those working in historic periods expect. This not 238 Alex Bayliss and Alasdair Whittle only gets us closer to past human experience in solar time but can also refine our understanding of possible causal chains. In turn, more precise estimates of both timing and duration in combination provide the opportunity to refine and deepen our understanding of the pace or tempo of change. The models for the appearance of Mesolithic huts, for example, give a real-life example of something that has been argued for hunter-gatherers in general, on the basis of analogy: that they did not live in a timeless world and that they did have history (e.g. Kelly 2013). Hunter-gatherers lived in a world that could change rapidly, and they could take action quickly. The model for the spread of causewayed enclosures across southern Britain suggests that the adoption of this new practice was at first relatively slow, from the late thirty-eighth into the early thirty-seventh century cal. BC, but that its pace then quickened over the next two generations or so, before a lull at the end of the thirty-seventh century cal. BC and a further briefer renewal in the thirty-sixth century cal. BC (Whittle et al. 2011: Chapter 14). This is not yet an archaeology of events, but it does sketch an eventful horizon, in which powerful forces of emulation and competition can be inferred, in ways previously attainable only by analogy or at a more theoretical level. Add to that also the models for other changes in material and connections in the later thirty-eighth and earlier thirty-seventh centuries cal. BC, even though all the details of the development of settlement, subsistence, materiality, and exchange networks at this time, for example, are still far from clear (Whittle et al. 2011: Chapter 14), and an altogether more complicated scenario for innovation emerges. We can claim that we can begin to write historical-scale narratives in prehistory and use the archaeological record to inform our understanding of history itself. What, finally, is the wider consequence of this improvement in the scale of chronological resolution? Is it just to be something akin to the availability of dendrochronological dates for individual wetland sites – intrinsically interesting but generally without much effect on how broader narratives about the past are constructed? The chronological default in radiocarbon-dependent archaeologies, especially in prehistory, has been that of the long term; the Braudelian longue durée has been attractive to many prehistorians, even if this was not really the scale at which Braudel wrote about the Mediterranean in detail. That scale was that of conjonctures and histoire événementielle; Bayesian modelling opens up now the former, and gives at least a glimpse of the possibility of the latter (Whittle et al. 2011: Figure 15.28). But how then are we to combine the different timescales, which should now become routinely available across radiocarbon-dependent archaeologies if the discipline adopts the procedures and strict criteria required, and given that this is a matter for combination rather than the choice of one scale to the exclusion of others? One kind of response involves recourse to the notion of tacking, alternately considering different scales in a kind of hermeneutic spiral (Gamble 1999; Hodder 1999; Wylie 2002: 163–165). We question whether this is enough. There is a real danger of assuming that we somehow intuitively know what is the scope and scale of the long term, and of then reifying that, of always being willing to surrender the mass of messy detail for shorter timescales into the comforting embrace of the long Uncertain on principle: combining lines of evidence to create chronologies 239 term. Depending on the kind of narrative aimed at, generalisation about trend is valuable: the sun will warm Planet Earth and eventually life on Earth will disappear; agriculture replaced hunting and gathering; unequal social relations gradually emerged (Flannery and Marcus 2012). But it simply makes no sense to operate longer-term timescales without worrying about what constitutes them in the first place, and it is hard to envisage medieval or modern historians being content to analyse only at a scale of several centuries (Braudel himself certainly did not!). Since shorter timescales have only recently begun to emerge, their potential implications and impact are for the moment wide open. Perhaps in the long run we may choose to revert again to broader narratives; for the present we believe that we should do all we can, in the creation of chronologies, to maximise the detail of more precise histories, even if they unavoidably contain elements of both epistemological uncertainty and ontological indeterminacy.

Note 1 Aoverall is a standard technical term in OxCal (Bronk Ramsey 2009a: 426).

Acknowledgements We are very grateful to Wolfram Schier and the editors for a critical reading of a first draft of this paper, and to Kirsty Harding for help with the figures.

References Adams, W. Y. and Adams, E. W. (1991) Archaeological Typology and Practical Reality: A Dialectical Approach to Artefact Classification and Sorting, Cambridge: Cambridge University Press. Bailey, G. (1983) ‘Concepts of time in Quaternary prehistory’, Annual Review of Anthropology, 12: 165–192. Bayliss, A. (2009) ‘Rolling out revolution: using radiocarbon dating in archaeology’, Radiocarbon, 51: 123–147. Bayliss, A. and Bronk Ramsey, C. (2004) ‘Pragmatic Bayesians: a decade integrating radiocarbon dates into chronological models’, in C. E. Buck and A. R. Millard (eds) Tools for Constructing Chronologies: Tools for Crossing Disciplinary Boundaries, London: Springer, 25–41. Bayliss, A. and O’Sullivan, M. (2013) ‘Interpreting chronologies for the Mound of the Hostages, Tara, and its contemporary contexts in Neolithic and Bronze Age Ireland’, in M. O’Sullivan, B. Cunliffe, G. Cooney and C. Scarre (eds) Tara From the Past to the Future, Bray: Wordwell, 1–80. Bayliss, A. and Whittle, A. (eds) (2007) Histories of the Dead: Building Chronologies for Five Southern British Long Barrows, Cambridge Archaeological Journal, 17(1): supplement. Bayliss, A. and Woodman, P. (2009) ‘A new Bayesian chronology for Mesolithic occupation at Mount Sandel, Northern Ireland’, Proceedings of the Prehistoric Society, 75: 101–123. Bayliss, A., Bronk Ramsey, C., van der Plicht, J. and Whittle, A. (2007a) ‘Bradshaw and Bayes: towards a timetable for the Neolithic’, Cambridge Archaeological Journal, 17(1): supplement, 1–28. 240 Alex Bayliss and Alasdair Whittle

Bayliss, A., Boomer, I., Bronk Ramsey, C., Hamilton, D. and Waddington, C. (2007b) ‘Absolute dating’, in C. Waddington (ed.) Mesolithic Settlement in the North Sea Basin, Oxford: Oxbow Books, 65–74. Bayliss, A., Healy, F., Whittle, A. and Cooney, G. (2011) ‘Neolithic narratives: British and Irish enclosures in their timescapes’, in A. Whittle, F. Healy and A. Bayliss Gathering Time: Dating the Early Neolithic Enclosures of Southern Britain and Ireland, Oxford: Oxbow Books, 682–847. Bayliss, A., Hines, J., Høilund Nielsen, K., McCormac, F. G. and Scull, C. (2013) Anglo- Saxon Graves and Grave Goods of the Sixth and Seventh Centuries AD: A Chronological Framework, London: Society for Medieval Archaeology. Bertsch McGrayne, S. (2011) The Theory That Would Not Die: How Bayes’ Rule Cracked the Enigma Code, Hunted Down Russian Submarines, and Emerged from Two Centuries of Controversy, New Haven, CT: Yale University Press. Brindley, A. L. (2007) ‘The dating of food vessels and urns in Ireland’, Bronze Age Studies, 7: 7–22. Galway: National University of Ireland. Bronk Ramsey, C. (1995) ‘Radiocarbon calibration and analysis of stratigraphy: the OxCal program’, Radiocarbon, 37: 425–430. Bronk Ramsey, C. (1998) ‘Probability and dating’, Radiocarbon, 40: 461–474. Bronk Ramsey, C. (2001) ‘Development of the radiocarbon calibration program’, Radiocarbon, 43; 355–363. Bronk Ramsey, C. (2009a) ‘Bayesian analysis of radiocarbon dates’, Radiocarbon, 51: 37–60. Bronk Ramsey, C. (2009b) ‘Dealing with outliers and offsets in radiocarbon dating’, Radiocarbon, 51: 1,023–1,045. Buck, C. E., Cavanagh, W. G. and Litton, C. D. (1996) Bayesian Approach to Interpreting Archaeological Data, Chichester: Wiley. Buck, C. E., Litton, C. D. and Smith, A. F. M. (1992) ‘Calibration of radiocarbon results pertaining to related archaeological events’, Journal of Archaeological Science, 19: 497–512. Dieks, D. (2006) ‘Becoming, relativity and locality’, in D. Dieks (ed.) The Ontology of Spacetime, Amsterdam: Elsevier, 157–176. Einstein, A. (1905) ‘Zur Elecktrodynamik bewegter Körper’, Annalen der Physik, 17: 891–921. Einstein, A. (1915) ‘Die Feldgleichungen der Gravitation’, Sitzungsberichte der Preussischer Akademie der Wissenschaften zu Berlin, 25, November: 844–847. Einstein, A., Podolsky, B. and Rosen, N. (1935) ‘Can quantum-mechanical description of physical reality be considered complete?’, Physics Review, 47: 777–780. Flannery, K. and Marcus, J. (2012) The Creation of Inequality: How our Prehistoric Ancestors Set the Stage for Monarchy, Slavery, and Empire, Cambridge, MA and London: Harvard University Press. Gamble, C. (1999) The Palaeolithic Societies of Europe, Cambridge: Cambridge University Press. Greenacre, M. J. (2007) Correspondence Analysis in Practice, 2nd edn, Boca Raton, FL: Chapman and Hall/CRC. Heisenberg, W. (1958) Physics and Philosophy: The Revolution in Modern Science, New York: Harper & Row. Hill, L. (2007) The Cambridge Introduction to Derrida, Cambridge: Cambridge University Press. Hodder, I. (1997) ‘“Always momentary, fluid and flexible”: towards a reflexive excavation methodology’, Antiquity, 71: 691–700. Hodder, I. (1999) The Archaeological Process: An Introduction, Oxford: Blackwell. Hodder, I. (2012) Entangled: An Archaeology of the Relationships between Humans and Things, Chichester: Wiley-Blackwell. Uncertain on principle: combining lines of evidence to create chronologies 241

Ingold, T. (1996) ‘Against the motion (1)’, in T. Ingold (ed.) Key Debates in Anthropology, London: Routledge, 112–118. Jacomet, S., Leuzinger, U. and Schibler, J. (2004) Die jungsteinzeitliche Seeufersiedlung Arbon Bleiche 3: Umwelt und Wirtschaft, Frauenfeld: Dept. für Erziehung und Kultur des Kantons Thurgau. Jones, A. M. (2012) Prehistoric Materialities: Becoming Material in Prehistoric Britain and Ireland, Oxford: Oxford University Press. Joyce, J. (1939) Finnegans Wake, London: Faber and Faber. Kelly, R. J. (2013) The Lifeways of Hunter-Gatherers, New York: Cambridge University Press. Lanting, J. N., Aerts-Bijma, A. T. and van der Plicht, J. (2001) ‘Dating of cremated bones’, Radiocarbon, 43: 249–254. Lindley, D. V. (1985) Making Decisions, 2nd edn, London: Wiley. Menotti, F. and O’Sullivan, A. (eds) (2013) The Oxford Handbook of Wetland Archaeology, Oxford: Oxford University Press. Miller, A. I. (1990) ‘Probability and the roots of Werner Heisenberg’s Uncertainty Principle paper’, in A. I. Miller (ed.) Sixty-Two Years of Uncertainty: Historical, Philosophical, and Physical Inquiries into the Foundations of Quantum Mechanics, NATO Science Series B, Physics, 226, New York: Plenum Press, 3–15. Minkowski, H. (1952) ‘Address to the 80th assembly of German natural scientists and physicians’ (1908), in H. A. Lorentz, A. Einstein, H. Minkowski and H. Weyl (eds) The Principle of Relativity: A Collection of Original Memoirs on the Special and General Theory of Relativity, New York: Dover, 75–91. Müller, J., Rassmann, K. and Kujundzˇic´-Vejzagic´, Z. (2013) ‘Okolisˇte — Rekonstruktion spätneolithischer und frühneolithischer Siedlungsprozesse des zentralbosnischen Visobeckens: Fragestellunge, Forschungsstrategien, Ergebnisse’, in J. Müller, K. Rassmann and R. Hofmann (eds) Okolisˇte 1 — Untersuchungen einer spätneolithischen Siedlungskammer in Zentralbosnien, Bonn: Habelt, 11–68. Neustupny´, E. (1998) ‘Structures and events: the theoretical basis of spatial archaeology’, in E. Neustupny´ (ed.) Space in Prehistoric Bohemia, Praha: Institute of Archaeology, Academy of Sciences of the Czech Republic, 9–44. Newton, I. (1687) Philosophiæ Naturalis Principia Mathematica, London: Joseph Streater for the Royal Society. Olivier, L. C. (2001) ‘Duration, memory, and the nature of the archaeological record’, in H. Karlsson (ed.) It’s About Time: The Concept of Time in Archaeology, Göteborg: Bricoeur Press, 61–70. Schier, W. (1996) ‘The relative and absolute chronology of Vincˇa: new evidence from the type site’, in F. Dras¸ovean (ed.) The Vincˇa Culture, its Role and Cultural Connections, Timis¸oara: The Museum of Banat, 141–162. Schier, W. (2000) ‘Measuring change: the Neolithic pottery sequence of Vincˇa-Belo Brdo’, Documenta Praehistorica, 27: 187–197. Schwartzschild, K. (1916) ‘Über das Gravitationsfeld eines Massenpunktes nach der Einsteinschen Theorie’, Sitzungsberichte der Preussischen Akademie der Wissenschaften,3 February: 189–196. Thiher, A. (2005) Fiction Refracts Science: Modernist Writers from Proust to Borges, Columbia, MI and London: University of Missouri Press. Waddington, C., Bailey, G., Bayliss, A. and Milner, N. (2007) ‘Howick in its North Sea context’, in C. Waddington (ed.) Mesolithic Settlement in the North Sea Basin, Oxford: Oxbow Books, 203–224. Waterbolk, H. T. (1971) ‘Working with radiocarbon dates’, Proceedings of the Prehistoric Society, 37(2): 15–33. 242 Alex Bayliss and Alasdair Whittle

Weninger, F., Steier, P., Kutschera, W. and Wild, E. M. (2010) ‘Robust Bayesian analysis: an attempt to improve Bayesian sequencing’, Radiocarbon, 52: 962–983. Whittle, A., Healy, F. and Bayliss, A. (2011) Gathering Time: Dating the Early Neolithic Enclosures of Southern Britain and Ireland, Oxford: Oxbow Books. Wylie, A. (2002) Thinking with Things: Essays in the Philosophy of Archaeology, Berkeley and Los Angeles, CA: University of California Press. Wysocki, M., Bayliss, A. and Whittle, A. (2007) ‘Serious mortality: the date of the Fussell’s Lodge long barrow’, Cambridge Archaeological Journal, 17(1): supplement, 65–84. 13 LESSONS FROM MODELLING NEOLITHIC FARMING PRACTICE

Methods of elimination

Amy Bogaard

Modelling Neolithic farming practice hinges on relevant comparisons between present-day crop husbandry regimes and fragmentary evidence of past cultivation methods. Such comparisons need to incorporate the wide range of farming scenarios featured in debate surrounding the transition to agriculture but also, crucially, to be framed in terms of characteristics that relate directly to key eco- logical variables distinguishing these regimes. This chapter considers a series of comparisons used in Bogaard (2004) to eliminate implausible Neolithic farming scenarios based on the ecological configuration of arable weed flora, which provide a detailed reflection of crop growing conditions and hence farming practices. The study made the case that the crop husbandry practices of Neolithic farmers more closely resembled ‘intensive garden cultivation’ – labour-intensive management of long-established arable plots – than it did shifting (or slash-and-burn) cultivation, extensive cultivation with the ard plough or floodplain cultivation. The strategy employed can be likened structurally to a dichotomous identifi- cation key, and constitutes an instance of ‘inference to the best explanation’ (ITBE), a form of inductive reasoning widely encountered in archaeology (Kelley and Hanen 1988; Fogelin 2007). Harman described the ITBE process as follows: ‘one infers, from the premise that a given hypothesis would provide a “better” explana- tion for the evidence than would any other hypothesis, to the conclusion that the given hypothesis is true’ (Harman 1965: 89). More recently, Lipton (2004) has emphasised that ITBE concerns inference to ‘the best of the available potential explanations’ (Lipton 2004: 58), and draws attention to two ‘filters’ at work: the selection of plausible candidate explanations, and the choice of the best from within this pool (Lipton 2004: 59). The following sections describe these steps in the case study, while the discussion reflects critically on the limitations of the work and forecasts two complementary strategies for evaluating and refining these inferences. 244 Amy Bogaard

Debate surrounding the nature of Neolithic farming The broad chronology of the spread of agriculture across Europe, and its mixed character as a cultivation and herding ‘package’, is well established (Price 2000). The domesticates involved were almost entirely of West Asian origin, with the addition of opium poppy from the central or western Mediterranean (Bakels 1982). It is also clear that the predominant crop and livestock species varied regionally and through time (e.g. Colledge and Conolly 2007; Colledge et al. 2013). Major points of ongoing debate surrounding Neolithic agriculture concern the nature of early farming as a set of practices: decisions about where and when to plant crops, how to prepare soil, whether/how much to manure/midden/water, how to sow, the extent of weeding during the growing season and so on. This web of decision- making is critical for understanding the social implications and long-term conse- quences of the agricultural transition. A long tradition of archaeobotanical recovery in Central Europe gives this region an unparalleled richness in direct evidence for farming practice. This evidence consists of the (usually charred) remains of crops and particularly the arable weeds harvested and preserved with them, which can provide nuanced ecological information on crop growing conditions (Bogaard 2004). A given species of crop can generally tolerate a range of growing conditions and so be grown under a variety of regimes; the identification of crop remains per se does not therefore offer detailed insight into farming methods. Particular constellations of weed species associated with crops, on the other hand, can pinpoint specific sets of crop growing conditions. It is the ecology of constituent weed species in an archaeobotanical assemblage, therefore, that offers key evidence of crop growing conditions and hence the specific agency of farming in any particular case. Meaningful interpreta- tion of this evidence requires comparison with weed survey data from relevant present-day agricultural regimes, a problem to be explored further below. For Central Europe, and the continent as a whole, four major models of early farming have emerged in the archaeological literature, deriving from a mixture of ethnographic analogy and what are often conflicting interpretations of indirect (non-archaeobotanical) evidence. To varying degrees these models have been linked into the central concern of many studies of Neolithic Europe: the origin and ethnic identity of the first farmers. Table 13.1 summarises the four models and shows how they can be distinguished by three ecological variables: permanence, intensity and seasonality. The initial question is: were cultivation areas transient or permanent? The shifting cultivation model fell out of favour in the 1970s (e.g. Modderman 1971) but has re- emerged in connection with interest in continuity between the Mesolithic and Neolithic, the idea being that a ‘transitional’ form of mobile cultivation would be conducive to adoption by indigenous hunter-gatherers (e.g. Whittle 1996; Thomas 1999). Shifting cultivation is associated ethnographically with communal land- holding and egalitarianism (e.g. Brown 1978). Also, livestock play no inherent role in this system (Rowley-Conwy 1981), and so this kind of Neolithic farming would imply that cultivation and herding were separate (if parallel) activities. Lessons from modelling Neolithic farming practice 245

TABLE 13.1 The relationship of the four major crop husbandry models to the three variables of permanence, intensity and seasonality (after Bogaard 2004: Table 3.1)

Models Permanence

Shifting cultivation Transient plots Intensity

Extensive ard Extensive Seasonality cultivation

Floodplain Fixed plots Spring sowing cultivation Intensive Intensive garden Autumn or spring cultivation sowing

If shifting cultivation can be eliminated, the next question is whether cultivation was extensive (large scale, relying on animal traction and ard ploughing) or inten- sive (small scale and garden-like) (see Table 13.1). The paradox of extensive cultivation is that, while yields per unit area are low, large-scale surplus production is possible given the efficiencies of specialised plough oxen and availability of sufficient labour at harvest time (Halstead 1995). Large-scale surplus production is associated with social stratification, including unequal access to land, as reflected in deployment of landless labourers at harvest (Goody 1976). Plough cultivation is usually a male activity, in part because extensive ploughing tends to take place far away from the home base. By contrast, intensive cultivation means that high inputs of labour are used to create high yields per unit area; inputs include careful tillage, regular weeding of crops, and use of manure and midden material to maintain high fertility levels, as well as watering (if needed). Intensive crop cultivation is restricted in scale by the labour capacities and requirements of families; ethnographic observations suggest that a farming family practising intensive cultivation of staple crops may require as little as a hectare or even less depending on yields (Bogaard 2004: 42–43). Unlike extensive ard cultivation, the intensive form does not necessarily imply social stratification, but it does harbour greater potential than shifting cultivation to encourage the development of social inequalities between households. Labour is the limiting factor for productivity in an intensive system, and this then raises the question of whose labour is critical; ethnographic evidence suggests that women and children play a central role in horticulture as a small-scale activity close to home (e.g. Watson and Kennedy 1991). If extensive ard cultivation is eliminated, the final question – for distinguishing floodplain and intensive garden cultivation – concerns seasonality, or sowing time: were crops autumn-sown, spring-sown or both? This issue was highlighted over 30 years ago by Karl Butzer’s suggestion that crop sowing times shifted from autumn to spring as farming spread from south-east to north-west; thus sowing time potentially reflects the adaptation of Mediterranean farming to Continental 246 Amy Bogaard and Atlantic European environments (Butzer 1972: 580). More immediately, crop sowing time has implications for labour scheduling and for crop yields (e.g. Gregg 1988: 70–78). Table 13.1 shows that shifting cultivation is distinguished from all other models by the transience of arable plots. Extensive cultivation is distinguished from the two remaining models by low cultivation intensity (relatively low levels of soil fertility and disturbance). The difference between floodplain and intensive garden cultivation is really a question of whether high levels of soil fertility and disturbance are present naturally (on alluvial soils, as suggested by Sherratt (1980)) or artificially, as the result of intensive labour inputs. Floodplain cultivation potentially differs from garden cultivation in that all crops must be sown in the spring (to avoid flood damage), but it should be emphasised that intensive cultivation could involve autumn sowing or spring sowing or both. (In fact, floodplain cultivation is suscep- tible to other, geomorphological lines of enquiry into the formation of floodplain habitats in the Early-Middle Holocene, and appears unlikely on those grounds (Bogaard 2012: 27).)

Weed ecological approaches Table 13.1 draws out major ecological gradients (permanence, intensity, season- ality) that are socially critical in a context of early farming and have a dramatic effect on weed flora. The key to operationalising this table is to find a meaningful way of comparing relevant present-day weed flora, developed under known conditions, and fragmentary archaeobotanical weed data. In archaeobotany there has been active interest in the interpretation of weed assemblages for several decades, particularly in Europe (e.g. Knörzer 1971; Wasylikowa 1981). This focus arose in part because archaeobotanical weed assemblages from early farming contexts tend to be quite different in their species composition to cereal weed communities of the recent past. This contrast would appear to imply a certain ‘otherness’ in early farming practice, but species differ- ences could simply be a function of comparing different times/places with contrast- ing flora. Moreover, the lack of taxonomic comparability between modern and archaeological weed floras sets up the key problem: a ‘common currency’ is needed in order to compare ancient and modern weed data. Different solutions to this problem have been proposed (for further discussion, see Jones 2002; Bogaard 2004: 5–7). One approach is to use phytosociology, based on the identification of plant ‘communities’ associated with certain growing condi- tions. Communities are groups of co-occurring species, arranged hierarchically in a ‘syntaxonomic’ system (e.g. Westhoff and van der Maarel 1973). There are two major difficulties with this solution: first, modern communities are contingent on changing species distributions, and on the creation and disappearance of ecologi- cally specific habitats; second, the link between species and growing conditions is based on field observations, not on any causal account of species’ requirements or tolerances. In practice, use of phytosociology enables only very broad comparisons Lessons from modelling Neolithic farming practice 247 between generalised modern weed communities (at high syntaxonomic levels) and archaeobotanical assemblages, resulting in problems of equifinality; similarities/ differences could be explained by one or more aspects of farming method (e.g. Jones 1992). The problems raised by phytosociology point toward a second requirement of the ideal ‘currency’: that it be based on characteristics that articulate with relevant ecological variables, and at sufficient resolution to disentangle different aspects of farming practice. In other words, we wish to construct relational analogies (Wylie 1985) between archaeological and present-day weed flora in order to assess similarity or difference in underlying causes – here, crop growing conditions and the farming practices that shaped them. Functional plant ecology and measurement of plant ‘functional attributes’ – physical or behavioural characteristics of plants that predict their potential success under different ecological conditions – provide just such a means of comparison. The ‘functional interpretation of botanical surveys’ (FIBS) approach was developed at the Unit of Comparative Plant Ecology, University of Sheffield as a method for analysing vegetation surveys and predicting consequences of changing land use (e.g. Hodgson et al. 1999). Each plant species has a unique set of morphological traits, which of course is what makes it possible for us to identify them; but the functional ecological approach recognises that species sharing the same habitat also tend to share certain functional attributes (Bogaard 2004: 7–9). For example, lamb’s succory (Arnoseris minima (L.) Schweigg. & Korte) and annual androsace (Androsace maxima L.) share certain obvious morphological similarities – both are short plants with small leaves – even though they belong to different taxonomic families and are not closely related. In functional attribute terms, we would say that they have a low canopy height – limiting their phytosynthetic potential and competitive ability under productive conditions – and a low specific leaf area (ratio of leaf area: dry weight) meaning that they grow relatively slowly (e.g. Reich et al. 1992). These are characteristics that function to adapt these species to open, relatively unproductive conditions; they are also autumn-germinating annuals, which gives them a long period of (slow) growth. By contrast, chickweed (Stellaria media (L.) Vill.) and knotweed (Polygonum aviculare L.) have medium canopy height and small but thin, quickly growing leaves; note also that they have ‘indeterminate growth’ structure – flowers within the axils of leaves – meaning that they may be flowering at one end but already producing seeds at the other. These characteristics, combined with a long germination period, enable them to grow under productive and highly disturbed conditions. Finally, black nightshade (Solanum nigrum L.) and maple- leaved goosefoot (Chenopodium hybridum L.) represent a third ‘functional type’: they have large canopies and large thin leaves, characteristics that enable them to compete under highly productive growing conditions, though they are more easily eradicated under frequently disturbed growing conditions than the medium-sized species. The specific advantages of using functional weed ecology to interpret and apply the weed survey data are that functional attributes provide a causal account of similarities and differences between modern and ancient weed flora; and that 248 Amy Bogaard they allow distinct ecological factors – such as fertility and disturbance – to be monitored independently (Charles et al. 1997; Jones 2002; Bogaard 2004: 7–9).

Methods for comparing present-day and ancient weed data The remaining challenge is to identify modern arable situations of relevance to the models of Neolithic farming summarised above (Table 13.1). Traditional farming areas and agricultural experiments provide weed survey datasets that can be compared with archaeobotanical weed data in terms of their ecological characteris- tics, as gauged through functional attribute measurements. Three modern weed survey studies provide floristic data relevant to the critical variables (permanence, intensity and seasonality) distinguishing the four models. Permanence was assessed using an experiment conducted in the Hambach Forest near Cologne in the 1970s, which provides data on the weeds growing in the first six years following clearance of long-lived deciduous woodland on loess (Bogaard 2002, 2004). The weed survey data suggest that woodland species are characteristic of recently cleared plots managed with little to no soil disturbance (tillage), as in a shifting cultivation regime, and that such weed flora are also dominated by peren- nials. General habitat category and life history of weed species are thus sufficient to confirm or exclude shifting cultivation. A weed survey study of traditional pulse cultivation on Evvia, Greece (Jones et al. 1999) encompassed a contrast between manually cultivated gardens that were manured and regularly hoed/weeded, and ard-ploughed fields managed with little manuring or weeding. Weed survey data from these plots had been subjected to functional ecological analysis, and a linear equation incorporating a set of attributes reflecting weed species’ potential in relation to soil productivity and disturbance was found to discriminate successfully between intensive and extensive manage- ment (Jones et al. 2000). Moreover, this discriminant equation was shown to classify intensively managed cereal plots surveyed in another study area, Asturias in northern Spain, correctly (Charles et al. 2002), even when based on simple presence/absence data for the constituent weed species rather than on more precisely quantified frequencies. This last point is important archaeologically, because archaeobotanical weed data are quantified in a completely different way to modern weed quadrat data – as seed counts – and presence/absence data thus provide the only way of comparing modern and ancient datasets directly. For seasonality, a modern phytosociological dataset from Central Europe was used that captured the contrast between autumn-sown and spring-sown crops (Bogaard et al. 2001). A functional attribute relating to the timing and duration of the flowering period was found to be a successful predictor of germination time, and to discriminate effectively between the two sowing regimes. This discriminant equation was also shown to classify autumn/winter cereal plots in Asturias correctly (Charles et al. 2002), even when based on presence/absence data. A final methodological point concerns the selection of suitable archaeobotanical samples for comparison with the modern datasets. Key criteria include richness in Lessons from modelling Neolithic farming practice 249 charred weed seeds, the associated crop content of the samples (as evidence of arable origin but also to assess the biasing effects of crop processing stages – winnowing, sieving, etc. – on weed content; see Jones 1992) and the integrity of the archaeo- logical (stratigraphic) context (Bogaard 2004: 60–68).

Archaeobotantical results The ‘dichotomous key’ of farming models summarised in Table 13.1 was applied to archaeobotanical sample weed data from Neolithic Central Europe (Bogaard 2004: 96–114). The modern weed survey datasets outlined above were used as analogues for the farming models, and the archaeobotanical weed data were com- pared to them in a step-wise fashion using relevant weed ecological characteristics as the basis for comparison. Shifting cultivation is unlikely given the dominance of annuals and rarity of woodland species in the archaeobotanical dataset; this model can therefore be excluded. The discriminant analysis function extracted to distinguish gardens and fields in the modern Evvia study classified almost all of the Central European samples as deriving from intensive cultivation (Bogaard 2004: Table 5.4). Finally, the discriminant analysis function extracted to distinguish autumn and spring-sown crops identified most of the Central European samples as deriving from autumn- sown crops, despite a bias imposed by crop processing in favour of spring sowing indicators (Bogaard 2004: Table 5.3; Bogaard et al. 2005); the reliability of the ‘spring sowing’ identification of some samples is therefore in doubt, though variability in sowing time would not be inherently surprising. Combined with geomorphological data suggesting that Neolithic floodplain habitats in the study region were generally unsuited to cultivation (Bogaard 2012: 27), floodplain cultivation can be excluded. ‘Intensive garden cultivation’ emerges as the most plausible model.

Discussion and conclusions: inference to a ‘better’ explanation? The elimination strategy outlined here consisted of a series of direct comparisons between archaeobotanical weed assemblages and modern weed datasets representing the various scenarios previously proposed for early farming. Relational analogies were framed on the basis of weed ecological characteristics that provide a causal account of species’ potential in relation to relevant environmental gradients. Two key ‘filtering’ processes define (and limit) this case study: the selection of farming models to be considered, and the method of elimination. Taking first the selection of models, a fundamental problem for ‘inference to the best explanation’ is suggested by Harman’s (1965) account: that by excluding certain possibilities, we make the further inference that the remaining possibility is ‘true’. Put another way, the reliability of the ‘answer’ we end up with ultimately depends on short-listing of the ‘true’ scenario among the alternatives under consideration. According to Lipton’s account of ITBE, however, a distinction must be drawn between actual 250 Amy Bogaard and potential explanations (Lipton 2004: 58), and the ‘best’ explanation is only a potential one, and hence provisional. The process could therefore be construed as one of ‘inference to a better explanation’, in that the ‘solution’ (merely) provides a better account of the evidence than the competing explanations; it may not be the ‘best’ explanation that will ever arise. In this case study, the four models to be distinguished represent the major alternatives that have emerged from archaeological debate surrounding early farming in Europe, or through ethnographic research motivated by that debate. These models are therefore the outcomes of background beliefs concerning the European Neolithic, beliefs that ‘are themselves the result of explanatory inferences whose function was to explain different evidence’ (Lipton 2004: 150). Concern over unconsidered alternatives is countered by allowing that researchers have considerable inductive powers, but also by the diversity of interests that gave rise to these disparate models in the first place through iterations of ITBE (Lipton 2004: 144). As the debate over early farming continues to evolve, new alternatives (as well as new evidence) will emerge, giving rise to further inferences to better explanations. Turning to the method of elimination, brief comment is warranted on several issues: 1 the robusticity of functional ecology as a basis of inference; 2 the suitability of the modern weed survey datasets for representing the various alternatives; and 3 the use of discriminant analysis (to assess the probability of intensive versus extensive cultivation, and of autumn versus spring sowing). First, functional ecology is a ‘soft approach to a hard problem’ (Hodgson et al. 1999): a series of easy-to-measure attributes are used as proxies for much more complex characteristics (e.g. specific leaf area as a proxy for growth rate). While individual attributes can be misconstrued, suites of attributes provide more robust indications of species’ ecological potential. Nevertheless, bringing a completely different kind of dataset to bear on the problem of growing conditions (such as stable isotope determinations on crop material – see below) would have the potential to expose systematic error introduced by the functional ecological method, and could also shed light on a different (and complementary) aspect of the ‘causal history’ of Neolithic crop growing conditions (see Lipton 2004: 144). By incorporating more strands of evidence, the explanation increases in breadth, becoming ‘richer’ as well as more reliable (e.g. Lipton 2004: 122). Second, the modern weed survey datasets provide ‘thin’ coverage of the models to be considered: each alternative is represented by only one modern study. This is a particular issue for the study of cultivation intensity on Evvia; while the Hambach Forest study of plots established in recently cleared woodland is corroborated by data from a more recent experiment (Rösch et al. 2002), and the German dataset pertaining to autumn versus spring sowing was itself a synthesis of numerous previous phytosociological surveys (Hüppe and Hofmeister 1990; Bogaard et al. 2001), the Evvian study represents a localised instance of contrast in soil productivity and disturbance (Jones et al. 1999, 2000). The problem is not only the narrow geographical scope of the study, but also that ‘intensity’ is itself a complex, composite variable that needs to be deconstructed further (see below). Lessons from modelling Neolithic farming practice 251

Third, the use of discriminant analysis as a means of choosing between alter- natives raises the question of how statistical probability fits into ITBE as a means of constructing a compelling explanation. The crucial point here is that the discriminant analyses were designed with explanatory considerations in mind (see Lipton 2004: 107–108): rather than using a haphazard selection of features to discriminate between alternatives, relevant functional attributes (i.e. proxies for plant ecological characteristics that relate mechanistically to species’ potential in response to pertinent environmental variables) were used as the discriminating variables to derive linear equations that maximise separation between contrasting modern regimes (autumn versus spring sowing, intensive versus extensive cultivation). The elimination framework described here could be made more robust in two ways. First, further modern weed datasets could be assembled to assess each model; this strategy is likely to be particularly worthwhile for the assessment of ‘intensity’, since this is a complex variable incorporating soil productivity and disturbance. The study of intensive and extensive pulse cultivation in Evvia, Greece, for example, was subsequently broadened by a study of intensive cereal cultivation in Asturias (Charles et al. 2002), and a new complementary study of extensive cereal cultivation in Haute Provence was undertaken in the summer of 2013 (Bogaard et al., unpublished data, under analysis) with a view to the construction of a new, cereal- based ‘intensity axis’ for comparison with archaeobotanical weed assemblages. Multiple ‘axes’ of intensity appear useful because each comparison represents a different balance between soil productivity and disturbance; it is envisioned, for example, that comparison of archaeobotanical data to the original Evvia study could be followed by further comparison(s) with other contrasting pairs of relatively intensive versus extensive management, in order further to refine inferences on the nature of past farming regimes that may lack a precise modern analogue. Second, methodological approaches different in kind to the functional ecological approach could be used to broaden the terms of comparison between modern and ancient weed datasets, both to allow for exposure of error introduced by any single method, but also potentially to deepen the explanation of complex ecological variables such as intensity. To this end, stable nitrogen and carbon isotope analysis of present-day crops grown under a range of traditional and experimental regimes has been undertaken, since this approach provides direct access to aspects of crop husbandry – manuring and water management – that are ecologically and socially significant, and relevant to assessing intensity (Bogaard et al. 2007; Fraser et al. 2011; Wallace et al. 2013). So far these new lines of evidence broadly support the case for intensive cultivation but also suggest a complex reality in which management intensity was contingent on a range of localised factors (Bogaard et al. 2013). Indeed, elimination of shifting, extensive ard and floodplain cultivation models, resulting in the inference to small-scale, labour-intensive cultivation, does not mean that early farming across Central Europe was homogeneous and uniform; rather, this general model brackets a range of potential scenarios. Farming regimes demonstrably varied among regions – within the LBK (Linearbandkeramik or Linear Pottery Culture) of Central Europe (c.5500–5000 cal. BC), for example (Bogaard 252 Amy Bogaard

2012) – but also within the arable landscape of a single community: recent analysis of a particularly well-documented LBK site in south-western Germany, Vaihingen an der Enz, shows that variation in soil disturbance and possibly soil pH levels related to ‘neighbourhood’ groupings of houses within the community, suggesting that blocks of arable land were farmed by these supra-household groups (Bogaard et al. 2011). This ‘social geography’ of farming only emerged through an exploration of diversity amongst archaeobotanical samples in relation to associated context and material culture. The elimination strategy outlined here complements this exploratory approach by establishing how the ancient weed data compare to modern datasets. A concluding ‘lesson’ from this case study might be that ‘inference to the best explanation’ entails a process of reiteration and refinement. New evidence and/or methodologies are brought to bear on the phenomenon to be explained in order to enhance the ‘virtue’ of the explanation that emerges (as an account of causal history, diverse data types and so on). This process is in tension with the propensity to infer that the ‘best’ explanation (at any given moment) is true (Harman 1965: 89).

References Bakels, C. C. (1982) ‘Der Mohn, die Linearbandkeramik und das westliche Mittelmeergebiet’, Archäologisches Korrespondenzblatt, 12: 11–13. Bogaard, A. (2002) ‘Questioning the relevance of shifting cultivation to Neolithic farming in the loess belt of western-central Europe: evidence from the Hambach Forest experiment’, Vegetation History and Archaeobotany, 11: 155–168. Bogaard, A. (2004) Neolithic Farming in Central Europe, London: Routledge. Bogaard, A. (2012) Plant Use and Crop Husbandry in an Early Neolithic Village: Vaihingen an der Enz, Baden-Württemberg, Frankfurter Archäologische Schriften, Bonn: Habelt-Verlag. Bogaard, A., Fraser, R. A., Heaton, T. H. E., Wallace, M., Vaiglova, P., Charles, M., Jones, G., Evershed, R. P., Styring, A. K., Andersen, N. H., Aerbogast, R. M., Bartosiewicz, L., Gardeisen, A., Kanstrup, M., Maier, U., Marinova, E., Ninov, L., Schäfer, M. and Stephan, E. (2013) ‘Crop manuring and intensive land management by Europe’s first farmers’, Proceedings of the National Academy of Sciences, 110(31): 12,589–12,594. Bogaard, A., Heaton, T. H. E., Poulton, P. and Merbach, I. (2007) ‘The impact of manuring on nitrogen isotope ratios in cereals: archaeological implications for reconstruction of diet and crop management practices’, Journal of Archaeological Science, 34: 335–343. Bogaard, A., Jones, G. and Charles, M. (2005) ‘The impact of crop processing on the reconstruction of crop sowing time and cultivation intensity, from archaeobotanical weed evidence’, Vegetation History and Archaeobotany, 14: 505–509. Bogaard, A., Jones, G., Charles, M. and Hodgson, J. G. (2001) ‘On the archaeobotanical inference of crop sowing time using the FIBS method’, Journal of Archaeological Science, 28: 1,171–1,183. Bogaard, A., Krause, R. and Strien, H.-C. (2011) ‘Towards a social geography of cultivation and plant use in an early farming community: Vaihingen an der Enz, south-west Germany’, Antiquity, 85: 395–416. Brown, P. (1978) Highland Peoples of New Guinea, Cambridge: Cambridge University Press. Butzer, K. W. (1972) Environment and Archaeology: An Ecological Approach to Prehistory, London: Methuen and Co. Lessons from modelling Neolithic farming practice 253

Charles, M., Bogaard, A., Jones, G., Hodgson, J. G. and Halstead, P. (2002) ‘Towards the archaeobotanical identification of intensive cereal cultivation: present-day ecological investigation in the mountains of Asturias, northwest Spain’, Vegetation History and Archaeobotany, 11: 133–142. Charles, M., Jones, G. and Hodgson, J. G. (1997) ‘FIBS in archaeobotany: functional interpretation of weed floras in relation to husbandry practices’, Journal of Archaeological Science, 24: 1,151–1,161. Colledge, S. and Conolly, J. (eds) (2007) The Origins and Spread of Domestic Plant in Southwest Asia and Europe, Walnut Creek, CA: Left Coast Press. Colledge, S., Conolly, J., Dobney, K., Manning, K. and Shennan, S. (eds) (2013) The Origins and Spread of Domestic Animals in Southwest Asia and Europe, Walnut Creek, CA: Left Coast Press. Fogelin, L. (2007) ‘Inference to the best explanation: a common and effective form of archaeological reasoning’, American Antiquity, 72: 603–625. Fraser, R., Bogaard, A., Heaton, T., Charles, M., Jones, G., Christensen, B. T., Halstead, P., Merbach, I., Poulton, P. R., Sparkes, D. and Styring, A. (2011) ‘Manuring and stable nitrogen isotope ratios in cereals and pulses: towards a new archaeobotanical approach to the inference of land use and dietary practices’, Journal of Archaeological Science, 38: 2,790–2,804. Goody, J. (1976) Production and Reproduction: A Comparative Study of the Domestic Domain, Cambridge: Cambridge University Press. Gregg, S. A. (1988) Foragers and Farmers: Population Interaction and Agricultural Expansion in Prehistoric Europe, Chicago, IL: University of Chicago Press. Halstead, P. (1995) ‘Plough and power: the economic and social significance of cultivation with the ox-drawn ard in the Mediterranean’, Bulletin on Sumerian Agriculture, 8: 11–22. Harman, G. H. (1965) ‘The inference to the best explanation’, Philosophical Review, 74: 88–95. Hodgson, J. G., Wilson, P. J., Hunt, R., Grime, J. P. and Thompson, K. (1999) ‘Allocating C-S-R plant functional types: a soft approach to a hard problem’, Oikos, 85: 282–296. Hüppe, J. and Hofmeister, H. (1990) ‘Syntaxonomische Fassung und Übersicht über die Ackerunkrautgesellschaften der Bundesrepublik Deutschland’, Berichte der Reinhardt Tüxen-Gesellschaft, 2: 61–81. Jones, G. (1992) ‘Weed phytosociology and crop husbandry: identifying a contrast between ancient and modern practice’, Review of Palaeobotany and Palynology, 73: 133–143. Jones, G. (2002) ‘Weed ecology as a method for the archaeobotanical recognition of crop husbandry practices’, Acta Palaeobotanica, 42: 185–193. Jones, G., Bogaard, A., Charles, M. and Hodgson, J. G. (2000) ‘Distinguishing the effects of agricultural practices relating to fertility and disturbance: a functional ecological approach in archaeobotany’, Journal of Archaeological Science, 27: 1,073–1,084. Jones, G., Bogaard, A., Halstead, P., Charles, M. and Smith, H. (1999) ‘Identifying the intensity of crop husbandry practices on the basis of weed floras’, Annual of the British School at Athens, 94: 167–189. Kelley, J. H. and Hanen, M. P. (1988) Archaeology and the Methodology of Science, Albuquerque, NM: University of New Mexico Press. Knörzer, K. H. (1971) ‘Urgeschichtliche Unkräuter im Rheinland: ein Beitrag zur Entstehungsgeschichte der Segetalgesellschaften’, Vegetatio, 23: 89–111. Lipton, P. (2004) Inference to the Best Explanation, 2nd edn, London: Routledge. Modderman, P. J. R. (1971) ‘Bandkeramiker und Wanderbauertum’, Archäologisches Korrespondenzblatt, 1: 7–9. Price, T. D. (2000) ‘Europe’s first farmers: an introduction’, in T. D. Price (ed.) Europe’s First Farmers, Cambridge: Cambridge University Press, 1–18. 254 Amy Bogaard

Reich, P. B., Walters, M. B. and Ellsworth, D. S. (1992) ‘Leaf life-span in relation to leaf, plant and stand characteristics among diverse ecosystems’, Ecological Monographs, 62: 365–392. Rösch, M., Ehrmann, O., Herrmann, L., Schulz, E., Bogenrieder, A., Goldammer, J. P., Hall, M., Page, H. and Schier, W. (2002) ‘An experimental approach to Neolithic shifting cultivation’, Vegetation History and Archaeobotany, 11: 143–154. Rowley-Conwy, P. (1981) ‘Slash and burn in the temperate European Neolithic’, in R. Mercer (ed.) Farming Practice in British Prehistory, Edinburgh: Edinburgh University Press, 85–96. Sherratt, A. (1980) ‘Water, soil and seasonality in early cereal cultivation’, World Archaeology, 2: 313–330. Thomas, J. (1999) Understanding the Neolithic, London: Routledge. Wallace, M., Jones, G., Charles, M., Fraser, R., Halstead, P., Heaton, T. H. E. and Bogaard, A. (2013) ‘Stable carbon isotope analysis as a direct means of inferring crop water status and water management practices’, World Archaeology, 45: 388–409. Wasylikowa, K. (1981) ‘The role of fossil weeds for the study of former agriculture’, Zeitschrift für Archäologie, 15: 11–23. Watson, P. J. and Kennedy, M. C. (1991) ‘The development of horticulture in the Eastern Woodlands of North America: women’s role’, in J. M Gero and M. W. Conkey (eds) Engendering Archaeology: Women and Prehistory, Oxford: Blackwell, 255–275. Westhoff, V. and van der Maarel, E. (1973) ‘The Braun-Blanquet approach’, in R. H. Whittaker (ed.) Handbook of Vegetation Science 5: Ordination and Classification of Communities, The Hague: Junk, 617–726. Whittle, A. (1996) Europe in the Neolithic, Cambridge: Cambridge University Press. Wylie, A. (1985) ‘The reaction against analogy’, in M. Schiffer (ed.) Advances in Archaeological Method and Theory 8, New York: Academic Press, 63–111. 14 EVIDENCE, ARCHAEOLOGY AND LAW

An initial exploration

Roger M. Thomas

At first blush (as lawyers sometimes put it), there might seem to be little or no connection between the subjects of archaeology and law. One is a research discipline, concerned with understanding the human past from material remains; the other occupies a central role in contemporary society, regulating conduct and, as far as possible, ensuring justice. A closer examination, however, reveals significant common ground between these two disciplines. This is because an important aspect of law is the use of evidence to try and establish what happened in the past. This is seen most clearly in contested trials, but comparable processes are also at work when investigations are undertaken in legal contexts (such as police or prosecutorial investigations when a crime has been committed). In this respect, archaeology and law share a good number of concepts and much vocabulary. Lawyers and archaeologists both talk about ‘evidence’, about inferences drawn from evidence, about ‘supposition’ and about ‘proving’ things. Notions of uncertainty, probability, dispute and debate are common currency in both dis- ciplines. The intellectual processes involved in archaeology and in law are also closely similar: formulation of hypotheses; collection of evidence; drawing inferences; marshalling arguments from multiple lines of evidence; testing and critical review of evidence and inferences; the complex interplay between evidence, hypothesis and conclusions. The concepts and terminology are very similar precisely because the processes have so much in common. Given all this, it is surprising that the parallels between archaeology and law as intellectual processes seem, certainly in archaeological circles, to have been little commented on previously. This lack of attention is all the more surprising for two reasons. First, forensic archaeology – that is, the collection of evidence, using archaeological techniques, as part of criminal investigations – has become a major specialism in its own right in recent years (e.g. Hunter et al. 1997). Second, a number of people have, at various points in their careers, worked or studied in both 256 Roger M. Thomas

fields: the distinguished British prehistorian Richard Bradley did his undergraduate degree in law; Giles Clarke, the excavator of the Lankhills Roman cemetery at Winchester (Clarke 1979), later changed career and become a barrister; and there are other examples of such movement between the two subjects. I do not think this is a coincidence. Because there are similarities in the intellectual processes involved, people who find intellectual satisfaction in archaeology, and have an aptitude for this kind of analytical thought, may find the same in relation to aspects of the law (and vice versa). The background to my own interest in the connections between the two subjects is that (as a practising archaeologist) I became interested in law, studied for a law degree at a mid-point in my career, and was called to the Bar in 2004. In the course of my legal studies, I came across Terence Anderson and William Twining’s excellent Analysis of Evidence (first edition: Anderson and Twining 1991; see now, Anderson, Schum and Twining 2005). This book is about evidence in legal contexts, but it makes reference to other disciplines that involve analysis of evidence and inference. Examples mentioned include military intelligence and archaeology. This was the only such reference to archaeology from a legal context of which I was aware at the time. I became intrigued by the possibility of applying legal methods of reasoning to archaeological evidence and interpretations, and that led eventually to this chapter. My debt to Professors Anderson and Twining in what follows is a substantial one.

Some general observations Let us consider in a little more detail some of the similarities between archaeological and legal thinking. Perhaps the first and most obvious point is that both are concerned with a past that actually happened – a ‘real past [which] exists’, as Shanks and Tilley (1987: 110) put it. I think the important point is that while, in both archaeological and legal contexts, the same evidence may be open to a number of different inter- pretations, there are some possibilities which are simply precluded by any reasonable interpretation of the evidence. This would seem a very obvious point to any lawyer, and it can be offered as an argument against the extreme relativism suggested by some post-processual archaeological theorists in the 1980s, described by Trigger (with reference to such authors as Shanks and Tilley 1987) as ‘hyper- relativism’ (Trigger 1989). While a particular body of archaeological evidence may support a variety of interpretations, it does not follow that it will support any interpretation. Both archaeologists and lawyers attempt to establish what happened in that ‘real past’ by collecting evidence, marshalling it, and using that evidence to construct a version of what happened. Arguments for that version are advanced, based on the evidence. Those arguments may be challenged, whether by offering an alternative interpretation of the evidence cited, by adducing different evidence, or by a combination of these. The original argument may then be defended, or the Evidence, archaeology and law: an initial exploration 257 alternative one countered, in the same way. We are all familiar with the television court-room drama, where opposing advocates each put forward their case, and then try to pick holes in the other side’s evidence or inferences. Every archaeologist has been involved in debates about archaeological interpretation – perhaps arguing on site about the date of a particular feature, with different people citing different pieces of evidence in support of their view, and seeking to argue away contrary evidence. The setting may be different – the excavation trench rather than the court-room – and the proceedings are obviously less formal, but the process is essentially the same. That is why, as outlined above, much vocabulary is shared between the two disciplines. Archaeologists and those involved in undertaking investigations for legal reasons (such as criminal investigations) also face a shared ‘philosophical’ problem in going about their work. It is this: in collecting evidence, do you pursue a particular hypothesis, which gives a clear focus to the investigation, but which may lead you off the scent if the hypothesis is wrong, or result in you collecting only that evidence which confirms your hypothesis? In other words, do you risk engaging in a self-fulfilling prophecy, in a process that could blind you to evidence that might lead in a different direction? Or do you try to collect evidence in a more neutral way, which may mean not really knowing what evidence to collect and what to ignore, because it is unclear what you are trying to find or what you are trying to prove? This problem is widely recognised in archaeology, where it is often characterised as the difference between ‘problem-oriented’ (or ‘hypothetico- deductive’) approaches and ‘inductive’ or empiricist ones. Both routes have their perils. In fact, the good investigator, or the good archaeologist, will employ both ‘hypothesis-testing’ and ‘inductive’ strategies. Archaeologists are certainly familiar with the process by which simply looking at and thinking about the evidence you already have can, in itself, suggest new lines of enquiry. In legal contexts, there is a concept of ‘abductive’ reasoning: ‘a creative process of using known data to generate hypotheses to be tested by further investigation’ (Anderson et al. 2005: 379). This is probably the approach that any good archaeologist adopts in practice; indeed, it is probably so much part of the archaeological investigative process that it generally takes place unremarked. Another problem shared by archaeology and law is that it is impossible to know with absolute certainty what actually happened in the past. The past did happen but it is, in practice, largely irrecoverable. Archaeologists, of course, are used to accepting this, and to recognising that many aspects of the past in which they might be interested are either completely irrecoverable, or can only be established with a very low degree of certainty. In the context of our comparison, the interesting point is that the law also recognises this ‘unknowability’ of the past, and addresses the problem in various ways, as will now be briefly explored. It is often popularly supposed that the purpose of a court case (a murder trial, say) is to find out ‘the truth’ about what happened. Strictly speaking, that is not correct. The purpose of a trial is to establish a particular proposition – that X murdered Y on a particular date, say – to a given ‘standard of proof ’. This 258 Roger M. Thomas proposition may be referred to as the ‘ultimate probandum’: the thing that, in the end, one wants to try and prove. Doing that may involve a series of ‘intermediate’ probanda: separate ‘lower-level’ propositions, each of which must be established if one is to prove the ultimate probandum. The proposition that X was close to the murder scene at the time of murder might be one such intermediate probandum: one of a series of strands, or steps, in an overall argument that leads to the ultimate probandum. In law, the ‘standard of proof ’ is the level of certainty to which the prosecution (or, in civil cases, the claimant) must prove their case in order to succeed. In English law, the standard of proof is different for criminal and civil trials. In criminal trials, the standard of proof is ‘beyond reasonable doubt’. That is close to, but stops short of claiming to be, ‘the absolute truth’. There is still some scope for doubt: ‘unreasonable’ doubt, but doubt nonetheless. In civil trials (for example, if you sue A for damages because you were injured in an accident which you believe was caused by A’s negligence) the standard of proof is ‘on the balance of probabilities’. In other words, is it more likely than not that the accident was A’s fault? That very clearly isn’t saying that the court’s decision is a statement of what actually happened in the past. It is a statement about what is more likely than not to have happened, based on the evidence presented in the case. Indeed, even ‘beyond reasonable doubt’ isn’t always clear-cut. Different people may come to opposing views, equally firmly and honestly held, on the basis of exactly the same evidence. This is recognised in jury trials, where, if the jury cannot reach a unanimous verdict, the judge may accept a majority verdict – say ten to two. The two people who took a different view from the rest are not regarded as being ‘wrong’; they have just formed a different perception of past events on the basis of the same evidence as that heard and seen by the majority. This is another acknowledgement of the fact that we cannot know the past with absolute certainty. Even after minute examination of detailed evidence from living witnesses about particular events that may only have happened a few weeks or months ago, uncertainties can remain.

Lawyers’ methods and archaeology In the preceding sections, I have drawn attention to what I think are some interesting parallels between legal and archaeological approaches, and between some of the problems facing lawyers and investigators in a legal context and those that are confronted by archaeologists. One can go on to ask: might legal techniques and approaches hold any interest for archaeologists? I would argue that they might. One way in which lawyers order their evidence and start to work up their case is by constructing a chronology: in essence, asking what happened when. Equally, for archaeologists, constructing a chronology (phasing a site, for example) is a basic archaeological task. Of course, the timescales will almost always be radically different for a legal as opposed to an archaeological instance, but the fact that developing a chronology is a basic task for lawyers in certain situations serves to Evidence, archaeology and law: an initial exploration 259 underline again how important evidence relating to time and sequence are for any investigation focused on the past. Issues of timing (‘was X at the crime scene by 7.15 a.m.?’) or of dating (‘is this radiocarbon date reliable?’) are often critical in both fields. Another tool of the lawyer is the so-called ‘case theory’ (Anderson et al. 2005: 153–158). This is the ‘story’ (or the argument) that the lawyer develops from the evidence: the version of events that the lawyer thinks will best accommodate and account for the available evidence. The lawyer in court has to do much more than simply present evidence; he or she has to present the evidence in a way that will persuade the judge or jury that his or her ‘story’ is more plausible than that presented by the other side. And the same goes for archaeology: the best archaeo- logical excavation reports, for example, use the evidence to lead to a set of clear and interesting conclusions. The dullest reports to read are those that do little more than simply describe features and finds.

Wigmore’s chart method In the preceding sections, I have drawn attention to how some aspects of legal methods and approaches find parallels in archaeological practice. I would now like to consider how one specific methodology from the field of law might be of value to archaeologists for managing their evidence, and for formulating and critically assessing their arguments and conclusions. Both lawyers and archaeologists are faced with a similar problem: how to use sometimes large bodies of evidence of varying kinds (‘mixed masses of evidence’, according to Anderson and Twining, Chapter 15 of this volume) to develop conclusions that are supported by that evidence. This involves marshalling the evidence (both for and against a proposition), defining how chains of inference link different pieces of evidence and lead to a particular conclusion, and assessing the strengths and weaknesses of each step in the argument. This may lead to a reconsideration of some conclusions, to a reassessment of the strength of one’s case, or to the identification of points on which additional evidence would be helpful (or to any or all of these in combination). One tool that can be used for this is the so-called ‘chart’ method, which was originally developed by the American lawyer John Henry Wigmore in the early twentieth century (e.g. Wigmore 1937). He was very keen on the idea that the principles of scientific inquiry could be applied to the analysis of disputed facts in a legal context. His work is accessibly summarised in Anderson et al. (2005). The ‘chart’ method is a way of analysing and depicting how different pieces of evidence and steps in an argument relate to each other. According to Anderson and Twining:

In a legal dispute, three questions must always be addressed: 1 What is the ultimate proposition that must be proved? 2 What are the data that are available as evidence? 260 Roger M. Thomas

3 What are the plausible relationships between these data and the ultimate proposition? (Anderson and Twining 1991: 159)

The chart method allows the way those three questions are addressed for a particular issue to be set out in a clear way. This can be especially useful for dealing with complex bodies of evidence, in which different lines of evidence and inference need to be analysed and combined to see whether (or how strongly) they support a particular conclusion. The aim of the chart (and the process of producing it) is to show how the ultimate proposition that the lawyer wants to prove (that A murdered B on 16 December, for example) can be demonstrated from all the relevant evidence available. The various pieces of evidence, inferences and probanda (things to be proved) are listed and numbered in a ‘key list’, and the relationships between each of them are then charted, using a standard set of symbols. The method is a way of ordering what may be a mass of different pieces of evidence, setting them in a logical relationship to each other, making clear what the inferential steps that link different pieces of evidence and argument are (thereby exposing any weaknesses or gaps in the argument) and showing the extent to which the evidence all converges (or not) on a particular conclusion. The charts are ‘dendritic’ in form, with the ultimate probandum at the apex and various lines of the argument progressively converging on this apex (see Figure 14.1). In a complex case, the main outlines of the argument can be shown in one chart, with the detail of some supporting arguments shown in separate (but linked) charts (e.g. Anderson 2003: Figures 4.3 and 4.4). The use of the chart(s) can be both ‘macroscopic’, marshalling the evidence and structuring the arguments for the case as a whole, and ‘microscopic’, identifying particularly crucial parts of the argument and subjecting them to the most careful dissection and analysis. Importantly, the process is a ‘dialectical’ one: the evidence and arguments both for and against a particular proposition are charted, so that the strength and weakness of particular parts of the case (and hence of the case as a whole) can be assessed. The process of compiling the key list and chart also requires thought and careful judgement at every stage. It is not a mechanical process; the chart method is an aid to thinking, not a substitute for it (see Anderson and Twining, Chapter 15 of this volume). In his original work, Wigmore used a large number of symbols, resulting in considerable complexity in the charts (Wigmore 1937). Anderson et al. (2005: 134–135) have produced a simplified version of Wigmore’s scheme, using fewer symbols (eight in total). These are as follows:

– the square for depicting testimonial assertions; – the circle for depicting circumstantial evidence or inferred propositions; > – the open angle to identify an argument that provides an alternative explanation for an inference proposed by the other side; – a vertical triangle to identify an argument that corroborates a proposed inference; Evidence, archaeology and law: an initial exploration 261

1

2 12

3 11 13 14 15 16

4 5 6 7 8 9 G 10

36

27 32 35

31

25 26 30

24 28 29 G 20 34

∞ 22 G 23 ∞∞∞22 22 22 ∞ 22 G 33

FIGURE 14.1 An example of a Wigmorean chart (from Anderson 2003: Figure 4.1). Note the ‘dendritic’ form, with different lines of argument converging on the ultimate probandum at the apex. The numbers denote different items in the ‘key list’ of evidence and propositions.

↑ – a line with a directed arrow to indicate the ‘direction’ of a proposed inferential relationship between a factum probans and a factum probandum, or between two factum probanda; – an infinity symbol to identify testimonial assertions or real evidence that the fact finders will hear or otherwise perceive with their other senses; ¶ – a paragraph symbol to identify facts the tribunal will judicially notice or otherwise accept without evidential support; and G – the letter ‘G’ to denote a generalisation that is likely to play a significant role in an argument in a case, but that is not a proposition that will be supported by evidence or that the tribunal will be formally asked to notice judicially. 262 Roger M. Thomas

In an archaeological context, ‘testimonial assertions’ could be read as covering two categories: first, verbal statements in field records, excavation reports or specialist reports; second, physical archaeological material (such as artefacts) and documents (such as photographs) – this second category would be referred to as ‘real evidence’ in law. An ‘inferred proposition’ could be a statement about chronology; for example, if Layer A is stated to lie on top of Layer B, then one would probably infer that Layer A was later than Layer B. In a legal context, the paragraph symbol (¶) is used to denote facts which are ‘judicially noticed’ (that is to say, so widely known and accepted that no evidence needs to be given on them – an example might be ‘the sun rises in the east’), or which are otherwise accepted without evidence. An archaeological equivalent might be the fact that the Roman period in England follows the Iron Age. Background generalisations (denoted by the symbol ‘G’) are very important to the chart, because of the principle that ‘every inferential step needs to be warranted by a background generalisation’ (see Anderson and Twining, Chapter 15 of this volume). In a legal context, for example, inferences drawn from alibi evidence (i.e. that X was elsewhere at the time of the murder) are based on the universal generalisation that a person cannot be in two different places at the same time. In an archaeological context, one such background generalisation might be the general stratigraphic principle that, if one layer lies on top of another, the overlying one will be later in date than the underlying one. This background generalisation would support (or ‘warrant’) the ‘Layers A and B’ example of an inference given above. In terms of wider archaeological or historical interpretation (see below), the question of ‘background generalisations’ is likely to be of great importance, and potentially contentious in itself. Wigmore thought that one of the sources of his ‘principles of proof’ was ‘general experience’: the ‘stock of beliefs’ or ‘common sense’, as some might call it, of a group or society (see Anderson and Twining, Chapter 15 of this volume). In looking at the behaviour of people in the distant past (as opposed to the properties of physical archaeological remains in the present), we need to accept that the ‘general experience’ of the people and societies that we seek to understand will have been very different from our own. This may require us to make explicit our views about how people may have thought and behaved, and how societies may have operated, in the past. This topic has, of course, been the subject of intense theoretical debate in archaeology in recent decades (e.g. Renfrew and Zubrow 1994; Dobres and Robb 2000).

Wigmore and Harris The archetypal context in which legal narratives are presented in their definitive form is in the court-room, as part of a trial, when it is necessary to put a clear and compelling account to the judge or jury. Anderson and Twining (see Chapter 15 of this volume, pp. 273 and 283) suggest that a Wigmorean chart is a useful device in preparing a case, but less suitable for presenting that case in court, where a more narrative approach is needed. In archaeology, the definitive account of research work is usually presented in a printed (or digital) publication. The final report on Evidence, archaeology and law: an initial exploration 263 an excavation is a classic example of this. I am not aware of any archaeological publications that use anything closely comparable to Wigmore’s chart method as a way of presenting evidence, arguments based on that evidence and the inferential links between these. It is of course possible that such charts are produced as working documents during the course of a project but are not presented in the published report, although I am not aware of this as a widespread practice in archaeological circles. One piece of archaeological apparatus does, though, have something in common with Wigmorean charts: the ‘Harris matrix’ (Harris 1989; see also Farid, Chapter 4 of this volume). This is a method for charting stratigraphic relationships (the relationships between different observed stratigraphic units or ‘contexts’). A particular excavation may result in hundreds or even thousands of discrete ‘contexts’ being observed and recorded. Each context has particular stratigraphic relationships with other contexts (A overlies B and C; D underlies B; C, E and F are actually part of the same deposit; and so on). The Harris matrix method allows these relationships to be systematically analysed and depicted in a logically correct way on a chart (or ‘matrix’) (see Figure 14.2).

1301

1301 1309

1309 1311

1331 1328 Coin of Victorinus 1328 1311 AD 269–71 1331

1353 1353

1360

1362 1360 1365

Animal bones 1368 1365 Complete pots 1367 1368 1368 Preserved timbers 1367 1362

01 metre

FIGURE 14.2 A simple example of an archaeological stratigraphic section and accompanying Harris matrix. Schematic east–west section through the fills of a middle Roman well at the town of Silchester. The numbers on the section represent individual archaeological ‘contexts’ (stratigraphic units), while the Harris matrix charts the stratigraphic relationships between contexts. (Drawing: Sarah Lucas; copyright: Amanda Clarke.) 264 Roger M. Thomas

The analysis on which the chart is based is founded on some general principles (which one might call ‘background generalisations’). For instance, if Layer A is on top of Layer B, then A is more recent than B (Harris’s ‘law of superposition’). The method involves using these general principles to order a mass of individual observations and logical inferences (for example, if A overlies B and B overlies C, then A is necessarily more recent than C). Other information, such as the dates of artefacts found in different contexts, can also be added to the chart to give a fuller understanding of the stratigraphic sequence, the chronology of the site and so on. The process of creating the matrix may highlight inconsistencies or contradictions in the evidence (e.g. the artefacts may indicate a different sequence than that suggested by the stratigraphy). Once identified, these issues can be explored further with a view to resolving them. Thus, the Harris matrix arguably performs a function similar to that of the Wigmorean chart: it is a device for marshalling a complex body of evidence, for establishing the relationships between different pieces of evidence, and for exposing flaws (discrepancies or circularities, for example) in the narrative thus created. This characterisation of the Harris matrix, which is specifically designed for dealing with one type of archaeological evidence (stratigraphic observations), serves to emphasise again the similarities between archaeological practice and some legal thinking and methods.

The Wigmorean chart in archaeology: a fictitious example To illustrate how Wigmorean charts might be used to represent and assess archaeological interpretations, let us consider a fictitious example. I have mentioned above the three central questions (what is the ultimate proposition that must be proved; what are the data that are available as evidence; what are the plausible relationships between these data and the ultimate proposition?). The chart method allows the way these questions have been addressed in a particular case to be set out and examined in a clear and systematic way. Say, for example, you have excavated a stone building somewhere in southern England. The dating of this building is not clear-cut, and is the subject of dispute. You believe it is Roman in date, and this is what you want to argue in the final report on the excavation. So, taking the three questions above and converting them to archaeological purposes:

(1) You want to show that the building is Roman in date (this is your ‘ultimate probandum’: the proposition which you ultimately want to prove).

(2) The available dating evidence consists of three strands (these are ‘the data available as evidence’) as follows: 1 a worn Roman coin minted in 300 AD from the foundation trench of the building; 2 sherds of pottery dated to around 350–400 AD from the floor of the building; Evidence, archaeology and law: an initial exploration 265

3 the fact that the building was sealed by a layer which contained twelfth- century AD pottery.

(3) The plausible relationships between the evidence and what you want to show (that the building is Roman in date) are as follows:

1.1 The coin was minted in 300 AD. 1.2 The coin is worn so must have been deposited well after the date it was minted. 1.3 The coin was securely sealed in the foundation trench. 1.4 The foundations must therefore have been laid well after 300 AD. 1.5 The building was therefore built well after 300 AD.

2.1 The pottery is dated to around 350–400 AD. 2.2 The pottery was found on the floor of the building. 2.3 The pottery is therefore contemporary with the use of the building. 2.4 The building was therefore in use in around 350–400 AD.

3.1 The pottery in the overlying layer dates to the twelfth century AD. 3.2 The layer must, therefore, have been deposited in the twelfth century AD at the earliest. 3.3 The building must, therefore, date to before the twelfth century AD. 3.4 Post-Roman stone buildings are unusual in this part of the country until the fourteenth century AD. 3.5 It is therefore likely that a pre-twelfth-century AD stone building in this part of the country is Roman.

This list of evidence and inferences constitutes the ‘key list’ (see above). Having produced this list, it can now be charted (see Figure 14.3). Here, I have used a ‘sub- divided’ numbering system, related to the separate lines of evidence, in contrast to the continuous sequences used by Anderson et al. (2005: 139 and Figure 5.2). These lines of evidence and argument all converge on a conclusion that the building is Roman in date (and late Roman at that, certainly after 300 AD and maybe well after that). Setting the evidence and the arguments out like this allows certain things to be seen: it exposes assumptions, and shows how an argument has been built up. In this case, the dating of the pottery is independent from the dating of the coin. The chart makes that clear. If, on the other hand, the pottery had been dated by the coin, then you could not validly use the pottery to date the building – that would be a circular argument. The chart also exposes another possible weakness in the argument. The layer overlying the building contains twelfth-century AD pottery. It is therefore true that this layer could not have been deposited before the twelfth century AD. Could it, though, have been deposited well after the twelfth century, and simply have included pottery that was already old when the layer was 266 Roger M. Thomas

Ultimate probandum: the building is late Roman in date

1.5 2.4 3.5

1.4 2.3 3.4

1.3 2.2 3.3

1.2 2.1 3.2

1.1 3.1

FIGURE 14.3 A Wigmorean chart for the fictitious archaeological example discussed in the text. (Copyright: author.) deposited? Equally, it identifies things that may need to be verified. Is it really true that post-Roman stone buildings were unusual in this part of the country until the fourteenth century AD, for example? What is the evidence for that statement, and how strong is it? By identifying such issues, the conclusion can be tested and perhaps modified or strengthened.

Chart analysis and approaches to archaeological publication Wigmorean charts may be useful for marshalling and assessing complex bodies of archaeological evidence and inference. Notwithstanding Anderson and Twining’s doubts about the usefulness of Wigmorean charts as a vehicle for presenting a case to others, at least in the context of a trial (see Chapter 15 of this volume, pp. 273 and 283), such an approach might also be relevant to the question of how excavations should be published. This has been a subject of long-running and continuing debate within the archaeological profession (e.g. Ancient Monuments Board 1975; Cunliffe 1983; Jones et al. 2003). In the past, much of that debate has centred on the question of whether ‘all the evidence’ should be presented in the report or whether simply a summary should be formally published, with the detailed supporting evidence remaining separate and being made available in other ways. The move towards digital (and online) publication makes it possible to move this debate forward. It is now perfectly practical and affordable to make all the evidence digitally available (e.g. Framework Evidence, archaeology and law: an initial exploration 267

Archaeology 2010), and potentially to integrate it fully with the main narrative account (by the use of hyperlinks). The scope for doing this throws the spotlight onto what has, in fact, always been the really important issue: making explicit the analytical links between archaeological evidence and the inferences which are drawn from it. Those links are the real focus of archaeological debate. How has a particular conclusion been reached? How strong is the evidence on which it is based? Does that evidence really support that conclusion? How strong or how tenuous is the link? Is a different inference supported as strongly, or more so, by the same piece of evidence? Does other evidence point in a different direction? The ‘chart’ method of legal analysis aims to allow those questions to be answered in relation to the evidence to be presented in a trial. It would be a very interesting experiment to try applying the same method to archaeological work – both by using it to analyse pieces of archaeological work that have already been published, and by using it as a tool in undertaking new archaeological projects and presenting the results. It would certainly be interesting to try and present the arguments of an archaeological report in the form of a Wigmorean-type key-list and chart, accompanied by a narrative commentary, rather than presenting everything as a continuous narrative account. An additional, and important, point is that the chart method forces hypotheses to be clearly defined and articulated. This parallels the archaeological injunction that projects should have clearly defined research objectives; framing these objectives formally as questions and devoting some care to the precise formulation of the questions can help to ensure that investigations have a clear focus from the start (Thomas 1999).

Some differences I have suggested that there are interesting and, in some respects, quite close similarities between legal and archaeological processes. There are also, I think, some differences, and it is worth looking briefly at these too. One significant difference concerns motivation, explanation and wider context. In large measure, the law is not much concerned with these questions. If the evidence (say, of an eyewitness) shows quite clearly that X committed a burglary, X’s particular reasons for committing that crime, and the wider social and economic context of the offence, are not particularly relevant as far as the question of guilt is concerned (although they may be relevant to the severity of the sentence that X receives). In archaeology, by contrast, there is often a strong desire to understand why things happened, as well as what happened – to explain what motivated people to behave in different ways at different times in the past, and why history unfolded in the way it did. The law, at least of the kind we have been talking about, tends to deal with single, individual events. It does not generally try to establish broad patterns of events or of human behaviour, and still less does it try to come up with explanations for these. 268 Roger M. Thomas

Thus, legal approaches may be especially useful for considering physical material evidence, such as excavation results. What is the date of that structure? What was its function? When did it go out of use? There is doubt, though, about how useful such approaches might be for answering broader questions, such as ‘what was the symbolic meaning of this structure?’ or ‘why was it built?’. This concern may be even more acute when it comes to considering broad questions about historical and social processes, such as ‘why was agriculture adopted?’ or ‘why did the economy of Roman Britain apparently collapse in the early fifth century AD?’. One could argue that legal-type approaches would be most useful in dealing with matters that are susceptible to ‘proof ’ from evidence, and least useful in relation to matters in which supposition or theoretical frameworks and positions play a large role. This may be because the ‘background generalisations’ that lie behind (for example) any Wigmorean analysis can differ so profoundly from one archaeological researcher to another. In some cases, the differing background generalisations involve deeply held theoretical views on the nature of human beings or the functioning of societies, such as the debate in the 1980s between ‘proces- sualist’ and ‘post-processualist’ theoretical positions (e.g. Shanks and Tilley 1987). These are not matters that are susceptible to being ‘proved’ from evidence. That said, there has been at least one attempt to apply Wigmorean analysis to a broad historical question. This is the work by Geller and Anderson on the proposition that the practice of cuneiform writing survived in Babylonia until the third century AD (Geller 2003a; Anderson 2003; Geller 2003b). Anderson, working in conjunction with Geller, undertook Wigmorean analysis of a lecture given by Geller on this subject. This work is discussed in more detail in the following chapter. Most relevant here is Anderson and Twining’s observation:

One strength of Wigmorean analysis is that it forces the clear articulation of ultimate and penultimate hypotheses and uses these as templates for relevance and anchors for an argument. If one has settled on a clear set of hypotheses, then one can sector and manage different strands in the argument by working downwards, that is, back from the ultimate probandum. (Anderson and Twining: Chapter 15 of this volume, p. 278)

Whether in relation to complex excavated evidence or for more broad-ranging historical inquiries and hypotheses, these sentiments are clearly relevant in an archaeological context. The systematic approach of the chart method, and its requirement for rigour and clarity, could thus make it as valuable to the archaeolo- gist as it is to the lawyer.

Conclusion This volume is about material culture as evidence, and about how archaeologists use this evidence. The concept of ‘evidence’ is, of course, also central to the practice of law. I have argued above that there are interesting parallels, worthwhile Evidence, archaeology and law: an initial exploration 269 to explore, between the disciplines of archaeology and law. Both are ‘evidential’ subjects, concerned with using evidence to try to establish what happened in the past. Some methods used in a legal context may also be applicable to the work of archaeologists. In particular, I suggest that Wigmorean ‘chart analysis’ may be of value to archaeologists, as a means of managing and ordering a ‘mixed mass of evidence’, building coherent arguments from this evidence and assessing the strengths and weaknesses of these arguments. This method may also offer pointers to new ways of presenting the results of archaeological excavations. My fictitious example above shows how this kind of approach can be applied to things that can be ‘proved’, such as function and dating of features, stratigraphic relationships and so on – in other words, to investigations of, and conclusions about, physical remains in the present. However, the general characteristics of Wigmorean analysis may assist in wider and more open-ended enquiries, such as those concerning processes of historical change. In neither case am I arguing that that this kind of approach will yield absolute truths (any more than it does in a legal context – and in practice of course the uncertainties are often much greater in an archaeological context). Nor am I claiming that it will always enable a decisive choice to be made between conflicting hypotheses based on the same evidence. It may, however, make it possible to reject some interpretations decisively. It is probably easy to overlook the importance of this. Some archaeological propositions, firmly believed at one time, have later been unequivocally rejected as a result of new evidence. The revision of the chronology of the British Neolithic following the advent of radiocarbon dating is an example (Darvill 2010: 19). New conclusions, when firmly rooted in the evidence, become part of the overall interpretative framework, and cease to be questioned or challenged. For that very reason, it is easy to forget that some interpretations are not open to reasonable challenge, simply because the evidence on which they are based is too strong. More generally, Anderson and Twining (Chapter 15 of this volume) stress the value of Wigmorean analysis for promoting clarity and rigour in the use of evidence and in drawing inferences. In the legal context, these qualities are achieved through searching examination and questioning at every stage. Any good investigator or lawyer will conduct just such a rigorous scrutiny of his or her own case before presenting it for scrutiny by others. Any good archaeologist should always wish to do the same.

Acknowledgements I am very grateful to Bob Chapman and Alison Wylie for inviting me to contribute to the Reading seminar in June 2010 from which this volume originates. A version of this chapter was originally presented at the Theoretical Archaeology Group (TAG) conference in Exeter in 2006. I am most grateful to Bob and Alison for their wise and careful guidance and great patience over this chapter. My greatest thanks, however, are to William Twining, whose work inspired me to think about 270 Roger M. Thomas the application of Wigmore to archaeology in the first place, and with whom I have had many very enjoyable and stimulating discussions in the course of producing this chapter.

References Ancient Monuments Board (England) (1975) Principles of Publication in Rescue Archaeology, London: Department of the Environment. Anderson, T. (2003) ‘Wigmore meets “the last wedge”’, in W. Twining and I. Hampsher- Monk (eds) Evidence and Inference in History and Law: Interdisciplinary Dialogues, Evanston, IL: Northwestern University Press, 140–215. Anderson, T. and Twining, W. (1991) Analysis of Evidence: How To Do Things With Facts, London: George Weidenfield and Nicolson. Anderson, T., Schum, D. and Twining, W. (2005) Analysis of Evidence, 2nd edn, Cambridge: Cambridge University Press. Clarke, G. (1979) Pre-Roman and Roman Winchester, Part 2: The Roman Cemetery at Lankhills, Oxford: Clarendon Press. Cunliffe, B. W. (1983) ‘The publication of archaeological excavations’, in Report of the Joint Working Party of the Council for British Archaeology and the Department of the Environment, London: Department of the Environment. Darvill, T. (2010) Prehistoric Britain, 2nd edn, Abingdon: Routledge. Dobres, M. A. and Robb, J. E. (eds) (2000) Agency in Archaeology, London: Routledge. Framework Archaeology (2010) Landscape Evolution in the Middle Thames Valley, Heathrow Terminal 5 Excavation, Volume 2, Oxford and Salisbury: Framework Archaeology. Geller, M. J. (2003a) ‘The last wedge’, in W. Twining and I. Hampsher-Monk (eds) Evidence and Inference in History and Law: Interdisciplinary Dialogues, Evanston, IL: Northwestern University Press, 122–139. Geller, M. J. (2003b) ‘Wigmorean analysis and the survival of cuneiform’, in W. Twining and I. Hampsher-Monk (eds) Evidence and Inference in History and Law: Interdisciplinary Dialogues, Evanston, IL: Northwestern University Press, 216–230. Harris, E. C. (1989) Principles of Archaeological Stratigraphy, 2nd edn, London: Academic Press. Hunter, J., Roberts, C. A. and Martin, A. (1997) Studies in Crime: An Introduction to Forensic Archaeology, London: Routledge. Jones, S., MacSween, A., Jeffrey, S., Morris, R. and Heyworth, M. (2003) ‘From the ground up’, The Publication of Archaeological Projects: A User Needs Survey, A Summary, York: Council for British Archaeology. Renfrew, C. and Zubrow, E. B. W (eds) (1994) The Ancient Mind: Elements of Cognitive Archaeology, Cambridge: Cambridge University Press. Shanks, M. and Tilley, C. (1987) Reconstructing Archaeology: Theory and Practice, Cambridge: Cambridge University Press. Thomas, R. (1999) ‘Writing archaeological reports: a question of questions?’, The Archaeologist, 34: 19. Trigger, B. G. (1989) ‘Hyperrelativism, responsibility and the social sciences’, Canadian Review of Sociology and Anthropology, 26(5): 776–797. Twining, W. and Hampsher-Monk, I. (2003) Evidence and Inference in History and Law: Interdisciplinary Dialogues, Evanston, IL: Northwestern University Press. Wigmore, J. H. (1937) The Science of Judicial Proof, Boston, MA: Little, Brown and Company. 15 LAW AND ARCHAEOLOGY

Modified Wigmorean Analysis

Terence J. Anderson and William Twining

In this chapter, we discuss the Wigmore chart method, which was originally developed to systematise arguments about disputed questions of fact, but which has also been used, in a modified form, to construct, represent and assess historical arguments in a range of other contexts. Building on the previous chapter, we consider here an application of Modified Wigmorean Analysis to an argument developed by Assyriologist Mark Geller for the claim that the Sumerian language survived in cuneiform for at least three centuries longer than historians had previously recognised (Anderson 2003; Geller (1994) 2003). The original exchange runs to almost 100 pages; the argument is summarised here, and the detailed analysis can be accessed at material-evidence.net.

Historical background Within the discipline of law, evidence is established as a distinct field, with specialised courses, treatises, conferences, journals, and scholars. For most of its history the study of evidence in law has been mainly focused on the Law of Evidence, viewed largely as a practical subject for lawyers involved in litigation. However, the great American evidence scholar John Henry Wigmore (1863–1943) maintained that the study of evidence in law consisted of two interrelated parts: the Law of Evidence (‘The Trial Rules’) and the Principles of Proof, ‘as found in “logic, psychology and general experience”’ (Wigmore (1913) 1937: Chapter 1). He argued that the latter subject was anterior to and more important than the Law of Evidence, but had been neglected by both scholars and practitioners. That is still true today, in that most treatises and students’ books still focus almost entirely on the Law of Evidence. Wigmore was the leading American scholar on the Law of Evidence for a generation, but his ideas on the Principles of Proof were not taken seriously for 272 Terence J. Anderson and William Twining over 60 years. Some dismissed his approach as ‘mere common sense’, others on the basis that it was too elaborate and time-consuming to be practical. Moreover, the idiosyncratic and folksy style of the Principles made it easy for academics to overlook it. Perhaps more important is the fact that the study of evidence in law has been narrowly assumed to be co-extensive with the Law of Evidence, and since the Principles was based on the proposition that ‘the law has no mandamus on the logical faculty’, i.e., that the law cannot dictate how people should reason (Thayer 1898: 314n), Wigmore’s ‘science’ was treated as not being a subject worthy of study. This was a serious mistake. Wigmore’s Principles was rediscovered in the 1970s, when a ‘new evidence scholarship’ began to emerge. The new movement was multi- disciplinary. To start with, it was mainly concerned with the application of statistical approaches, especially Bayes’s theorem, to evaluating the weight of evidence, and also with questions of witness psychology, especially identification evidence (Carson et al. 2007; Cutler 2008). However, in a third strand a few individuals, notably David Schum, Terry Anderson and William Twining, took Wigmore’s ‘logic of proof ’ seriously and modified it (see below) for use in teaching students, not only in law but also in engineering and intelligence analysis, a basic method for constructing and criticising arguments based on complex masses of data (Anderson et al. 2005; Schum (1994) 2001; Twining 1985, 2006; Twining and Hampsher- Monk 2003). The crucial assumption was that the method involved the application of basic principles of ordinary inferential reasoning in specific contexts. Roger Thomas (see Chapter 14 of this volume) has perceptively outlined some of the main overlaps and differences between legal and archaeological approaches to evidence. He has also introduced Wigmore’s chart method of analysis for mixed masses of evidence.1 Building on Thomas’s account, this section briefly explains the relationship between the Law of Evidence and the logic of proof and elaborates on some features of Modified Wigmorean Analysis (MWA), referring to an example of its application to a complex historical problem that requires the handling of multiple lines of evidence.

The Law of Evidence Wigmore’s teacher, James Bradley Thayer, articulated the basic view of the Anglo- American Law of Evidence that still represents the predominant common law orthodoxy, as follows:

1 That nothing is to be received which is not logically probative of some matter requiring to be proved; and 2 that everything which is thus probative should come in, unless a clear ground of policy of law excludes it. (Thayer 1898: 530, cf. 266)

This is significant for several reasons. First, the key concept is relevance, which is interpreted to mean ‘tends to support’ or ‘tends to negate’ a probandum (a fact Law and archaeology: Modified Wigmorean Analysis 273 to be proved). Second, the Law of Evidence mainly concerns admissibility of relevant evidence. In short, it can be restated as a single principle: ‘Exclude all irrelevant evidence; include all relevant evidence unless there is some special reason to exclude a particular proposition.’ Thus the Law of Evidence is a collection of disparate exceptions to ‘free proof ’, i.e., the application of ordinary (common-sense) inferential reasoning unconstrained by formal or artificial rules. Since Thayer, the scope and importance of these exceptions has been steadily eroded over time, so that in practice there are very few peremptory rules of exclusion left (Twining 2006: 210–218).2 Third, the exclusionary rules and discretions operate within a framework of argumentation based on propositional logic: to say that an item of evidence is not admissible means that it may not be used in this argument in this case. On this view, the logic of proof in law is anterior to the rules of evidence, but shares some of the key concepts, notably materiality, relevance, probative force and admissibility. (On these basic concepts, see Anderson et al. 2005: Chapter 3).3 The most important principle is exclude all irrelevant evidence.4 The Wigmorean approach to ‘proof ’ is based on two assumptions:

1 The principles of logic – abductive, inductive, and deductive – can be applied to any mass of data that constitutes evidence from which we wish to reconstruct particular past events or situations in order to resolve disputed or doubtful questions of fact. 2 The propositions established by these data and those proffered as possible inferred propositions can be articulated and can be organized and mar- shaled as arguments bearing upon the resolution of such disputed questions of fact. (Anderson et al. 2005: 98)

The chart method Traditionally, lawyers have used several analytic devices for organising data in preparing for trial. These include chronological tables, narratives, classification by source and elaborate ‘trial books’, which combine several methods (see Anderson et al. 2005: 317–324 and Chapter 4). Today, they can use sophisticated records management, indexing and support systems. Wigmore’s chart method is different from all of these in that it operates in a framework of argument. Generally speaking, these methods are complementary rather than mutually exclusive. A Wigmore chart is a chart or map of an argument, not a story. So what is the chart method? It is a rigorous technique for the construction, reconstruction and evaluation of arguments about questions of fact. It involves articulating every step in an argument, breaking down the argument into simple propositions, and then mapping or ‘charting’ all the relations between those propositions. It is most useful in dealing with complex arguments based on ‘mixed masses of evidence’. The foundation of the method is a simple form of informal propositional logic in the English empiricist tradition of Francis Bacon, Jeremy 274 Terence J. Anderson and William Twining

Bentham, John Stuart Mill, Alfred Sidgwick, Stephen Toulmin, Jonathan Cohen and Douglas Walton (e.g. Bentham 1827; Mill (1843) 1974; Toulmin 1964; Walton 1989). The logic is simple; the complexity lies in the materials to be analysed and in the relationships between the propositions in an extensive argument based on a mass of evidence. The logic is binary: every relevant proposition either tends to support or to negate directly or indirectly a single hypothesis or conclusion (the ultimate probandum). The technique is dialectical: the aim of the chart-maker should be to construct the most cogent argument possible for and against the ultimate conclusion and to relate the opposing arguments within a single coherent structure. The ‘chart method’ has two main aspects: first, it is an extremely useful technique for managing facts, marshalling evidence and structuring arguments in complex cases. This is achieved by working backwards from the ultimate conclusion (the top of the chart), and organising different phases of the argument into discrete, manageable sectors. The key lies in anchoring the arguments about the case as a whole in a clear standpoint, clear questions, and precisely formulated hypotheses or conclusions (see below). We shall refer to this aspect as ‘macroscopic analysis’. The second aspect of the method is as an analytical tool for identifying crucial or important phases in a complex argument and subjecting selected phases to rigorous detailed analysis. This ‘microscopic analysis’ is often perceived to be extremely arduous and time-consuming. However, the skilful analyst learns some basic principles of economy, by identifying crucial or important phases within an argument and focusing mainly on these. If a person’s jugular vein is severed, death is almost inevitable. In advocacy, ‘going for the jugular’ means concentrating one’s attack on the weakest point(s) in an opponent’s argument or, more positively, building up support for a proposition that, if established, will ensure success. Wigmorean analysis is particularly useful in identifying key propositions in an argument or, in investigation, in guiding enquiry by identifying potentially key propositions that are needed to build or destroy a case. The two main products of the chart method are a ‘key list’ of all the propositions used in the argument and an overview of their relations to each other presented in a single chart. Wigmore called his method ‘the only thorough and scientific method’, but this is misleading (Wigmore (1913) 1937: 858; Twining 1985: 125–142). The method is disciplined and systematic, but a chart is a picture of the chart-maker’s beliefs about what constitutes the strongest argument and its construc- tion at every stage involves judgement, selection, skill and analytical capacity. Because the method is binary and is compatible with modern computing tech- nology, people can easily be tempted into believing that the technique is mechanical and, given a sophisticated program, can be handed over to machines. This is a dangerous mistake. The chart method is an aid to thinking, it should not be envisaged as a substitute for thinking. Law and archaeology: Modified Wigmorean Analysis 275

Background generalisations One of the most important claims for the chart method is that it brings to the surface and forces articulation of exactly what is being argued. This is especially significant in respect of background generalisations. The model of reasoning is based on the premise that every inferential step needs to be warranted by a background generalisation (Wigmore’s ‘general experience’). For example, we infer that X happened from the fact that witness A said, ‘I saw X happen’. The warrant for this is that witnesses usually (but not always) tell the truth. For a universal generalisation, such as ‘no person can be in two different places at the same moment of time’, we can deduce an alibi (absence of opportunity) as a conclusion that follows necessarily from the premises. But that is exceptional, in that most of the background generalisations we must rely on are not universal; rather, they are qualified by terms such as ‘many’, ‘most’, ‘some’, ‘it is possible that’, ‘in 60 per cent of known cases of X, Y happens’ and so on. So nearly all inferential steps have to be expressed in terms of likelihoods or probabilities falling short of certainty. Of course, the reliability and precision of background generalisations vary considerably. Some generalisations are based on well-grounded science; some on evolving science; some on dubious, contested or junk science; some on practical expertise based on experience; some on everyday experience; some on isolated experiences; many on hearsay; some on speculation; some on faith; some on prejudice. Some generalisations are value judgements masquerading as factual propositions (e.g. sexist, racial, ageist prejudices, such as ‘all/most people of type X have characteristics Y’). Some generalisations are derived from stories, myths and fairy tales about particular events (Cinderella: if you want to win a prince, lose a shoe). All generalisations are fallible. Background generalisations are necessary and warranted in inferential reasoning, but many are dangerous for the above reasons (Anderson et al. 2005: Chapter 10; Anderson 2011). From where do we get our background generalisations? The orthodox answer is from a society’s (or group’s) ‘stock of knowledge’, or more accurately its ‘stock of beliefs’. As noted above, such ‘common sense’ includes well-founded, evidence- based generalisations, but it also includes many others that cover a wide spectrum of reliability, down to speculation, myth and prejudice. Reasoning in this way also has to postulate a cognitive consensus, the idea that most of these beliefs are shared (another dangerous assumption, especially in a multicultural society). In archaeology, there is a further complication, because inferential reasoning has to take account not only of researchers’ stock of beliefs but also of the subjects of enquiry. Background generalisations, especially when unstated, are often the most vulnerable part of an argument. Ordinary common-sense practical reasoning has to proceed on fragile foundations, but what is the alternative? (See Twining 2006: 120–125.) 276 Terence J. Anderson and William Twining

Modified Wigmorean Analysis (MWA) Since the 1970s, we have modified Wigmore’s chart method in several respects.5 First, we have suggested a quite flexible protocol in the form of an intellectual procedure for applying the method to a wide range of contexts. Anderson followed this protocol in ‘The Last Wedge’, an analysis of the historical reasoning involved in establishing how long the Sumerian language survived (Anderson 2003). In outline the procedure involves seven steps:

1 clarification of standpoint, purpose and role; 2 formulation of potential ultimate probandum or probanda; 3 formulation of potential penultimate probanda; 4 formulation of theory and themes of the case – choice of strategic ultimate, penultimate and intermediate probanda; 5 compilation of a key-list; 6 preparation of the chart(s);6 and 7 completion of the analysis.

Steps 1, 2, and 3 of this protocol merit some comment.

Clarification of standpoint MWA requires a careful and precise clarification of standpoint before starting on any exercise. This requires asking questions along the lines of: ‘who am I?’; ‘at what stage in what process am I?’; ‘what am I trying to do?’; ‘what question(s) am I trying to address?’; ‘what data is already or potentially available to me?’ Clarification of standpoint is particularly important in law because a wide range of participants in legal processes need to analyse and evaluate evidence from different standpoints, at different stages in the process and for different purposes7 (Anderson et al. 2005: 115–117 and 124–125). In a single case, a legal adviser, a police investigator, an advocate preparing for trial and thereafter in court, a juror and a parole board will have significantly different vantage points, access to information, roles, objectives and perspectives in relation to potential and actual evidence. Participants are significantly different from observers, who also need to be differentiated. In teaching MWA, we usually advise students to start by adopting the standpoint of a historian today making an ex post facto judgement about what happened (the event) on the basis of a finite body of data. In this context, a historian may define and pursue a single objective: the construction and evaluation of the arguments for and against a specific conclusion (e.g. does the evidence support beyond a reasonable doubt the conclusion that it was O. J. Simpson who murdered his wife?). This move filters out many of the technical factors that participants in legal processes have to take into account, such as roles, procedural requirements, trial strategy and tactics, client’s instructions and, of course, the Law of Evidence. Law and archaeology: Modified Wigmorean Analysis 277

Adopting the standpoint of a ‘historian’ cuts out most of these complicating ‘noise’ factors, opening the way to concentrating on the relationship of the data about a specific hypothesis in relation to a particular past event, thereby focusing attention on the logic. Of course, there are many kinds of historian, with differing vantage points, lenses, biases, objectives, methods, types and extent of data and specialised background knowledge. This means that a more nuanced clarification of standpoint may often be required for archaeologists than the simple standpoint adequate for law students who use this as a simplifying device. On the other hand, historians do not have a formal concept of materiality that defines the ultimate and penultimate probanda. A historian exploring an alleged miscarriage of justice in a case of murder has to employ the legal concept of murder (Twining and Weis 2006).

Hypotheses, material facts, ultimate and other probanda In empirical enquiries, it is important to distinguish conceptually between hypothesis formation and hypothesis testing, even though the distinction becomes blurred in practice. One standard way of expressing this is to say that hypotheses are abducted from particulars, but are tested by deduction and induction from evidence (Anderson et al. 2005: 55–60; Eco and Sebeok 1983). That is a bit simple, but will do for present purposes. In The Boscombe Valley Mystery, Sherlock Holmes stated: ‘You know my method. It is founded on the observance of trifles’ (Doyle 1936, as discussed in Anderson et al. 2005: 58–59). For example, in The Silver Blaze, the fact that the dog did not bark in the night led immediately to the suspicion that the groom was the guilty party (Doyle 1936). This did not ‘solve’ the case; rather it set up a plausible hypothesis that focused attention on one suspect and guided the investigation. In reasoning from ‘trifles’ to hypotheses, Holmes’s main method was abductive, even though Arthur Conan Doyle referred to these feats of reasoning as ‘deductions’. In police investigation and other enquiries abduction plays a large role, but contested trials are primarily concerned with testing already formulated hypotheses. In detective stories, cases are sometimes said to be ‘solved’ without sufficient evidence to satisfy legal standards of proof. In law, ‘material facts’ (see below, n. 3), ultimate, penultimate and interim probanda perform much the same function as hypotheses in historical enquiries. They provide templates of relevance. In this respect, the main difference from historical enquiries is that in legal proceedings it is the substantive law that prescribes what the proponent has to prove to succeed in this case. Typically, the ultimate probandum in a contested trial involves more than one proposition. For example, in murder the ultimate probandum may be ‘X murdered Y’; under English law the penultimate probanda will be the elements of the crime of murder – ‘Y is dead’, ‘the cause of death was an unlawful act’, ‘X committed the unlawful act’, ‘X committed the act with criminal intent’. The prosecution has to prove each of these beyond reasonable doubt. X may try to negate all or some of these probanda. If, for example, he puts forward an alibi, the argument can be expressed in the form: ‘it was not X who murdered Y, because he lacked opportunity; he was at place Z, 278 Terence J. Anderson and William Twining

25 miles away at the time of the murder; a person cannot be in two places at the same moment of time; his girlfriend G will testify that X was with her at Z at the time’. Obviously the prosecution may wish to challenge the credibility of both X and Z. They may adduce several reasons for disbelieving Z, including, for example, her close relationship with X and that she has several convictions for prostitution. And so on. The important point is that all of the operative propositions are hypotheses – X murdered Y, X had opportunity, X did not have opportunity, Z testifies that X did not have opportunity, and Z is a prostitute – are probanda: ultimate, penultimate and intermediate. They are all probanda, i.e., hypotheses needing to be supported or negated by evidence. The proposition that Z is a prostitute, apart from the question as to whether this affects her credibility, is indirectly relevant to the proposition that X murdered Y. Its relevance is dependent on other hypotheses. One strength of Wigmorean analysis is that it forces the clear articulation of ultimate and penultimate hypotheses and uses these as templates for relevance and anchors for an argument. If one has settled on a clear set of hypotheses, then one can sector and manage different strands in the argument by working downwards, that is, back from the ultimate probandum. In a contested trial at law, the ultimate and penultimate probanda are precisely prescribed by the time the fact-finder comes to decide the relevant evidence is identified, and the task of constructing a stable argument about the case as a whole is simplified. Without hypotheses the argument has no anchor. It may be objected that investigators are regularly warned not to jump to conclusions prematurely and not to focus on a single hypothesis too soon. For example, police investigators are warned of the dangers of ‘suspect-driven enquiries’, in which, having settled on a single suspect, all of their effort is focused on collecting evidence against that suspect. The enquiry should be kept open until a late stage. Hypotheses can be adjusted or refined as the investigation proceeds. There are standard procedures for listing possible (rival) suspects and shortening the list by eliminating as many as possible, one by one, for example by checking alibis early on. When a scenes of crime officer (SOCO) first comes to a potential crime scene, he or she is meant to keep an open mind and to look out for potentially significant ‘trifles’. However, the SOCO does not just ask: ‘What have we here?’ Rather, he or she uses a stock of questions to guide the enquiry and a set of generalisations (a specialised ‘stock of knowledge’) that helps to identify and interpret suspicious, unusual or puzzling trifles.8 Similar considerations apply to historians.

The last wedge: introduction and macroscopic analysis In his inaugural lecture at University College London, the Assyriologist Mark Geller ((1994) 2003) argued that the Sumerian language, as written in cuneiform, survived to the third century AD, rather than the second or first centuries BC. This dating was important for Geller, who expressed the view that the death of a language marked a ‘clearly definable historical boundary’, because ‘a civilization is Law and archaeology: Modified Wigmorean Analysis 279 lost when its successors no longer have access to its ideas or written legacy, and hence its language’ (Geller (1994) 2003: 123). The evidence for the ‘death’ of cuneiform included linguistic analysis (how late was the script?), the dating of tablets on which it was written, a small number of tablets that had cuneiform on one side and a Greek transliteration on the other (written by whom?) and inscriptions, literary sources and archaeological evidence relevant to the question of the length of time Mesopotamian rituals and temple complexes survived. In two cases, Geller produced ‘witnesses’ who were taught Babylonian language, customs and history (Iamblichus, see below) or adopted Babylonian practices (the Emperor Elegabalus) in Rome in the late second and third centuries AD. In addition to these ‘witnesses’, Geller ((1994) 2003: 131–132) also noted examples of ‘counterevidence’ (e.g. the visits of the Roman emperors Trajan and Septimus Severus to Babylon in 115 AD and 199 AD respectively, both recording that the city was in ruins and deserted). Anderson (2003) subjected Geller’s argument initially to macroscopic, and later to microscopic Modified Wigmorean Analyses, highlighting some weak points in that argument and leading Geller to respond (see below). In its published version, the exchange between Anderson and Geller covers over 100 pages. Presented below is a summary of Anderson’s critique, with a link to the original exchange, in order to illustrate how the method can be applied to this kind of material.

Standpoint In undertaking an analysis of Geller’s arguments, Anderson was an academic lawyer adept in using Wigmore’s chart method of analysis in legal contexts, but with no significant knowledge of Assyriology or Babylonian history. Anderson’s objectives were to determine whether the method could be applied to a body of Assyriological arguments and to discover what, if any, useful insights the effort might generate.

The ultimate probandum Unlike Geller or an archaeologist looking at the available evidence, Anderson did not have to formulate and reformulate hypotheses during the analysis. Because his database was the text of Geller’s argument, the source of the ultimate probandum in ‘The Last Wedge’ was internal, not external. The central proposition that Geller was advancing had to be derived from his text. As Anderson framed it, the ultimate probandum was:

1 The class of persons who could read cuneiform survived until the Sasanian conquest of Babylonia in 256 AD.

The penultimate probanda For ‘The Last Wedge’, the penultimate probanda were not defined by the ultimate probandum. Anderson was attempting to interpret and chart the arguments that 280 Terence J. Anderson and William Twining

Geller was making. Given his standpoint, Anderson’s task was to understand Geller’s theory and to organise and reconstruct the arguments in a way that would facilitate the analysis and an understanding of the resulting product. The penultimate probanda were propositions stating the main points developed in Geller’s argument, as Anderson understood them. In that context, the primary function of the formulation and strategic reformulation of these probanda was to organise and categorise Geller’s arguments in a form that facilitated the analysis and charting of the relationships between the evidential and inferential propositions to the ultimate proposition that Geller had set out to establish. Anderson defined the penultimate probanda as follows:

2 One or more active temple centres where cuneiform was read, studied and taught survived into the third century AD. 12 The Sassanian conquest in 256 AD was the probable cause of the termination of the class of persons who could read cuneiform.

(Note: the numbers directly above and below refer to the proposition numbers listed in Anderson 2003.) Anderson further subdivided the penultimate probanda into subordinate major points, penultimate probanda (see Figure 15.1):

3 The astronomical diary produced in 75 AD, the quantity of tablets produced in the Hellenistic period [circa 321 BC – 256 AD] as evidenced by the surviving fragments and the Greek script inscribed on the Graeco-Babyloniaca tablets provide physical evidence that supports the conclusion that the class of persons who could read cuneiform survived until the second or third century AD.

1

2 12 FIGURE 15.1 The beginnings of Anderson’s macroscopic Wigmorean Analysis, showing the relationship between the ultimate and penultimate probanda of Geller’s arguments. (Prepared by Terence L. Anderson.) 3 8 Law and archaeology: Modified Wigmorean Analysis 281

8 There is additional evidence in the documentary record left by Babylonians and non-Babylonians suggesting that the ability to read cuneiform survived until the third century AD (see Iamblichus analysis, below).

In all, Anderson listed 273 propositions that expressed the factual and inferential assertions of Geller’s argument. Based on details of all the propositions and the macroanalysis chart of Geller’s argument, Anderson argued that the propositions and chart ‘are sufficient to demonstrate that a method for depicting the logic involved in arguments about disputed questions of fact developed in one discipline (Law) could be applied to depicting arguments about disputed hypotheses in another (Assyriology)’ (Anderson 2003: 155). Anderson’s charts of Geller’s argu- ments based upon the physical evidence demonstrated to his satisfaction that these arguments strongly supported his conclusion. The charts demonstrated that his argument based on Iamblichus’ testimony were far weaker.

The Last Wedge: microscopic analysis and the Iamblichus argument In making his case for the later dating of the cuneiform script, Geller argued from one literary source as follows:

I call as my first witness one Iamblichus, a writer who appears in The Library of Photius. Iamblichus was a Syrian who composed erotic tales in Greek in approximately 200 AD. He was conversant with Babylonian magic, but he was educated as a Greek. The scholiast adds that Iamblichus ‘was a native Syrian, and not a Greek-Syrian, who knew Syriac, but also had a tutor who taught him the Babylonian language, as well as (Babylonian) customs and history’. The source in Photius goes on to add that Iamblichus later continued his career as a writer in Rome, residing there in the reign of Septimus Severus, circa 200 A.D. So here we have an independent record of a Syrian who already speaks the glossa Syriaca, or the local dialect of Aramaic, and wishes to learn the glossa Babyloniaca, which in my view is most likely to be Akkadian, and during the reign of Septimus Severus. But who was Iamblichus’ tutor? How and where did he find a teacher for cuneiform in, say, the mid-second century A.D.? The scholiast adds this important detail: Iamblichus’ teacher was a Babylonian who was taken prisoner by the Romans during Trajan’s campaign there, and later sold into slavery in Syria. The teacher had obviously got on well as a pedagogue or scribe in Syria, since he had been well-educated. The scholiast is clear on this point: Iamblichus’ teacher was skilled in ‘barbarian learning’, having served as a royal scribe in Babylon. ‘Barbarian learning’ never refers to Greek culture; Iamblichus’ teacher was a scholar who most likely knew cuneiform. He had served as a royal scribe, conversant in cuneiform in Babylon in the beginning of the second century A.D., before being captured by the Romans, and he 282 Terence J. Anderson and William Twining

remembered cuneiform well enough to teach Iamblichus later in the century. (Geller (1994) 2003: 129–130).

The principal source for the arguments made in that paragraph was a cluster of notes in the margin of a copy of Photius’s Bibliotheca, a book containing summaries written in the middle of the ninth century of some 280 books written in Greek (Encyclopedia Britannica 1910). One of these books, Babyloniaca or A Babylonian Story, was written by an author identified as Iamblichus. The notes were written in the margin of Photius’ summary of that story by an unknown scribe, a scholiast, who was preparing a copy of the Bibliotheca early in the tenth century.

Standpoint Here, Anderson wanted to take an argument that he perceived as ‘weak’ and use the chart method to determine and illustrate exactly why he thought this was the case and what its implications might be. For this analysis, he used three additional sources to enable him to overcome, to the extent feasible, the knowledge barriers to the analysis that the first exercise had revealed and illustrated. Otherwise, his standpoint was the same as stated for the macroanalysis.

Ultimate probanda, penultimate probanda and theory Either of two propositions could be viewed as the ultimate proposition that the arguments in the paragraph were intended to support:

62 Iamblichus knew and could read Akkadian cuneiform until the beginning of the third century AD. 51 The tutor knew and could read Akkadian cuneiform until the middle of second century AD.

The ambiguity made no difference for the purposes of this analysis because proposition 51 is a necessary step in the argument supporting proposition 62. As before, the penultimate probanda emerged from Anderson’s ‘working up’ from the evidential propositions in the text. The theory is clear, albeit ambiguous: the scholiast’s notes support an inference that cuneiform remained a ‘living language’ within Geller’s definition until either the middle of the second century or the beginning of the third century AD. Anderson next drafted a key-list of some 250 propositions and, following the MWA protocol, constructed four charts illustrating key phases of his argument. He then reflected on what this application of the method brought to our under- standing of a complex historical problem. He concluded that ‘the chart method of analysis [can] be usefully applied to evidence-based arguments in another discipline’ (Anderson 2003: 173). Mark Geller, in reply, acknowledged that ‘the use of Law and archaeology: Modified Wigmorean Analysis 283

Wigmorean analysis has been instructive in pointing out certain areas of weakness in the argument’, but robustly argued that he would still publish the article in its original form because it also ‘pointed to what may be possible, not provable’, and ‘to identify from where future evidence might be forthcoming’ (Geller (1994) 2003: 228–230). This is quite compatible with MWA, which claims to be able to identify what propositions, not yet backed by evidence, might be useful in strengthening or modifying an argument. Geller also produced some modifications of his argument in Wigmorean form. Pressure of space precludes presenting Anderson and Geller’s microscopic analyses, but these are available at material- evidence.net.

Conclusion MWA is very flexible and can be applied in a variety of contexts. It can be used to structure an argument through marshalling a complex body of existing data; in investigation it can also guide enquiry, by suggesting lines of questioning: ‘what proposition(s) would I like to have?’; ‘what if . . . ?’; ‘how can I test/boost/weaken the credibility of this source?’; ‘what information is needed to eliminate this particular suspect/hypothesis?’; ‘what is the weakest phase in the argument so far?’; and so on. The chart method is essentially a way of structuring arguments, so that a complex case can be made more manageable by breaking it up into sectors and identifying which evidence is potentially relevant to one or more sectors, while maintaining a coherent picture of the whole. The main limitation of the method is that it provides very little guidance on evaluating the strength of particular relations between propositions and of the argument as a whole (see Anderson et al. 2005: Chapters 8 and 9). It is very useful in identifying with precision the weak points in an argument and focusing attention on these. The process is useful for helping the chart-maker to clarify what exactly is being argued and how this can be strengthened; it is less useful for communicating or presenting an argument to an audience. It is therefore more useful in preparation for trial than for presentation at trial. This is where persuasive storytelling and other methods come into their own. Similar considera- tions apply to historical enquiries: marshalling and managing the data; constructing, eliminating and testing hypotheses; organising, refining and clarifying the argument are all important, usually necessary preliminaries to communicating to particular audiences. For the method to be of practical use in archaeology, it needs to be adapted and applied to concrete archaeological examples. This is something that can only be done by archaeologists.

Notes 1 Roger Thomas follows Wigmore and tradition by treating the contested jury trial as the paradigm case. In practice, such trials are wholly exceptional (less than 1 per cent of all criminal proceedings), but by that stage in the process the ultimate and penultimate 284 Terence J. Anderson and William Twining

probanda have been settled, and all or most of the evidence has been assembled. Modified Wigmorean Analysis (MWA) extends the application of the approach to other decisions in legal contexts involving inferential reasoning from evidence, including different stages of criminal investigation, decisions to prosecute, appeal and review of decisions at first instance, decisions by parole boards on questions of dangerousness, etc. One important characteristic of all of these enquiries is that they involve practical reasoning to arrive at and justify decisions, rather than conclusions. It still helps clarity of exposition to use the contested trial as the standard example. 2 We emphasise this here because most books on the Law of Evidence are not much use to archaeologists. Of course, a few rules – for instance, those governing confessions and improperly obtained evidence – are of considerable practical importance in some cases. 3 ‘Materiality’ relates to probanda (what has to be proved in a given case) and is governed by substantive law; for example, the law of murder prescribes what the prosecution has to prove in a case of murder. Material facts perform the same function as hypotheses, but generally speaking historians do not have a concept of materiality, i.e., their choice and formulation of hypotheses is not governed by formal rules. ‘Relevance’ refers to relations between propositions (A is relevant to B, i.e., A tends to support or negate B directly or indirectly); ‘weight’ or ‘probative force’ refers to evaluation of evidence, the strength or weakness of the relationship between particular propositions and between all of the evidence and the case as a whole. In Wigmore’s usage, ‘relevance’ and ‘weight’ are both governed by ‘logic and general experience’; lawyers tend to find it useful to keep conceptually distinct the question ‘is there any connection?’ (relevance) from the question ‘what is the strength of the connection?’ (probative force). ‘Admissibility’ refers to whether a given item (or class) of relevant evidence should be excluded (be treated as inadmissible), and is the only one of the four governed by the Law of Evidence. (See further Twining 2005: Chapter 6, ‘What is the Law of Evidence?’). 4 Mark Geller makes the point that in ancient history and archaeology it may be prudent to keep a note of seemingly irrelevant – but potentially relevant – items of evidence in case new data becomes available (Geller (1994) 2003: 230). That is fine and applies generally to investigations. But such evidence cannot properly be used as part of an argument based on available data. 5 So far as law is concerned, the main modifications to Wigmore’s approach have been: to simplify the symbols; to systematise the method through a flexible seven-step intellectual procedure; to emphasise clarification of standpoint (below); to treat story construction as complementary rather than as a rival method; to broaden the focus from contested jury trials to all phases of the legal process; and to extend its application to all enquiries into particular past events (e.g. history, genocide, archaeology) and some other enquiries (e.g. intelligence analysis, spotting insurance fraud). For archaeologists, the most significant advance may be the application of the approach to police investigation at different stages of enquiry (see Anderson et al. 2005: Chapter 2). 6 For detailed advice on charting, see Anderson et al. 2005: Chapter 5. 7 Theoretical aspects of the concept of standpoint and its significance are discussed, partly in relation to R. G. Collingwood’s theory of history, in Twining 2002: Chapters 2 and 3. 8 I am grateful to Richard Leary for this example.

References Anderson, T. (1999) ‘On generalizations I: a preliminary exploration’, South Texas Law Review, 40: 455. Anderson, T. (2003) ‘Wigmore meets “The Last Wedge”’, in W. Twining and I. Hampsher- Monk (eds) Evidence and Inference in History and Law, Evanston, IL: Northwestern University Press, Chapter 4. Law and archaeology: Modified Wigmorean Analysis 285

Anderson, T. (2011) ‘Generalisations and evidential reasoning’, in A. P. Dawid, W. Twining and M. Vasilaki (eds) Evidence, Interference, and Enquiry, London: British Academy, Proceedings of the British Academy, 171: Chapter 8. Anderson, T. and Twining, W. (1989) ‘Analysis of evidence’, in N. Gold, K. Mackie and W. Twining (eds) Learning Lawyers’ Skills, London: Butterworths, Chapter 8. Anderson, T., Schum, D. and Twining, W. (2005) Analysis of Evidence, 2nd edn, Cambridge: Cambridge University Press. Bentham, J. (1827) Rationale of Judicial Evidence, ed. J. S. Mill, 5 vols, London: Hunt and Clarke. Berger, P. and Luckman, T. (1967) The Social Construction of Reality, New York: Doubleday. Carson, D., Milne, R., Pakes, F., Shalev, K. and Shawyer, A. (eds) (2007) Applying Psychology to Criminal Justice, Chichester: Wiley. Cohen, L. J. (1977) The Probable and the Provable, Oxford: Oxford University Press. Cohen, L. J. (1980) ‘The Logic of Proof ’, Criminal Law Review, 2(1): 91–103. Cutler, B. L. (ed.) (2008) Encyclopedia of Psychology and Law, Thousand Oaks, CA: Sage Publications. Dawid, A. P. (2005) ‘Probabilities and proof ’, an online appendix to T. Anderson, D. Schum and W. Twining (eds) Analysis of Evidence, 2nd edn, Cambridge: Cambridge University Press. Dawid, A. P., Schum, D. and Hepler, A. (2011) ‘Inference networks: Bayes and Wigmore’, in P. Dawid, W. Twining and M. Vasilaki (eds) Evidence, Interference, and Enquiry, London: British Academy, Proceedings of the British Academy, 171: Chapter 5. Dawid, A. P., Twining, W. and Vasilaki, M. (eds) (2011) Evidence, Inference and Enquiry, London: British Academy, Proceedings of the British Academy, 171. Dingley, A. (1999) ‘The ballpoint case: a Wigmorean analysis’, in M. Malsch and J. H. Nijboer (eds) Complex Cases: Perspectives on the Netherlands Criminal Justice System, Amsterdam: Thela Thesis, Chapter 9. Doyle, A. C. (1936) The Complete Sherlock Holmes, Garden City, NY: Doubleday, Doran and Co. Eco, U. and Sebeok, T. (eds) (1983) The Sign of Three: Dupin, Holmes, Peirce, Indianapolis, IN: Indiana University Press. Gaskins, R. (1992) Burdens of Proof in Modern Discourse, New Haven, CT: Yale University Press. Geller, M. ((1994) 2003) ‘The Last Wedge’, in W. Twining and I. Hampsher-Monk (eds) Evidence and Inference in History and Law, Evanston, IL: Northwestern University Press, Chapter 3. Gold, N., Mackie, N. and Twining, W. (eds) (1989) Learning Lawyers’ Skills, London: Butterworth. Haack, S. (2009) Evidence and Inquiry, 2nd edn, Amherst, MA: Prometheus Books. Hanson, S. (2010) Legal Method, Skills and Reasoning, 3rd edn, London: Routledge-Cavendish. Kadane, J. and Schum, D. (1996) A Probabilistic Analysis of the Sacco-Vanzetti Evidence, New York: Wiley. Malsch, M. and Nijboer, J. F. (eds) (1999) Complex Cases: Perspectives on the Netherlands Criminal Justice System, Amsterdam: Thela Thesis. Memon, A., Vriej, A. and Bull, R. (eds) (2003) Psychology and Law: Truthfulness, Accuracy and Credibility, Chichester: Wiley. Mill, J. S. ((1843) 1974) Collected Works 7–8: System of Logic, Ratiocinative and Inductive, Toronto: University of Toronto Press. Palmer, A. (2010) Proof and Preparation for Trials, 2nd edn, Pyrmont, NSW: Thomson Reuters (Australia). 286 Terence J. Anderson and William Twining

Roberts, P. and Zuckerman, A. (2010) Criminal Evidence, 2nd edn, Oxford: Oxford University Press. Robertson, B. (1990) ‘John Henry Wigmore and Arthur Alan Thompson: an example of Wigmorean Analysis’, Victoria University Law Review, 20: 181–212. Schum, D. A. (1987) Evidence and Inference for the Intelligence Analyst, Lanham, MD: University Press of America. Schum, D. A. (1994/2001) Evidential Foundations of Probabilistic Reasoning, Chichester: Wiley and Evanston, IL: Northwestern University Press. Thayer, J. B. (1898) A Preliminary Treatise on Evidence at Common Law, Boston, MA: Little, Brown. Toulmin, S. (1964) The Uses of Argument, Cambridge: Cambridge University Press. Twining, W. (1985) Theories of Evidence: Bentham and Wigmore, London, Weidenfeld and Nicolson. Twining, W. (2002) The Great Juristic Bazaar: Jurists’ Texts and Lawyers’ Stories, Aldershot: Ashgate/Dartmouth. Twining, W. (2006) Rethinking Evidence, 2nd edn, Cambridge: Cambridge University Press. Twining, W. and Hampsher-Monk, I. (eds) (2003) Evidence and Inference in History and Law, Evanston, IL: Northwestern University Press. Twining, W. and Weis, R. (2006) ‘Reconstructing the truth about Edith Thompson: the Shakespearean and the Jurist’, in W. Twining Rethinking Evidence, Cambridge: Cambridge University Press, Chapter 12. Walton, D. (1989) Informal Logic: A Handbook for Critical Argumentation, New York: Cambridge University Press. Weis, R. (2001) Criminal Justice: The True Story of Edith Thompson, London: Penguin. Wigmore, J. H. (1908) ‘Review of C. Moore, A Treatise on Facts, or the Weight and Value of Evidence’, Illinois Law Review 3: 477–478. Wigmore, J. H. ((1913) 1937) The Principles of Judicial Proof, Boston, MA: Little, Brown (1937 sub. nom. The Science of Judicial Proof ). Wigmore, J. H. (1983) Treatise on Evidence, Vol. 1, rev. P. Tillers, Boston: Little, Brown. Wylie, A. (2011) ‘Critical distance: stabilising evidential claims in archeology’, in P. Dawid, W. Twining and M. Vasilaki (eds) Evidence, Interference, and Enquiry, London: British Academy, Proceedings of the British Academy, 171: Chapter 14. 16 TRADITIONAL KNOWLEDGE, ARCHAEOLOGICAL EVIDENCE, AND OTHER WAYS OF KNOWING

George Nicholas and Nola Markey

How do we come to identify and then interpret what can never be observed directly, namely the actions and associated material culture of past peoples? This question frames the archaeological endeavor, and it is a challenge of considerable magnitude, especially when those societies are often many thousands of years distant and likely representing behaviors and values that have no modern analog. As a profession, archaeologists are committed to the premise that the past is knowable, and indeed have had remarkable success in teasing apart the history of human achievement, evolution, and lifeways. But two unavoidable questions are: how do we know what we know about the archaeological record; and what types of evidence suffice for providing adequate “proof ” for our interpretations? Archaeologists have long depended upon ethnographic sources of information to help develop or test interpretations about past lifeways and events, but have had an uneasy alliance with oral histories and traditional knowledge.1 On the one hand, local knowledge is valued when it supports or supplements archaeological evidence. However, when the situation is reversed, when traditional knowledge is seen to challenge archaeological “truths,” its utility is questioned. This situation goes beyond the nature of “evidence” per se to reveal the challenges – and opportunities – that exist at the intersection of different ways of knowing (or different “know- ledge systems”). The emphasis on empiricism has also been manifested strongly in the broader realm of anthropology where achieving the status of a Kuhnian paradigm “became critical in the eyes of many social scientists, reinforcing their claims to do ‘real’ science (that being understood as a good thing)” (Darnell 2001: 3).2 In this chapter we explore the nature of archaeological knowledge, its creation and application, by examining the question of what constitutes “evidence” in the context of traditional use studies, ethnographic research, and oral histories produced or provided by Indigenous peoples. We do this by asking in what circumstances, 288 George Nicholas and Nola Markey and why, do archaeologists ignore local knowledge as evidence? And in what circumstances, and why, do Indigenous peoples ignore archaeological information and interpretations as evidence? We center our discussion on the intersection of archaeological evidence and oral traditions relating to the First Nations in western Canada, but also cite examples elsewhere in North America.3 The cases range from instances of concordance to contradiction with the archaeological record, yet each offers insights into evidential reasoning and, more broadly, the nature of knowledge. The tension that exists at the intersection of Western and Indigenous ways of knowing offers a rich but challenging arena in which to examine the scientific nature of traditional knowledge and oral histories.

Evidentiary reasoning Archaeologists and others may be skeptical . . . that this elder can as easily translate [rock art] images that may be a few hundred years old as those that may be thousands of years old. “How do you know this to be so?” may be answered by elders stating, “we don’t need to know how.” (Nicholas 2005: 82)

The interpretation of archaeological data has traditionally been informed by ethnoarchaeological research and experimentation, filtered through a largely Western set of values and expectations. This is very much an exercise in empirical observation, analogy, and inference. Yet despite some concerns about the adequacy of evidence or the means of interpretation, archaeology is no less inferential than other disciplines. In fact, to bridge that often deep chasm between the “stones and bones” of the past and the behaviors and activities they reflect, archaeologists have wrestled with the nature of inference and its shortcomings (see Wylie 2002) to the point that inference represents a strength of the discipline. At the same time, there has been considerable debate as to what constitutes “evidence,” a discussion that has been both shaped and propelled by the very public debate between scientists and humanists, manifested by the so-called science wars, by debates between processual vs postprocessual archaeologists, and by the often contentious relationship between archaeologists and descendant communities. In each case, science is promoted as real, objective, and the foundation for knowledge creation. Archaeology has long relied on ethnographic and ethnohistoric sources to reveal or provide proxies for the behaviors responsible for patterning found in the archaeological record (David and Kramer 2005). For example, studies of twentieth-century Nunamiut land use practices in Alaska have informed dis- cussions about hunter-gatherer logistical organization (e.g. Binford 1982), while insights gained from Ngatatjara informants in Western Australia provided possible explanations for unanticipated patterns of resource selection (e.g. Gould 1980: 154–155). Traditional knowledge, archaeological evidence, and other ways of knowing 289

In a practical sense, information from Indigenous peoples, whether framed as “ethnographic” or “local,” is comfortably viewed as one acceptable line of evidence used to support a variety of applications. These include:

• providing glimpses into the experiences of non-literate or marginalized peoples, such as enslaved Africans in eighteenth-century America (e.g. Taylor 1999), who have been conspicuously absent in written accounts; • correcting inaccurate or incomplete historical accounts of what transpired at the Battle of Greasy Grass (Little Big Horn) in 1876 by acknowledging the Lakota version (Marshall 2007); • adding new contributions to world “literature,” as illustrated by Robert Bringhurst’s (2011) translation and re-presentation of the major myths of the Haida of the Pacific Northwest Coast; • providing information in land claims or legal cases to establish a record of tenure or traditional practices that may have left no material remains (Miller 2011); • corroborating or extending scientific observations and models of climate change, such as using Inuit traditional knowledge (Qaujimajatuqangit) to explain changes in sea ice (e.g. Laidler 2006); and • yielding a new understanding of different ways of life, seen to add to an appreciation of humankind (e.g. Maybury-Lewis 1992).

Clearly, local knowledge has been used to great benefit by archaeologists in explaining archaeological patterning, constructing culture histories, and identifying broad cultural trends. However, what has been poorly articulated is the interface of such understandings with local or traditional knowledge.4 More to the point, the notion that such local knowledge can make substantial contributions to archaeology has often been challenged. Daryl Stump recommends that

. . . archaeologists should be wary of claims for the incorporation of indigenous knowledge in the absence of a methodology for doing so; should be skeptical of assertions that archaeology can unproblematically repair or restore lost local knowledge; should be critical regarding the archaeological visibility of local knowledge; and should be aware of the conflations of methodologies and nomenclatures that treat indigenous knowledge in the past and indigenous knowledge in the present as if these are susceptible to the same methods of inquiry. (Stump 2013: 285)

Likewise, Ronald Mason is cautious about oral traditions because: 1) the veracity of memory-based records must be regarded as “highly dubious”; 2) they reflect the “concerns of their contemporary reciters more than conditions and events of the past”; 3) they are emic and “largely impervious to external challenge”; and 4) some 290 George Nicholas and Nola Markey information is considered sacred or restricted, and not accessible to outsiders (Mason 2006: 11). From a scholarly perspective then, some see oral histories as a second-class type of information to be used with caution and in concert with other, more reliable sources. Rather than add to this already broad and rich discourse on oral tradition (e.g. Vansina 1985; Whiteley 2002), we shift now to the larger context – Indigenous traditional knowledge – of which oral histories form just one part.

Indigenous knowledge systems Indigenous or traditional knowledge (IK, TK) has been defined as the collective body of knowledge, experience, epistemology, ontology, traditions, and values held by societies that explain, record, and perpetuate their relationship with the world and all it contains (Aikenhead and Ogawa 2007; Bannister and Solomon 2009; Bruchac 2014; Smith 1999). Intersecting past, present, and future,

Indigenous knowledges are conveyed formally and informally among kin groups and communities through social encounters, oral traditions, ritual practices, and other activities. They include oral narratives that recount human histories, cosmological observations and modes of reckoning time, symbolic and decorative modes of communication, techniques for planting and harvesting, hunting and gathering skills, specialized understandings of local ecosystems, and the manufacture of specialized tools and technologies . . . (Bruchac 2014: 3,814)

This system/body of knowledge is based on individual and collective learned experiences and explanations, verified by elders, and conveyed through and preserved by oral traditions and other record keeping. In contrast to Western knowledge, which tends to be text-based, reductionist, hierarchical, and based on categorization, Native science (Cajete 1999) does not strive for a universal set of explanations but is particularistic in orientation, generally situated in specific environments (e.g. Inuit Qaujimajatuqangit). It is considered all- inclusive, as Mi’kmaw Stephen Augustine notes:

The fact that Native science is not fragmented into specialized compartments does not mean that it is not based on rational thinking, but that it is based on the belief that all things are connected and must be considered within the context of the interrelationship. (Augustine 1997: 1)

Notably, both Western and Indigenous knowledge are constantly verified through repetition and verification, inference and prediction, empirical observations and recognition of pattern events, and both forms of knowledge are always subject to improvement. Traditional knowledge, archaeological evidence, and other ways of knowing 291

What has become known as traditional ecological knowledge (TEK) refers to the integrated principles, practices, and beliefs that reveal and perpetuate the inter- connectedness of people, animals, plants, natural objects, supernatural entities, and environments (McGregor 2009; RCAP 1996). TEK may include harvesting techniques, stewardship of lands and resources, close origins with ancestral lands, medicinal and technological uses of plants, cosmology, protocols, and a philosophy of respect, all of which contribute to Indigenous worldview. Aboriginal peoples have accumulated knowledge and experience of the ecosystems in which they live, an understanding transmitted across generations and practiced daily. TEK has been prominent in the discourse around land and resource management and policy decision-making (Berkes et al. 2000), and represents an area of congruence with Western science. TEK and Western science are used as complementary tools in the environmental impact assessment (EIA) process, land use planning practices, and for resource management purposes.

Congruence or conflict? How does the nature of Indigenous and Western evidentiary reasoning affect knowledge and understanding of the past? Are these systems categorically antithetical? Or do they offer multiple points of entry into the complex and highly nuanced nature of both “heritage” and the “archaeological record,” and what each means in different cultural contexts? To explore these questions, we offer a series of examples that reveal different types (and degrees) of congruence and non- congruence or conflict between Indigenous and Western knowledge systems. These cases illustrate, singularly or in combination, three important considera- tions concerning the intersection of different knowledge systems – 1) selection of information for specific purposes; 2) consequences; and 3) opportunities – as discussed below.

Congruence The historical record helps to explain North American social settings as the product of traceable processes rather than as an expression of a timelessly rigid “ethnographic present.” Oral traditions and the archaeological record both reveal the workings of these processes, and both provide important know- ledge about the ancient past. (Echo-Hawk 2000: 288)

It is not uncommon to find connections between traditional knowledge systems and archaeological or historic records. For example, traditional resource manage- ment practices are a well documented, and indeed widely sought dimension of traditional knowledge (e.g. Anderson 2006; Turner 2005). Many elders are open to, indeed often enthusiastic about, sharing their knowledge with outsiders of past and present traditional knowledge about landscape management practices 292 George Nicholas and Nola Markey

(Davidson-Hunt 2003), fire ecology (Boyd 1999), lithic sourcing (Bernard et al. 2011), migrations (Whiteley 2002), tsunamis and other catastrophic events (Budwha 2002), or medicinal properties of plants (e.g. Bannister 2000). The following exam- ples reveal different types of congruence between oral accounts and archaeological information.

The travels of Glooscap The travels of Glooscap, a major figure in Abenaki oral history and worldview, are inscribed throughout the Mi’kmaw homeland of the Maritime provinces of eastern Canada. Like other transformers, Glooscap went about his business, creating many landscape features along the way:

In cutting open a beaver dam at Cape Chignecto, a small portion of the earth floated away, and Glooscap changed it into a moose and set his dogs on it. The moose took to the bay and made off; whereupon Glooscap turned him back into land, made him an island – the Isle of Holt – and fixed him there. He changed the dogs into rocks, which may be seen to this day . . . [at] . . . Ooteel. (Sable 2011: 162)

Today, many Mi’kmaw are reminded daily of Glooscap through place names, which refer to events that took place at specific locales. Some archaeologists and community members are actively seeking to demonstrate correlations between distribution of lithic sources and Mi’kmaw “mythical” sites. A community-based workshop held in Debert, Nova Scotia, in 2005 explored corroboration between archaeological, geological, and paleoclimatic sources and traditional Mi’kmaw knowledge, to better understand the Sa’qiwe’k L’nuk (Ancient People) (Bernard et al. 2011). There, elder Murdena Marshall stated that the Mi’kmaw needed to understand these other ways of knowing

so that our conclusions will not be so far apart. You believe in proof, we believe in oral traditions. We need to let those two work together to reach a common ground. . . . For me I did not need archaeology, anthropology or geology to prove who I am or where I came from. My oral traditions, which are part of my everyday life, told me so a long time ago. (Marshall 2011: 173–174) lo xwi (clam gardens) While the role of Indigenous peoples as resource managers is well documented on land (e.g. fire ecology, wetland canals), less apparent is management of coast resources, or mariculture. Long known to local residents, it was not until the 1990s that rock arrangements noted by a geomorphologist conducting a helicopter survey Traditional knowledge, archaeological evidence, and other ways of knowing 293 in coastal British Columbia brought outside attention to these features. These features were eventually identified as lo xwi,5 or “clam gardens” (Figure 16.1): inten- tionally rock-walled constructions that provide ideal habitat for shellfish (Lepofsky and Caldwell 2013). Kwakwaka’wakw Tom Sewid stated that, “My ancestor says: ‘We’ve made a productive beach because of working on it.’ It would become their property they’d guard with their lives because in a still harsh climate, clams could mean the difference between life and death” (cited in Williams 2006: 36). When this “discovery” was first reported to the provincial archaeology branch, the rock features were explained as likely fish traps: “What remains perplexing is that no clam researcher, anthropologist, or archaeologist appears to have seen or considered important the Waiatt clam terraces of the more than 350 sites [geomorphologist] Harper calculated he found . . .” (Williams 2006: 37). Today the identification of these features as clam gardens is widely accepted, and current studies are evaluating their effectiveness (Groesbeck et al. 2014).

Medicinal properties of balsamroot In the Plateau region of British Columbia, the Secwepemc depended heavily upon balsamroot (Balsamorhiza sagittata), “one of the most versatile food plants used in the region due to its edible roots, root crowns, young shoots, and seeds (Turner 1997: 93). Elder Mary Palmantier called it “the plant to end all plants” (cited in Bannister 2000: 41). Necessarily baked to make it edible, archaeological roasting pits attest to its use over thousands of years, and it is still harvested today. Balsamroot and an array of other plants were also used for their purported medicinal properties

FIGURE 16.1 Pre-contact clam garden, Waiatt Bay, Quadra Island, BC. (Photograph courtesy of Amy Groesbeck.) 294 George Nicholas and Nola Markey

TABLE 16.1 Secwepemc medicinal plants grouped according to medicinal use (after Bannister 2000: 24).

Plant use category Number of plant species

Tonics and general medicines 26 Purgatives, laxatives, emetics 4 Colds, coughs, tuberculosis, influenza, other respiratory ailments 34 Poultices, salves or washes for wounds, infections, burns, sores 40 Arthritis, rheumatism, muscular aches and pains 23 Kidney and urinary ailments 8 Sexually transmitted diseases 3 Eye medicines 13 Stomach and/or digestive tract medicines 16 Medicines for women (especially at childbirth) 12 Medicines for “cancer” (or disorders perceived as cancer) 2 Medicines for circulatory system 8 Miscellaneous or unspecified uses 53 to treat wounds, infections, urinary and digestive tract disorders, and other condi- tions (see Table 16.1). Kelly Bannister, with the consent of the Secwepemc Cultural Education Society, conducted a phytochemistry-based study of 68 plant species used by the Secwepemc. The study, which aimed to identify anti-microbial properties, determined that 88 percent of the plant species tested had antibacterial activity, 75 percent had antifungal activity, and 25 percent had antiviral activity (see Bannister 2000: ii). There is thus a strong correlation between local knowledge of plant properties and lab-based results.

Traditional use studies In Canada, traditional use studies (TUS) and traditional land use studies (TLU) are community-based research projects designed to record the long-term Indigenous land and resource use and ecological knowledge of an area (Markey 2001). Documenting archaeological and ecological knowledge on hunting, trapping, fishing, gathering, settlement, and traveling is important in identifying adverse impacts to Indigenous and treaty rights, and evaluating the impact of proposed development (e.g. logging, mining, hydroelectric) upon their lands, waters, and resources. TLU studies generally include literature and archival research, ethno- graphic interviews, and land use mapping (Brody 1981; Freeman 1976) to elicit information on traditional activities, the harvesting of resources, environmental knowledge, and personal use of traditional territories. TLU studies have commonly been conducted for the purposes of legal court cases, land use planning, or integrating traditional knowledge into environmental assessments, which has accelerated due to the increase in northern resource develop- ment and revised provincial and federal legislation (e.g. Canadian Environmental Traditional knowledge, archaeological evidence, and other ways of knowing 295

Assessment Act 2012), and landmark court cases addressing Aboriginal title and rights (e.g. Delgamuukw vs British Columbia 1997; Tsilhqot’in Nation vs British Columbia 2007) or consultation and accommodation matters (e.g. Haida Nation vs British Columbia 2004 and 2010; Rio Tinto Alcan Inc. vs Carrier Sekani Tribal Council 2010). Many TLU studies have been conducted in Canada over the past two decades, with Western scientific methods influencing how traditional knowledge is gathered, documented, and interpreted in archaeology and heritage management.

Non-congruence . . . much evidence must be discarded or discounted not because the wit- nesses are not decent, truthful persons but because their evidence fails to meet certain standards prescribed by law. (from Chief Justice Allan McEachren’s decision dismissing Delgamuukw title, cited in Cruikshank 1992: 26)

How do differences in knowledge systems outlined in the previous section play out regarding what constitutes evidence for archaeologists and for Indigenous peoples? Here we present five cases where there is a clear lack of congruence between Indigenous knowledge and the archaeological or historical records.

Stein River pictographs In the late 1980s, anthropologist Richard Daly interviewed Nlaka’pamux elder Annie York to interpret pictographs in the Stein River Valley of south-central British Columbia. Working from photographs, since Annie had never been there, she drew on knowledge acquired from elders to interpret many of the images (see Figure 16.2):

This is another hunting dream. The thing at the top looks like a hunting bag. The two sets of zigzags, that’s a way of showing the number of days of his dreaming. The top one is three-and-a-half days, and the lower one is four- and-a-half days. You count them from point to point. It shows two dreamings with a total of eight days no eating. He dreams the deer, see? Beside that is ancient animal tracks. Those old dreamed ones sometimes have those tails on them. Maybe that animal was dragging his feet, or it’s in snow weather. . . . In front of the deer is a deadfall log for snaring animals. (York et al. 1999: 102)

Most archaeologists are very uncomfortable with elders’ apparent knowledge of images painted hundreds, if not thousands of years ago, recalling Mason’s earlier cited concerns about oral histories. How can Annie York know what happened during a youth’s vision quest? Or is there a vocabulary of symbols known to the Nlaka’pamux that has carried through for generations? Conversely, what do 296 George Nicholas and Nola Markey

FIGURE 16.2 Pictograph “dreamings” interpreted by Annie York, Stein River Valley, BC (York et al. 1999: 102).

community members require as proof of their beliefs? What suffices as evidence here? For the Nlaka’pamux, oral histories include detailed accounts of the activities of such beings as Sun, Moon, Morning Star, and Coyote, as well as the first arrival of European explorers in the early nineteenth century (Wickwire 1995), and each is considered valid.

The Nataawaau bones In northern Québec, Cree oral histories refer to a battle that took place with invading Nataawaau warriors, possibly Iroquois (Denton 1997: 107). Working with members of the Nemaska Band, archaeologist David Denton examined an area elders had previously identified as where some members of the Nataawaau raiding party were killed by the Cree. However, testing revealed no human remains or artifacts, leading one elder to suggest that this was the wrong area: “Did we dig in the wrong place . . . Or were we in the correct location but still missed the evidence? Could the elder have been mistaken regarding the location? Or might the story have no factual basis?” (Denton 1997: 118). Traditional knowledge, archaeological evidence, and other ways of knowing 297

Certainly, Cree stories about invading groups have a strongly mythical quality. Yet, there is a long history of unfriendly relations, kidnapping of women, and raids by outsiders. In considering oral accounts and their correlation to historic accounts, Denton suggests that

. . . while oral traditions represent unbroken chains of knowledge of past events passed from generation to generation, they may also be responsive to material evidence from the past, and these remains may be interpreted (and re-interpreted) within the context of existing knowledge of local and broader regional history and within the framework of possibilities set out in local cultural tradition. (Denton 1997: 121)

Delgamuukw One of the most important legal title cases in Canada was Delgamuukw vs Britich Columbia, in which the Gitksan and Wet’suwet’en First Nations brought suit against the province of British Columbia. The initial 1991 ruling in favor of the province was overturned on appeal by the Supreme Court of Canada in 1997 (Culhane 1998). A crucial aspect of the case was the role of oral histories – the adaawk of the Gitksan, the kungax of the Wet’suwet’en – as a central line of evidence of tenure in the region. These oral histories embed and embody their traditions, principles of governance, social norms, economics, land use, resource management, and worldview. When questioned as to their accuracy, Chief Gyoluugyat answered, “In Gitksan law all Adaawks are true . . . How could the Adaawk be repeated if it’s not true to the Gitksan people? They got to be accurate and I know this from experience myself that they are accurate” (Napoleon 2005: 123). While there has been much written by anthropologists on the non-linear, fluid, multi-layered nature of oral accounts, what stands out in the cross-examination of the witnesses and Justice McEachern’s decision (see epigraph above), is the characterization of oral evidence as incomplete, untrustworthy, or corrupted by historical events (Cruikshank 1992). Notable is that anthropologists served as expert witness for both sides, indicating a difference of opinion professionally as to the integrity of oral histories as evidence, at least within the strict parameters of Western law (see Napoleon 2005). In their opening statement, the Gitksan and Wet’suwet’en hereditary chiefs acknowledged the difficulty for outsiders in fully understanding the nature of adaawks and kungax. “The challenge for this court is to hear this evidence in all its complexity, in all its elaboration, as the articulation of a way of looking at the world which pre- dates the Canadian Constitution by many thousands of years” (Culhane 1998: 118).

Navajo ethnogenesis There is strong disagreement today between Navajo accounts of their tenure in the Four Corners region of the American Southwest and archaeological and linguistic 298 George Nicholas and Nola Markey evidence that suggests they first appeared around AD 1500 (Wilshusen 2010). This issue has played out, sometimes very publicly, in the use of the term “Anasazi,” and who is represented by it, relative to Navajo claims of ethnogenesis. Warburton and Begay ask:

Are some Navajos related to the Anasazi? This question, which lies at the root of our discussion, is posed and mostly answered in the scientific realm. It is not a question for the Navajo, because the answer inheres in various traditional ceremonies and clan origin stories. (Warburton and Begay 2005: 537)

Traditional beliefs aside, bioarchaeological evidence of recent migration is compromised by centuries of intergroup relations: “The Navajo willingness to adopt the Pueblo lifestyle, as well as the Apache raiding and capturing of neighboring women, probably brought the B2a haplogroup, as well as others, into the Southern Athapaskan mtDNA gene pool . . .” (Achilli et al. 2013: 14,310). However, information derived from ancestral remains is difficult to assess as evidence because of the restricted ways it can be transmitted, i.e. only in the Navajo language, or only in ceremonial contexts (Warburton and Begay 2005: 535). Today Navajo history is being told in different ways because

. . . the archaeological literature about Navajo people was written by and for non-Navajo academicians. The resultant archaeological concept of “Navajo” does not align with Diné worldview and conceptions of self and history, thus reducing Diné oral history to the realm of “alternative” history. (Thompson and Towner, forthcoming: 1).

“We have always lived in pithouses . . .” In 1993, one of us (Nicholas) was conducting an archaeological excavation on the Kamloops Indian Reserve in British Columbia. A Secwepemc elder observed for a while, and then stated, “Our ancestors always lived in pithouses,” referring to the earth-covered, semi-subterranean structures (Figure 16.3) that were in use at the time of contact in the late eighteenth century. He went on to say, with absolute confidence, that salmon had always been the mainstay of their diet and economy. The lifestyle he was describing matched perfectly what is known of the historical period and the late prehistoric period. The archaeological record, however, suggests that this riverine-oriented way of life took form only in the last 5000 or so years, representing what is termed the Plateau Pithouse Tradition (Richards and Rousseau 1987). Yet, here is a case where archaeology provides a dynamic view of the past, one in which people responded to new social and environmental conditions by shifts in land use or new modes of social organization, while some contemporary Secwepemc adhere to a view of the past that is ahistorical, static, and unchanging. Traditional knowledge, archaeological evidence, and other ways of knowing 299

FIGURE 16.3 Replicated pithouse, Secwepemc Heritage Park, Kamloops, BC (side- entrance not traditional). (Photograph: George Nicholas.)

Discussion The cases presented above reflect an array of intersections between different forms of evidence derived from Indigenous and Western knowledge systems. For example, the four examples of congruence – Glooscap, lo xwi, balsamroot, and TUS – point to an overlap in what constitutes evidence, including story elements representing historical events and empirical observations. The instances of non- congruence are more challenging, as each reflects a different reason for rejecting evidence: Annie York’s interpretations of pictographs are not supported by standard research methods; archaeological investigation does not document the Cree encounter with the Nataawaau; the Delgamuukw oral history did not meet legal standards of evidence according to some scholars and the court; and both the Navajo and Secwepemc cases provide examples (albeit at different scales) where there is absolute disparity between oral history and the archaeological record. Three aspects of the evidentiary reasoning associated with these knowledge system intersections provide insight: factors relating to the selection of evidence; the consequences of evidence selection; and the opportunities that emerge from the tension found at the intersection of Indigenous and Western ways of knowing. 300 George Nicholas and Nola Markey

Choosing lines of evidence There is a strong element of selectivity in archaeologists’ acceptance of oral histories, and vice versa. Local knowledge has long been sought to support archaeo- logical interpretations, but when the situation is reversed, when traditional knowledge challenges archaeological “truths,” its utility is questioned. For example, archaeologists have no problem with discerning regional rock art styles, or identifying changes in style or motif over time; however, when an elder reads their meaning that knowledge is questioned. There is also an ironic aspect to elders sharing their knowledge: “Aboriginal plant medicine, once deemed ‘unscientific’ . . . and viewed as ‘at best fraudulent and at worst satanic’ . . . is now valued within the scientific search for new drugs and other plant products of benefit to humanity . . .” (Bannister 2000: 15). This is not a sign that traditional knowledge is now recognized for what it represents to the source community; rather, its monetary and medicinal value to outsiders has become apparent. Another issue is that it often takes considerable time before scholars are willing to accept or corroborate oral histories. Archaeological features may not be recognized at first, as is the case of the clam gardens, despite local knowledge. Likewise, the Lakota version of the Greasy Grass Battle was not considered reliable until validated by archaeological studies (Fox 1993); this is also true for information provided in the so-called slave narratives (e.g. Samford 2007; Singleton 1985). Examples of selectivity and delay are also evident in the use of archaeological evidence by Indigenous peoples. This includes hesitancy in seeking scientific lines of support (especially archaeological, bioarchaeological, and genetic evidence) to support land claims, tribal recognition cases, and other needs. Serious concerns over modern and ancient DNA research involving Indigenous peoples (Pullman and Nicholas 2012) are responsible.

Consequences of choice Teasing apart different lines of reasoning reveals both positive and negative conse- quences. For example, there is often a significant disparity when different knowledge systems come into contact, and this has negative consequences; as Cruikshank notes, “Participants predictably take their own meanings to represent ‘common sense,’ but the more powerful parties regularly build their meanings into legal arguments” (Cruikshank 2005: 11). In the case of Delgamuukw, discussions about the reliability of oral history are not restricted to academic discourse but had real consequences for the communities, questioning their ontology and epistemology. However, the Delgamuukw appeal has led to oral histories now having some weight in law. Traditional use studies provide one means of compiling and protecting local knowledge and otherwise protect community interests. However, the type of inventory-based research that typically ensues reduces traditional knowledge to symbols on a map, which is problematic because local knowledge is not so Traditional knowledge, archaeological evidence, and other ways of knowing 301 arbitrarily bound (Cruikshank 1998). Likewise, traditional ecological knowledge (TEK) is often less about knowledge content, and more about the processes used to gather and apply knowledge in the world. Most often, non-Indigenous researchers design the research, collect and catalog the data, and interpret the results due to little capacity in communities and the limited time allotted to provide the information (Markey 2010). Finally, TEK is usually translated from the original language to English, which affects the context. The clam gardens and balsamroot examples demonstrate strong acknowledge- ment of the scientific nature of local knowledge. The knowledge holders may find satisfaction that their traditional knowledge has been “validated,” and the source community and outsiders may benefit from health benefits or more efficient food production. But there may be negative consequences related to the commodifica- tion of that knowledge and other intellectual property issues (Bannister and Solomon 2009). In the case of Glooscap and other transformers, the correlation of events from “mythical” times with known archaeological sites supports traditional beliefs. Contrary to one elder’s conviction that ancestral land use had been unchanging, the Secwepemc (and other groups) acknowledge (and actively support) a more dynamic view of their past through archaeological evidence. However, when archaeologist’s use and interpretation of ethnographic sources is filtered through Western values and expectations (Sahlins 1968; Wobst 1978), the understanding of the truth that results may be very different from that of insiders.

Opportunities at the intersection One key attribute of Western science is developing and then testing hypotheses to ensure rigor and replicability in interpreting empirical observations or making predictions, criteria that are generally neither attributed to Indigenous knowledge nor totally absent from it. Whether or not Indigenous knowledge systems are truly compatible with archaeological reasoning, even contradictory lines of evidence have value. Employing them within a framework of multiple working hypotheses ensures consideration of a variety of predictive, interpretive, or explanatory possibilities not constrained by Western expectation or logic. As Hanen and Kelly note:

Because archaeologists can never use all of the available evidence, selectivity is inevitable. This selectivity may at times operate to restrict the field of hypotheses from which we wish to choose the best. At other times . . . it may cause us to eliminate alternative hypotheses prematurely, one of which may ultimately turn out to be the best. (Hanen and Kelly 1989: 17)

Following their lead, we suggest that using a method of multiple working hypotheses that incorporates traditional knowledge-derived evidence can push 302 George Nicholas and Nola Markey archaeologists toward unanticipated conclusions, which thus increases objectivity. Such a direction for responsible research is supported by Bruce Trigger’s contention that

By contradicting accepted interpretations of the past, these ideas stimulated research that tested both old and new ideas. . . . The greatest obstacle to making progress in archaeology is intellectual complacency. Without the ability to imagine alternative explanations, archaeology languishes. On the other hand, without the opportunity and determination to test ideas, imagination is of little value. (Trigger 2003: 190)

Grappling with contradictions between archaeological theory and practice may be valuable in the abstract (e.g. Wallace 2011), but there is a very real need for continued and expanded discussion as to what happens at the intersection of different knowledge systems (e.g. Sillitoe 2007). This can only benefit all parties by promoting good research practices that promote a more anthropological archaeology, a more satisfying engagement with descendant communities, and a better understanding of the utility and limitations of both lines of evidence.

Conclusions On what basis may we assess the strengths and weaknesses of each version of the past? Are we always obliged to adopt a relativist position in presenting parallel and equally valid stories about the past, or can some historical facts be derived by looking for congruence in diverse data sets? From whose perspective does the search for congruence take place? In what way may the different stories be seen as relating to different aspects of the same past, and hence be complementary rather than contradictory? (Denton 1997: 106)

How do we negotiate between the modes of evidentiary reasoning and inference discussed here? Assessing the points of intersection between different types of evidence is clearly important for archaeologists as an aid in interpreting lifeways and worldviews very different from their own. Despite a long-standing reliance in archaeology on using ethnohistoric sources to support interpretation and modeling, there is an enduring mistrust of traditional knowledge (e.g. oral vs “real” history), and a perception that it lacks credibility as evidence. Ignoring local knowledge may result in a selective reinterpretation of material culture or behaviors that differ significantly from that of the source communities. The inaccurate use of such information may subsequently jeopardize the integrity of archaeologically derived evidence. Indigenous epistemology and ontology, on the other hand, rely on evaluative criteria generally distinct from those of Western science, and a self-understanding Traditional knowledge, archaeological evidence, and other ways of knowing 303 of cultural tradition and history operates independently of outsiders’ knowledge. Evidence derived from oral histories thus needs to be used with care, a point urged not only by Ronald Mason but also Roger Echo-Hawk, who writes that

Scholars must stand their ground . . . when they are urged to accept origin stories as literal history. The intellectual legacy of academic scholarship requires that every presumption of historicity be subjected to critical exami- nation no matter how much it may anchor any specific cultural pattern. (Echo-Hawk 2000: 287)

That care also extends to using local knowledge responsibly, mindful of harm that can be caused to descendant communities (Nicholas and Wylie 2012). Local information, however manifested, provides an opportunity to evaluate archaeologically derived evidence, while also generating new questions that can inform our understanding of both past lifeways and contemporary heritage concerns. For their part, Indigenous peoples have the opportunity to examine archaeological information, to be as selective as archaeologists in determining what constitutes meaningful evidence for them, and to make their own judgments as to the goodness of fit between their ways of knowing and those that of Western science. Finally, heritage is as multifaceted as a wampum belt, which simultaneously represents to northeastern Native peoples a promissory note, a record of diplomacy, and a living entity. Writing of the Haudenosaunee agreement known as the “Two- Row Wampum,” Bruchac observes:

This diplomatic tradition invites two disparate cultures to coexist, side by side, without undue interference with one another [which] parallels the efforts to integrate Western Knowledge and Indigenous knowledge. (Bruchac 2014: 3,822)

Sometimes it will be useful and sometimes not, but both groups need to have that control over the interpretation.

Acknowledgements We thank Alison Wylie and Bob Chapman for their invitation to develop this chapter. Anne Salomon, Amy Groesbeck, Chris Arnett, John Welch, and Dana Lepofsky kindly provided useful resources and assistance.

Notes 1 While there are distinctions between them, we use “local knowledge,” “traditional knowledge,” and “Indigenous knowledge” more or less synonymously, but avoid “ethnographic knowledge” since it (intentionally or not) can be seen to “legitimize” or make more “reputable” oral histories by moving them from their original context into the Western realm of “ethnography.” 2 See additional discussions in Kuznar (2008) and Marks (2009). 304 George Nicholas and Nola Markey

3 There is a much broader literature (e.g. Johnson and Hunn 2010; Laudine 2009). 4 This is more than an academic point of interest since imposing a Western epistemological scheme upon Indigenous communities can have real consequences, especially when there is no need for an external retelling of their history (Nicholas 2005). 5 Williams refers to a Kwakwaka’wakwa song about building an lo xwi, a term that Franz Boas recorded without understanding its significance.

References Achilli, A., Perego, U. A., Lancioni, H., Olivieri, A., Gandini, F., Kashani, B. H., Battaglia, V., Grugni, V. N., Rogers, M. P., Herrera, R. J., Woodward, S. R., Labuda, D., Smith, D., Cybulski, J. S., Semino, O., Malhi, R. S., and Torroni, A. (2013) “Reconciling migration models to the Americas with the variation of North American native mitogenomes,” Proceedings of the National Academy of Science, 110(34): 14,308–14,313. Aikenhead, G. S. and Ogawa, M. (2007) “Indigenous knowledge and science revisited,” Cultural Studies of Science Education, 2: 539–620. Anderson, K. (2006) Tending the Wild: Native American Knowledge and the Management of California’s Natural Resources, Berkeley, CA: University of California Press. Augustine, S. J. (1997) “Traditional Aboriginal knowledge and science versus Occidental science,” Ottawa, ON: Biodiversity Convention Office of Environment Canada. Bannister, K. P. (2000) “Chemistry rooted in cultural knowledge: unearthing the links between antimicrobial properties and traditional knowledge in food and medicinal plant resources of the Secwepemc (Shuswap) Aboriginal Nation,” unpublished dissertation, University of British Columbia. Bannister, K. and Solomon, M. (2009) “Indigenous knowledges,” in A. Iriye and P.-Y. Saunier (eds) The Dictionary of Transnational History, Basingstoke: Palgrave Macmillan. Berkes, F., Colding, J., and Folke, C. (2000) “Rediscovery of traditional ecological knowledge as adaptive management,” Ecological Applications, 10(5): 1251–1262. Bernard, T., Rosenmeier, L. M., and Farrell, S. (eds) (2011) Ta’n Wetapeksi’k: Understanding from Where We Come, Truro, NS: Eastern Woodland Publishing. Binford, L. (1982) “The archaeology of place,” Journal of Anthropological Archaeology, 1: 5–31. Boyd, R. (ed.) (1999) Indians, Fire, and the Land in the Pacific Northwest, Corvallis, OR: Oregon State University Press. Bringhurst, R. (2011) A Story as Sharp as a Knife: The Classical Haida Mythtellers and Their World, Toronto, ON: Douglas and McIntyre. Brody, H. (1981) Maps and Dreams: Indians and the British Columbia Frontier, Vancouver, BC: Douglas and McIntyre. Bruchac, M. M. (2014) “Indigenous knowledge and traditional knowledge,” in Encyclopedia of Global Archaeology, New York: Springer. Budwha, R. (2002) “Correlations between catastrophic paleoenvironmental events and Native oral traditions of the Pacific Northwest Coast,” unpublished thesis, Simon Fraser University, Burnaby, BC, Canada. Cajete, G. (1999) Native Science: Natural Laws of Interdependence, Santa Fe, NM: Clear Light Books. Cruikshank, J. (2005) Do Glaciers Listen? Local Knowledge, Colonial Encounters, and Social Imagination, Vancouver, BC: UBC Press. Cruikshank, J. (1992) “Invention of anthropology in British Columbia’s Supreme Court: oral tradition as evidence in Delgamuukw v. BC,” BC Studies, 95: 25–42. Cruikshank, J. (1998) The Social Life of Stories: Narrative and Knowledge in the Yukon Territory, Vancouver, BC: University of British Columbia Press. Traditional knowledge, archaeological evidence, and other ways of knowing 305

Culhane, D. (1998) The Pleasure of the Crown: Anthropology, Law and First Nations, Vancouver, BC: Talon Books. Darnell, R. (2001) Invisible Genealogies: A History of Americanist Anthropology, Lincoln, NE: University of Nebraska Press. David, N. and Kramer, C. (2005) Ethnoarchaeology in Action, Cambridge: Cambridge University Press. Davidson-Hunt, I. (2003) “Indigenous lands management, cultural landscapes and Anishinaabe people of Shoal Lake, northwestern Ontario, Canada,” Environments, 31(1): 21–41. Denton, D. (1997) “Frenchman’s Island and the Nataawaau bones: archaeology and Cree tales of culture contacts,” in G. P. Nicholas and T. D. Andrews (eds) At a Crossroads: Archaeology and First Peoples in Canada, Burnaby, BC: Simon Fraser University Archaeology Press. Echo-Hawk, R. C. (2000) “Ancient history in the new world: integrating oral traditions and the archaeological record in deep time,” American Antiquity, 65(2): 267–290. Fox, R. A. Jr (1993) Archaeology, History, and Custer’s Last Battle: The Little Big Horn Reexamined, Norman, OK and London: University of Oklahoma Press. Freeman, M. (1976) Inuit Land-Use and Occupancy Project, Ottawa, ON: Thorn Press Ltd. Gould, R. A. (1980) Living Archaeology, Cambridge: Cambridge University Press. Groesbeck, A. A., Rowell, K., Lepofsky, D. and Salomon, A. K. (2014) “Ancient clam gardens increased shellfish production: adaptive strategies from the past can inform food security today,” PLOS ONE, 9(3): 1–13. Hanen, M. P. and Kelly, J. H. (1989) “Inference to the best explanation in archaeology,” in V. Pinsky and A. Wylie (eds) Critical Traditions in Contemporary Archaeology, Albuquerque, NM: University of New Mexico Press. Johnson, L. M., and Hunn, E. A. (eds) (2010) Landscape Ethnoecology: Concepts of Biotic and Physical Space, New York: Berghahn. Kuznar, L. A. (2008) Reclaiming a Scientific Anthropology, 2nd edn, Lanham: Altamira Press. Laidler, G. J. (2006) “Inuit and scientific perspectives on the relationship between sea ice and climate change: the ideal complement?,” Climate Change, 78: 407–444. Laudine, C. (2009) Aboriginal Environmental Knowledge: Rational Reverence, Farnham, UK: Ashgate. Lepofsky, D. and Caldwell, M. (2013) “Indigenous marine resource management on the Northwest coast of North America,” Ecological Processes, 2: 1–12. McGregor, D. (2009) “Living traditional knowledge and environmental practice in Ontario,” Journal of Canadian Studies, 43(3): 69–100. Markey, N. M. (2001) “Data ‘gathering dust’: an analysis of traditional use studies conducted within Aboriginal communities in British Columbia,” unpublished thesis, Simon Fraser University, Burnaby, BC, Canada. Markey, N. M. (2010) “The ‘other’ accidental archaeologist,” in G. P. Nicholas (ed.) Being and Becoming Indigenous Archaeologists, Walnut Creek, CA: Left Coast Press. Marks, Jonathan (2009) Why I Am Not a Scientist: Anthropology and Modern Knowledge, Berkeley, CA: University of California Press. Marshall, J. (2007) The Day the World Ended at Little Bighorn: A Lakota History, New York: Viking. Marshall, M. (2011) “L’nuit’si: Mi’kmaw tribal consciousness,” in T. Bernard, L. M. Rosenmeier and S. Farrell (eds) Ta’n Wetapeksi’k: Understanding From Where We Come, Truro, NS: Eastern Woodland Publishing. Mason, R. J. (2006) Inconstant Companions: Archaeology and North American Indian Oral Traditions, Tuscaloosa, AL: University of Alabama Press. 306 George Nicholas and Nola Markey

Maybury-Lewis, D. (1992) Millennium: Tribal Wisdom for the Modern World, New York: Viking. Miller, B. G. (2011) Oral History on Trial: Recognizing Aboriginal Narratives in the Courts, Vancouver, BC: University of British Columbia Press. Napoleon, V. (2005) “Delgamuukw: a legal straightjacket for oral histories?,” Canadian Journal of Law and Society, 20(2): 123–155. Nicholas, G. P. (2005) “The persistence of memory, the politics of desire: archaeological impacts on Aboriginal Peoples and their response,” in C. Smith and H. M. Wobst (eds) Decolonizing Archaeological Theory and Practice, New York: Routledge. Nicholas, G. P. and Wylie, A. (2012) “‘Do not do unto others . . .’: cultural misrecognition and the harms of appropriation in an open-source world,” in R. Coningham and G. Scarre (eds) Appropriating the Past: Philosophical Perspectives on Archaeological Practice, New York: Cambridge University Press. Pullman, D. and Nicholas, G. P. (2012) “Intellectual property and the ethical/legal status of human DNA: the (ir)relevance of context,” Inuit Studies, 35(1+2): 146–164. Richards, T. H. and Rousseau, M. K. (1987) Late Prehistoric Cultural Horizons on the Canadian Plateau, Burnaby, BC: Simon Fraser University Archaeology Press. Royal Commission on Aboriginal Peoples (RCAP) (1996) “From time immemorial: demographic profile,” Report of the Royal Commission on Aboriginal Peoples, Vol. 2, Ottawa: Canada Communications Group. Sable, T. (2011) “Legends as maps,” in T. Bernard, L. M. Rosenmeier, and S. Farrell (eds) Ta’n Wetapeksi’k: Understanding From Where We Come, Truro, NS: Eastern Woodland Publishing. Sahlins, M. (1968) “Notes on the original affluent society,” in R. B. Lee and I. Devore (eds) Man the Hunter, Chicago, IL: Aldine. Samford, J. (2007) Subfloor Pits and the Archaeology of Slavery in Colonial Virginia, Tuscaloosa, AL: University of Alabama Press. Sillitoe, P. (2007) “Local science vs global science: an overview,” in P. Sillitoe (ed.) Local Science vs Global Science: Approaches to Indigenous Knowledge, New York: Berghahn. Singleton, T. (1985) The Archaeology of Slavery and Plantation Life, Orlando, FL: Academic Press. Smith, L. T. (1999) Decolonizing Methodologies: Research and Indigenous Peoples, New York: Zed Books. Stump, D. (2013) “On applied archaeology, indigenous knowledge, and the usable past,” Current Anthropology, 54(3): 268–298. Taylor, Y. (1999) I Was Born a Slave: An Anthology of Classic Slave Narratives, Vol. 1: 1770–1849; Vol. 2: 1849–1866, Chicago, IL: Lawrence Hill. Thompson, K. F. and Towner, R. H. (forthcoming) “Navajo archaeology and a century of alternative Navajo history,” in B. J. Mills and S. Fowles (eds) Oxford Handbook of Southwest Archaeology, Oxford: Oxford University Press. Trigger, B. G. (2003) Artifacts and Ideas: Essays in Archaeology, New Brunswick, NJ: Transaction Publishers. Turner, N. J. (1997) Food Plants of Interior First Peoples, Royal British Columbia Museum Handbook, Vancouver, BC: University of British Columbia Press. Turner, N. J. (2005) The Earth’s Blanket: Traditional Teachings For Sustainable Living, Vancouver, BC: Douglas and McIntyre. Vansina, J. (1985) Oral Tradition as History, Madison, WI: University of Wisconsin Press. Wallace, S. (2011) Contradictions of Archaeological Theory: Engaging Critical Realism and Archaeological Theory, London: Routledge. Warburton, M. and Begay, R. M. (2005) “An exploration of Navajo-Anasazi relationships,” Ethnohistory, 52: 533–561. Traditional knowledge, archaeological evidence, and other ways of knowing 307

Whiteley, P. M. (2002) “Archaeology and oral tradition: the scientific importance of dialogue,” American Antiquity, 67: 405–415. Wickwire, W. C. (1995) “To see ourselves as the other’s other: Nlaka’pamux contact narratives,” Canadian Historical Review, 75(1): 1–20. Williams, J. (2006) Clam Gardens: Aboriginal Matriculture on Canada’s West Coast, Vancouver, BC: New Star Books. Wilshusen, R. H. (2010) “The Dine at the edge of history: Navajo ethnogenesis in the northern Southwest 1500–1750,” in L. L. Scheiber and M. D. Mitchell (eds) Across a Great Divide: Continuity and Change in Native North American Societies, 1400–1900, Tucson, AZ: University of Arizona Press. Wobst, H. M. (1978) “The archaeo-ethnology of hunter-gatherers or the tyranny of the ethnographic record in archaeology,” American Antiquity, 43(2): 303–309. Wylie, A. (2002) “The interpretive dilemma,” in Thinking from Things: Essays in the Philosophy of Archaeology, Berkeley, CA: University of California Press. York, A., Arnett, C. and Daly, R. (1999) They Write their Dreams on the Rock Forever: Rock Writings in the Stein River Valley of British Columbia, Vancouver, BC: Talon Books.

Legislation and litigation Canadian Environmental Assessment Act, 2012 Delgamuukw vs British Columbia [1997] 3 SCR 1010 Tsilhqot’in Nation vs British Columbia, 2007 BSCS 1700 Tsilhqot’in Nation vs British Columbia, 2014 SCC 44 William vs British Columbia, 2013 SCC 34986 Haida Nation vs British Columbia (Minister of Forests), [2004] 3 SCR 511 Rio Tinto Alcan Inc. vs Carrier Sekani Tribal Council, 2010 SCC 43 This page intentionally left blank PART IV Broader perspectives: material culture as object and evidence This page intentionally left blank 17 EVIDENCE OF WHAT?

On the possibilities of archaeological interpretation

Gavin Lucas

Prologue Viđey is a small island just off Reykjavík, the capital city of Iceland. Recently, I ran an archaeological project exploring the material histories connected to an abandoned village at the eastern end of the island. The village is not old; it was founded at the start of the early twentieth century by a joint stock company and abandoned in the early 1940s, its existence bound up with the rise of industrialized fishing in Iceland (Lucas and Hreiđarsdóttir 2012). My interest in the village was connected to a larger, collaborative project studying aspects of modern ruins (see www.ruinmemories.org; also Olsen and Pétursdóttir 2014) but one of the key themes of my study was looking at how objects came together and dispersed at this location (Lucas 2014). A pivotal question that arose on encountering the site was: where did everything go? The place is decidedly eerie; not by any stretch of the imagination is it a classic ghost town, where the rooms survive with things strewn about as if abandoned all of a sudden. Iceland has plenty of these places, especially in the countryside. No, this site is in some ways more archaeological, in that for the most part, all that remain are stone and concrete foundations, cracking and covered with moss, slowly being covered by vegetation (Figure 17.1). A site halfway to be being buried – neither quite a conventional archaeological site, nor a classic ruin. It is precisely this quality though that provoked the question: where did everything go? There was clearly something here once, but very little is visible any more. If the site had been a ghost town, I might have asked where all the people had gone, but I probably would not have been concerned with the things. If the site had been completely buried and required extensive excavation to uncover it, I might have simply shown surprise (or disappointment) at how much (or little) was buried under the ground. But the question that prompted me to explore the processes of how the village came together and fell apart – materially speaking – does have a flip side, which I have not considered until now. In reflecting on where 312 Gavin Lucas

FIGURE 17.1 One of the abandoned houses at the fishing village on Viđey, Iceland. (Photograph: author.) everything went, I neglected to ask what is perhaps the most important archaeo- logical question of all: why have some things lingered? Why have these things remained behind? This question was not uppermost in my thoughts, simply because so little seemed to be there (a fact of course quickly disproved when I began small-scale excavation and uncovered large amounts of discarded material culture), and it is a question perhaps naive in the extreme. But sometimes the most naive questions lead to the most interesting conclusions. At the risk of sounding presumptious, it is this question I wish to take up now because I feel it goes to the heart of another question, that which is captured in the title of this chapter: what exactly, is the archaeological record evidence of ? By the end of this chapter, I hope the relation between these two questions will become apparent.

A critique of archaeological reason It is very unfashionable to talk about the limits of archaeological interpretation or inference these days. In fact you can see I hesitated to use it in my title and instead opted for a more positive phrasing. This is the legacy of half a century of theoretical archaeology, by which I mean the emergence of an explicit and self-defined sub- field of archaeology that was spawned by the New Archaeology in the 1960s and continued into processualism, post-processualism and whatever -ism defines now. Back in the beginning, the promise was about expanding the limits of what we can say about the past (Binford 1962, 1968); optimism ruled and the only negative voices were the old guard, warning the new generation about the limits of over- Evidence of what? On the possibilities of archaeological interpretation 313 interpreting the evidence, exemplified in Christopher Hawkes’s famous epistemic ladder (Hawkes 1954; see also Smith 1955). And just when the new generation began to pull back a little, show more caution, along came another lot who threw the ladder away altogether and opened up an endless vista of interpretive possi- bilities (Hodder 1986). Of course, I have grossly caricatured this disciplinary history, but I think there remains a strong sense that it is bad form or reactionary to set limits on archaeological interpretation, even if we are fully aware that it is not a case of ‘anything goes’. Setting limits remains a purely negative or empty threat: there are limits but we do not try to define them, we just know they exist. Perhaps we should call it a negative epistemology. Of course there are good reasons for this stance: critiques of capitalist reason (Adorno and Horkheimer 1997; Horkheimer 2004), Western reason (Spivak 1999; Young 1990) or masculine reason (Haraway 1988; Hartsock 1983; Harding 1986; Wylie 1992) made within critical theory, postcolonialism or feminist epistemology and science studies have all made us highly sensitive to prescriptive epistemologies that spell out what is and what is not to count as proper knowledge. Nonetheless, such critiques do not so much undermine the epistemic status of limits as make them more mobile; think of Spivak’s strategic essentialism or feminist standpoint theory. In this chapter, I want to explore what you might call a critique of archaeo- logical reason – that is, I want to try and understand something of the nature of the limits of interpretation in archaeology. To do that, I suggest we need to understand something about the nature of archaeological reality as we apprehend it – i.e. the archaeological record. This is of course a very conventional Kantian undertaking insofar as it posits a close link between the nature of archaeological phenomena and the nature of archaeological reason. But it is also a very particular and contingent project because it makes no claims to universality but rather applies to a very specific standpoint: the archaeological. I want to begin by looking at how archaeological reality – the archaeological record in effect – is conventionally articulated. To this end, I will explore two very different metaphors and corresponding conceptions of the archaeological record: the fragment and the relic. I will suggest that whereas the former tends to view the archaeological record as incomplete, fragmentary, the latter views it in terms of survival or continuity. Now there exist archaeological discourses around both of these ideas, but I would suggest that the former has tended to dominate and especially been most influential in discussions about archaeological interpreta- tion and knowledge. In ending, I offer some thoughts on how to re-imagine the archaeological record in such a way that it makes us re-think the nature of archaeological interpretation. We will ask, what is it really evidence of ?

The fragment Various words have been used to characterize the nature of the archaeological record: remains, vestiges, ruins, relics, traces and fragments. The list could go on, and while many terms are still current, some have a decidedly antiquarian ring to 314 Gavin Lucas them (e.g. vestiges). By looking at the words we use to define the stuff that makes up the archaeological record in this generic sense, I think one can discern an underlying tension between two broader concepts, that of time and totality. By totality I mean a sense of completeness or wholeness, while by time I mean specifically a sense of the contemporary versus the antiquated, anachronistic or out of date. Thus if one were to classify these terms on axes of time and totality, one can begin to see how they differ. Some terms, like ‘fragment’, can have almost no temporal connotation but simply be linked to the notion of totality, whereas others, like ‘relic’ are much more temporal and evoke a sense of belonging to another time or age regardless of their completeness. Many other terms, of course, fall in between – the ruin, for example, implies both incompleteness and something from the past. Now the point of this observation is really to direct our attention to how these two framing concepts of totality and contemporaneity have shaped the way we interpret the archaeological record. It seems to me that it is the former (i.e. totality) that has powerfully dominated our framing of the nature of the record, while the latter (contemporaneity) has remained very much suppressed – though it was not always that way. Let me give some brief historical examples to illustrate this (see also Lucas 2012 for a more extended discussion). The idea of the archaeological record as incomplete or fragmentary runs through archaeological discourse since the inception of the discipline (also see Burtsröm 2013). Its origins can perhaps be traced to Romanticism and art history through the concept of the ruin (Ginsberg 2004), where the fragment acted as the stimulus for reflection on the missing parts, the former whole, whether from an epistemological perspective (Vismann 2001) or an aesthetic one (Patrik 1986). The power of the fragment lies in evoking or underlining an absence or loss and, as such, initiates a discourse or poetics on the relation between the fragment and the missing parts or between the fragment and the whole. Now of course this is inevitably entwined with time, but by emphasizing the fragmentary, time is coloured in a very specific way – it is conceived as a gradient of diminution or reduction. The whole pot reduced to a few sherds, the pristine sword corroded to a rusty, flaky stick of iron. Indeed, precisely because these are strongly material properties we can directly experience, it is perhaps unsurprising that they are metaphorically extended to define the archaeological record in toto. Following Lakoff and Johnson, we might say that potsherds and rusty iron objects are good metaphors to live by, as archaeologists (Lakoff and Johnson 1980). From them, we can generate an abstract conception of archaeological remains in terms of incompleteness. Arguably, the idea of incompleteness has very much dominated our conception of the archaeological record since the late nineteenth century. Classic statements of this conception emerged in the middle of the twentieth century: formation theory, sampling theories and source criticism, all of which stressed the loss of evidence or data in the form of a failure to survive – whether in terms of organic materials or ephemeral phenomena like behaviour itself. Moreover, even Schiffer’s transformation theory, which was ostensibly a corrective to naive, reductionist or entropic views (e.g. Schiffer 1987: 8–11), can still be seen as part of this same Evidence of what? On the possibilities of archaeological interpretation 315 conception. Materially, formation theory may not inherently be reductionist, but in terms of information lost through re-shuffling or re-ordering of this material, then it remains an essentially entropic theory from the point of view of the systemic context – even if new patterning or information is created in the process. However, it is not only the nature of the archaeological record that is conceived primarily through the notion of the fragment, it is also the nature of archaeological practice. The idea of practice being guided by an ethic of totality is also evident throughout the history of the discipline through the ideal of the total record. Beginning in the late nineteenth century with German classical archaeology – which stressed the need for a full recovery of all remains, not just the fine art specimens (Schmidt 2002; Marchand 1996), and the aspiration for complete collections of material for analysis – these ideals carried through into the late twentieth century and its idealization of the archive objective: that the archive should stand in for the site itself (Roskams 2001). Of course, the ideal of the total record was always ambiguous and perhaps seen as an impossible ideal (e.g. Carver 1990). Yet this ambiguity played to the heart of the issue of incompleteness because it was always about mediating the relation of frag- ment to whole (also see Burström 2013). Whether it was a question of how many specimens were necessary for a full classification or corpus or how big a sample was necessary for reliable interpretation, the issue was always about gaining control over the problem of the fragment. Indeed, the abstraction and generalization of this problem even extended to the nature of archaeological interpretation itself, both in terms of an inductive epistemology or the hermeneutic question of nomothetic versus idiographic science (Lyman and O’Brien 2004). How many facts did you need to prove or test a hypothesis? How much generalization is possible in a humanist science like archaeology? These larger epistemic issues tie directly into the question of the fragment, because how you respond to them affects how you judge issues such as the relation between types and examples, or target and sample populations, let alone your very model of interpretation or explanation. I have inevitably condensed what is a very rich and complex discourse here (see Lucas 2012), but it hopefully suffices to illustrate my point about the important ways in which the archaeological record has been conceptualized through the idea of totality – or incompleteness. Yet, this is not the only way in which we have thought about archaeological remains; besides the question of incompleteness, there is also that of time and particularly, the concept of contemporaneity. It is to that I now want to turn.

The relic While the fragment is a powerful metaphor for the archaeological record, it lacks a temporal dimension. This is not to suggest it cannot be aligned with time, only that time becomes secondary and therefore coloured by the idea of incompleteness, hence the dominant connotation of entropy or loss. If one foregrounds time, however, then arguably a very different conception emerges about archaeological 316 Gavin Lucas remains: that of the remainder, the left-over or left-behind; residues, relics, vestiges – terms that conjure up a rather different image of the archaeological record. Such remains may indeed even be complete or whole; what defines them is not so much their incompleteness but their untimeliness; they are from another era, anachronisms, survivals. This conception of the archaeological record in fact may have been the more dominant one in the nineteenth century, only superceded in the twentieth by the idea of the fragment. Such a view is very clearly expressed when one considers the close relationship that existed between material remains and what came to be called survivals, that is extant customs or practices which no longer have any obvious purpose, like vestigial organs. In fact, up to the middle of the nineteenth century, the word antiquities was commonly used to refer to both; it was only with the crystallization and separation of academic disciplines in the last quarter of the nineteenth century that such differences also started to emerge, and with it, a loss of sensibility to issues such as the temporality of the archaeological record (see also Shanks 2012). It is very instructive to read Tylor’s work on survivals in this regard, because he quite explicitly saw survivals and relics (terms Tylor himself used) as the twin pillars of the study of the past. What is important about this temporal dimension to the archaeological record is that it is very much defined in relation to the contemporary; what makes relics and survivals significant is their non-contemporariness. Now this is quite interesting because this temporal quality of the archaeological record has nothing to do with absolute age or relative chronology – it is quite simply about their status vis-à-vis the time of the scientist. I find this articulation through the concept of the contemporary very provocative because it immediately resonates with the oft- repeated statement that the archaeological record is a contemporary phenomenon (e.g. Binford 1983). It is the juxtaposition of the two views that is particularly troubling; for on the one hand, archaeological remains are contemporary – they exist here and now and they are what we engage with, even if our objective is to wrestle out a story about the distant past. At the same time, they are also of or from the past – and not simply in the sense of having persisted or endured in a neutral, chronological sense, but they have endured into a time in which they do not seem to belong. That is the crux of the concept of survival – an anachronism, an object or custom out of its proper time. Of course, this raises the question of what constitutes the ‘proper’ time for an object, and no doubt archaeology itself has helped to create the perceived proper temporal order for things, but more generally this is probably the legacy of modernist thinking insofar as the trope of modernity defines the very possibility of something being untimely. In considering the legacy of this temporal conception of the archaeological record, it is actually hard to discern its theoretical impact, which is not the case for its counterpart, the fragment (but see Harrison 2011). Indeed, it has been very much superceded by the concept of incompleteness, especially in the key method- ological theories like formation theory, source criticism or sampling theory. Let me take one example to demonstrate how this shift has occurred: the concept of the palimpsest. This idea was first proposed in the 1920s by British landscape archaeolo- Evidence of what? On the possibilities of archaeological interpretation 317 gists like O. G. S. Crawford, who likened the landscape to a re-used manuscript: it was composed of all kinds of elements from different eras. Looking at the layout of a twentieth-century village with an eye to the past meant understanding how its present or contemporary morphology was directly due to survivals of earlier organizational systems. They regarded the palimpsest as a very useful metaphor for the landscape. However, the term as it is usually employed today draws from a very different genealogy: the 1980s ideas of Binford, Foley and Bailey, who used the term to refer to the way successive events erase or disturb previous material configurations, which had both negative and positive consequences (e.g. see Bailey 2007). The negative ones involved the destruction of the integrity of short-term contexts, which reduced one’s ability to examine the past on a micro-scale. The positive ones involved the manifestation of longer-term processes and the potential of archaeology for seeing what was invisible to other disciplines with a shorter termporal focus. Thus looking at a deposit as a source of information on past behaviour at an ethnographic scale was severely constrained due to the greater fragmentation of the original context, but taking a longer-term view meant also having a more coherent data set to work with. The difference between the 1920s and 1980s conceptualizations of the palimpsest illustrates the way in which the idea of the relic has yielded to the idea of the fragment as a way to think about the archaeological record. From seeing a palimpsest as an idea to articulate the left-over, the residue to a notion which defines incompleteness in terms of scales of resolution demonstrates how effectively this shift has happened. In the next section, I want to re-visit the idea of the relic and see if there is any scope for reviving it.

Survival of what? One of the problems with the idea of the fragment is that it emphasizes the limits of archaeological interpretation – even in an age when we want to downplay such a notion, it haunts archaeology like a shadow. The fragment elicits an absence, either of the whole or just missing pieces, and makes archaeological interpretation a mediation between the fragment and what is absent. The relic, however, has rather more positive implications, because it emphasizes the continuity of the past into the present; the question it elicits is more like that raised when a gatecrasher appears at a party: what are you doing here? The archaeological is what does not belong, temporally speaking. Yet we do not usually feel like that, which is the really strange thing. Our sensibilities have been so dulled, so familarized to this gatecrasher (in fact, we actively seek him out) that this question never occurs to us. We know these things do not belong – we are not concerned with what they are doing here, only with where they came from. Of course we show some interest in how they got here – but only for methodological reasons (e.g. formation theory, source criticism). Theoretically, the question of their lingering, or persistence, is a given. But let me try to recover some of our original naivety here (if we ever had it), and pose the question theoretically: why are these things still here? 318 Gavin Lucas

Back on our Icelandic island, we excavated a few small test trenches into house- hold middens to recover some artefact assemblages, and of course we unearthed a mass of finds. These included the usual stuff one might expect from early twentieth- century sites: lots of ceramics, glass vessels and structural ironwork, among other things. We also found some scraps of textile and, in one trench, remains of several leather ankle boots. Every archaeologist knows that some materials survive better than others – the glass and ceramics could have been dropped yesterday; the iron nails, however, are mostly rusted and corroded, while organic remains are in an even worse condition. Now conventionally, we might say that this stuff is still here because it physically endures; because it was discarded and buried in an environment relatively favourable to preservation; because no major earth-moving activities have occurred on the site since deposition. Through consideration of taphonomy and formation processes, we routinely take account of factors that have facilitated the preservation or destruction of what survives – and of course, the configuration of its survival. Schiffer’s c and n transforms if you like. But is that all there is to it? More importantly, what if we do not treat these as methodological issues, factors to consider when trying to reconstruct the ‘systemic context’, but as fundamental to the way we interpret the archaeological record? What if we consider the question of survival to be central to any interpretation we make? If we do take this seriously, then we really need to ask more carefully what it is that we mean by survival. Straightaway we need to question any simplistic idea of physical preservation. For something to survive, it needs to retain a degree of identity as an object, which means its has to cohere, or hold together. This can involve what we commonly refer to as the physical properties of an object vis-à-vis its environment, e.g. the material structure and form of ceramic or bone. And of course one also has to consider the agencies responsible for bringing an object and its environment into contact – someone discarding a bone down a well, humans digging up and displacing buried remains to build a well. Thus far we remain within the familiar territory of formation theory. But the question of survival does not just pertain to artefacts like pots or other remains like bones; it also pertains to larger assemblages of things, and I use the concept of assemblage here in a rather different sense to the way an archaeologist might – as a depositional or typological collection of material (see Lucas 2012 for a more in-depth discussion). By an assemblage, I mean a recurrent and enduring gathering of things connected to practices or performances; my kitchen is an assemblage, as is my house. But so is a gathering of people for a dinner party or a funeral. Assemblages are, as the word implies, gatherings, and these can be very ephemeral (e.g. dinner party) or very long-lived (building). The survival of assemblages is both very different and, on another level, exactly the same as the survival of objects like a ceramic plate. It is different because different agencies are involved, but it is similar because the same issues are very much at stake: the coherence or holding together of the entity, which is a question of both its internal constitution and its external environment. Let me take an example from my island village. Consider an assemblage like one of the houses. What held it together, internally? Well, we know from documents Evidence of what? On the possibilities of archaeological interpretation 319 and photographs that it was made from timber and had a series of tenants over its lifetime; what held it together was a combination of factors, including the structural properties of its walls, roof and foundations and the maintenance activities carried out by its tenants. More generally, it survived as a viable assemblage so long as its critical parts remained enchained through recursive practices – cooking, eating, sleeping, heating, standing, sweeping, painting, etc. Looking at it now – a stump of foundations – it clearly has not survived as this assemblage. Parts of it have survived, either in the same location, like the foundations, or in new locations, like ceramics and stove parts, which ended up in a midden. But the assemblage itself has not survived. We can ask, why not? It may be due to a combination of factors, but in this case it most likely relates to a change in its external environment. Indeed, next we need to ask, what held this assemblage together externally? The answer is the viability of the village as a whole and its relation to the company which established it, as well as the wider network of fishing and trade within Iceland and the North Atlantic. When this company when bankrupt, when the provision of services to the island decreased, when the opportunities for survival of its human inhabitants diminished and they left the island, then critical factors which facilitated the cohesion of this assemblage no longer existed. Consequently, the internal cohesion of the assemblage was threatened. People left and took usable parts with them, including most of the timber superstructure. The assemblage was disas- sembled. Now all this is straightforward in many ways. Of course, the details of this disintegration process could be greatly fleshed out, but the issue here has been to show how, in thinking of the archaeological record in terms of survival at a more general, theoretical level, what was previously just a methodological issue now becomes quite obviously a susbstantive one regarding interpretation. This is because we have subtly shifted the nature of the archaeological object. It is not simply the stuff we dig up – potsherds, flints, bones, house foundations. The theoretical question of survival is not primarily about these things at all, but about assemblages of things, insofar as an assemblage is understood not in the conventional way but in the sense outlined above. That is, an assemblage is a group of things connected either as part of a common practice – an orchestra of players, instruments, chairs, music stands, etc. – or as part of a larger, functioning entity – a building comprised of bricks, pipes, cables, carpets, tables and so on. The assemblage is what matters here. Let me, in ending, expand on this last point and finally answer the question posed by my title.

Evidence of what? What is the archaeological record evidence of? This is a question which archaeolo- gists will probably not agree on, yet perhaps there are still identifiable com- monalities. Some will claim they give us evidence of ideologies, religous beliefs or gender systems, others evidence about economies, ecologies or technologies. Maybe most archaeologists will agree that all of these are accessible, to varying 320 Gavin Lucas degrees depending on the nature of the remains and methods of analysis used. Archaeologists might also vary on the level of detail we can hope to gain – small- scale events and practices or long-term processes and structures? Individuals or collective identities? Again, it is easy to dichotomize, but in practice I suspect most archaeologists would accede that all of these are possible to infer, depending again on the quality of data and rigour of method. What seems to be at stake is not so much the ontology of what is being interpreted, but the applicability of any given interpretation in relation to the nature of the evidence. And yet some quite profound ontological leaps are routinely assumed in archaeology: we go from pieces of flint to exchange networks, from graves to ideology, from table settings to capitalism. On the one hand, such connections can be lauded for showing the relationship between ‘small things forgotten’ and larger-scale structures and processes defining historical periods, societies and cultures. They show the power of archaeology to contribute to questions that count. On the other hand, such connections also seem to slip too easily from one ontological register into another – from the concrete objects we excavate and analyse to the conceptual abstractions that litter our discourse. For example, it would be very easy to talk about the abandonment of the village on Viđey in terms of wider political and economic forces – the shift towards a service-oriented economy and the decline of the fishing industry. But we lose so much in making such broad leaps. These political and economic forces themselves are made up of specific material configurations: ships, harbours, shops. To understand the material processes involved in making and unmaking the village, we need to connect them to other material processes elsewhere. These connections are the economic and political contexts. This is a point made before by others in relation to the social sciences and humanities in general, most notably by Bruno Latour, and connects to general critiques of the social (e.g. Latour 2005). I do not want to dwell on this critique here for lack of space, but another way to articulate this problem is in terms of the distinction between evidence and evidence of. Evidence in archaeology typically comprises the data and patterns in data derived from the physical remains of the past. But we would never claim that a pot excavated from a grave was evidence of the pot having existed in the past; we would surely cede the point, but it is so obvious it does not need stating. No, the pot is really evidence of something else – production technology, exchange networks, craft specialization, and so on. But what if we take the existence of the pot as evidence of itself – that is, of its persistence? What if survival is the point here? The pot, materially speaking, is the stabilization of other elements – water, clay and temper brought together under certain conditions. This stabilization, or persistence, is also what the pot, quite concretely, is evidence of. This is not a call for archaeology as artefact physics or confining interpretation to the lowest rung of Hawkes’s ladder, because we can approach the village on Viđey in exactly the same way: as the stabilization of elements – stone, wood, nails, stoves, people – held together under certain conditions. This is how I have tried to articulate it in the previous section. But there is another dimension of survival to tack on here. Evidence of what? On the possibilities of archaeological interpretation 321

In the previous section, I stated that the village has not survived, as an assemblage. This is true insofar as the town across the bay – Reykjavík, which existed contemporaneously with the village – is still surviving, albeit as a mega- assemblage. But while one aspect is about the survival of an assemblage as a functioning entity or gathering, the other concerns its survival as a recognizable assemblage, even if now defunct. From a potsherd you can still discern the whole pot; from house foundations and middens you can still discern the village. But an underfired pot that melts into the soil may leave no more trace of what it once was than a gathering of people on a sunny day for a picnic, even if one may find a discarded bottle here or a bone there. Finding remnants of an assemblage is not the same as finding the assemblage itself if the relation between those parts is lost. Moreover, the integrity of an assemblage has no necessary relation to its durability; traces of a short-term event like a burial could endure and be visible archae- ologically much more than a building that stood for 30 years. What my earlier discussion of survival was intended to do was to foreground a study of the archaeological record in these terms. The real significance of thinking about the archaeological record in terms of survival is that it asks us: survival of what? I would argue that this question is identical to the other question: evidence of what? If we make the link between these two questions, then we start to see things somewhat differently. If the archaeological record is really about survival, specifically the survival of assemblages as material entities, then it is evidence of a reality that is of the same nature as the remains themselves. We have what DeLanda (2002) calls a flat ontology, not a dual one. How has all this changed the way we might think about the archaeology of the village on Viđey? What does viewing in terms of relics instead of fragments mean for the archaeological record there? Let me return to the story I opened with and the questions that I raised there. If the question that prompted this chapter has been, ‘why have some things survived?’, it nonetheless still connects to my original research question: where did everything go? Foregrounding the notion of survival is common to both. In my work on the village, the goal was not to reconstruct some former totality from fragments – the village that once was – but rather to explore why some things remained and some things did not. It was about the flows of objects that combined into assemblages and then disassembled. It was about exploring how assemblages come to exist and then cease to be. The point of seeing the archaeological record in terms of the relic, in terms of survival, is that our focus is directed towards identifying the constitution of those objects or entities that the archaeological record is evidence of. It is fundamentally an ontological endeavour. It is therefore also one which opens possibilities for exploring new and different kinds of objects in the archaeological record and the past. Elsewhere, I have explored the possibilities of a multi-subject archaeology where the focus of the narrative is, among others, a settlement that is viewed as a form of extended organism (e.g. Lucas 2013). Such an approach connects up with the broader field of posthumanities (e.g. see Braidotti 2013). It remains an open question to what extent such a perspective might alter the practices of fieldwork; my fieldwork on 322 Gavin Lucas this project was in many ways very conventional, but in foregrounding the notion of relic it may well be that a very different set of methodological concerns could come into play.

Epilogue At the start of this chapter, I discussed the limits of archaeological interpretation, yet in my title, I used the word possibilities; for me, the two are inextricably linked, and in fact if we shift the question of limits to that of possibilities, then we might find ourselves thinking about the archaeological record and archaeological interpretation in different ways. Arguably, at its most basic level, archaeology is only possible because things from the past have survived into the present. If no material traces remained, archaeology would not exist. Survival then becomes the necessary precondition for archaeology; by no means is it sufficient, but it is surely a necessary condition. It is why antiquarians and early archaeologists used terms that emphasized this particular quality – vestiges, remains, relics. If we do accept that point (which no doubt could be contested), then I think it should also be reasonable to define the limits of interpretation through the same concept. Now it might appear that we already do – after all, is not the idea of incompleteness, the fragment, the key trope for articulating all the critical problems of archaeological interpretation? It is, but as I have suggested in this chapter, incompleteness itself has no necessary relation to the idea of survival and in fact it is a concept we have adopted over survival. I would not argue that the idea of incompleteness has no utility, but it does raise the question: incompleteness of what? Now we could debate that endlessly, but unless one grounds it in the temporal nature of archaeological remains themselves – unless one subordinates that question to the question, ‘survival of what?’, then I think we remain in danger of losing sight of what we can and cannot accomplish with material culture as evidence in archaeological contexts.

References Adorno, T. and Horkheimer, M. (1997) Dialectic of Enlightenment, London: Verso. Bailey, G. N. (2007) ‘Time perspectives, palimpsests and the archaeology of time’, Journal of Anthropological Archaeology, 26: 198–223. Binford, L. (1962) ‘Archaeology as Anthropology’, American Antiquity, 28(2): 217–225. Binford, L. (1968) ‘Archaeological Perspectives’, in L. Binford (ed.) New Perspectives in Archaeology, Chicago, IL: Aldine Publishing Co., 5–32. Binford, L. (1983) In Pursuit of the Past, London: Thames & Hudson. Braidotti, R. (2013) The Posthuman, Cambridge: Polity Press. Burström, M. (2013) ‘Fragments as something more: archaeological experience and reflection’, in A. González-Ruibal (ed.) Reclaiming Archaeology: Beyond the Tropes of Modernity, London: Routledge, 311–322. Carver, M. (1990) ‘Digging for data: archaeological approaches to data definition, acquistion and analysis’, in R. Francovich and D. Manacorda (eds) Lo scavo archeologico: dalla diagnosi all’edizione, Firenze: All’Insegna del Giglio SAS, 45–120. Evidence of what? On the possibilities of archaeological interpretation 323

DeLanda, M. (2002) Virtual Science, Intensive Philosophy, New York: Continuum. Ginsberg, R. (2004) The Aesthetics of Ruins, Amsterdam: Rodopi. Haraway, D. (1988) ‘Situated knowledges: the science question in feminism and the privilege of partial perspective’, Feminist Studies, 14(3): 575–599. Harding, S. (1986) The Science Question in Feminism, London: Cornell University Press. Harrison, R. (2011) ‘Surface assemblages: towards an archaeology in and of the present’, Archaeological Dialogues, 18(2): 141–161. Hartsock, N. (1983) ‘The feminist standpoint: developing the ground for a specifically feminist historical materialism’, in S. Harding and M. Hintikka (eds) Discovering Reality, Dordrecht, Holland: D. Reidel Publishing Company, 283–310. Hawkes, C. (1954) ‘Archaeological theory and method: some suggestions from the Old World’, American Anthropologist, 56(2), 155–168. Hodder, I. (1986) Reading the Past: Current Approaches to Interpretation in Archaeology, Cambridge: Cambridge University Press. Horkheimer, M. (2004) Eclipse of Reason, London: Continuum. Lakoff, G. and Johnson, M. (1980) Metaphors We Live By, Chicago, IL: Chicago University Press. Latour, B. (2005) Reassembling the Social, Oxford: Oxford University Press. Lucas, G. (2012) Understanding the Archaeological Record, Cambridge: Cambridge University Press. Lucas, G. (2013) ‘Archaeology and the science of new objects’, in B. Alberti, A. Jones and J. Pollard (eds) Archaeology after Interpretation: Returning Materials to Archaeological Theory, Walnut Creek, CA: Left Coast Press. Lucas, G. (2014) ‘Conduits of dispersal. Dematerializing an early twentieth century village in Iceland’, in B. Olsen and þ. Pétursdóttir (eds) Ruin Memories: Materialities, Aesthetics and the Archaeology of the Recent Past, London: Routledge. Lucas, G. and Hreiđarsdóttir, E. (2012) ‘The archaeology of capitalism in Iceland: the view from Viđey’, International Journal of Historical Archaeology, 16(3): 604–621. Lyman, R. L. and O’Brien, M. J. (2004) ‘Nomothetic science and idiographic history in twentieth century Americanist anthropology’, Journal of the History of the Behavioural Sciences, 40(1): 77–96. Marchand, S. L. (1996) Down from Olympus: Archaeology and Philhellenism in Germany, 1750–1970, Princeton, NJ: Princeton University Press. Olsen, B. and Pétursdóttir, þ. (eds) (2014) Ruin Memories: Materialities, Aesthetics and the Archaeology of the Recent Past, London: Routledge. Patrik, L. (1986) ‘The aesthetic experience of ruins’, Husserl Studies, 3: 31–55. Roskams, S. (2001) Excavation, Cambridge: Cambridge University Press. Schiffer, M. B. (1987) Formation Processes of the Archaeological Record, Salt Lake City, UT: University of Utah Press. Schmidt, D. (2002) ‘Refuse archaeology: Virchow – Schliemann – Freud’, Perspectives on Science, 9(2): 210–232. Shanks, M. (2012) The Archaeological Imagination, Walnut Creek, CA: Left Coast Press. Smith, M. (1955) ‘The limitations of inference in archaeology’, The Archaeological Newsletter, 6(1): 3–7. Spivak, G. (1999) A Critique of Postcolonial Reason. Toward a History of the Vanishing Present, Cambridge, MA: Harvard University Press. Vismann, C. (2001) ‘The love of ruins’, Perspectives on Science, 9(2): 196–209. Wylie, A. (1992) ‘The interplay of evidential constraints and political interests: recent archaeological research on gender’, American Antiquity, 57: 15–34. Young, R. (1990 ) White Mythologies: Writing History and the West, London: Routledge. 18 MEETING PASTS HALFWAY

A consideration of the ontology of material evidence in archaeology

Andrew Meirion Jones

What is material evidence?

Archaeologists have always recognised the malleable nature of their relationship to past material culture. In his magisterial study of the history of archaeology, The Discovery of the Past, Alain Schnapp (1993: 30) draws our attention to the imaginative character of archaeological interpretation. He argues that from its very inception those who discovered and studied past material culture recognised that their interpretations were an exercise in imaginative reconstruction. This recog- nition has continued throughout the history of the discipline, whether acts of reconstruction were framed as objective and scientific enterprises (e.g. Binford 1983: 45–57; Clarke 1973; Randall-McIver 1933; Watson et al. 1971) or as hermeneutic processes (e.g. Hodder 1992; Shanks and Tilley 1987; Shanks and Hodder 1995; Tilley 1993). Whatever the precise character of the interpretative process, throughout the history of the discipline antiquarians and archaeologists have been remarkably consistent in their characterisation of material evidence: material evidence is conceived as the material trace of past human activity. It is this characterisation of material evidence that this chapter seeks to question. In doing so, I will offer an alternative characterisation of material evidence, by considering the relational characteristics of material evidence, while also briefly examining the ontological status of this evidence. The relationship between archaeologists and the object of their study is summed up neatly in Chris Tilley’s statement that ‘without the interpretative work of the archaeologist the past would remain dead and gone’ (Tilley 1993: 20). It would seem that the malleable interpretative work of archaeology rests on the assumed solid foundations of materials. A number of issues are caught up in the tacit characterisation of materials as a form of evidence. The notion that the past remains ‘dead and gone’ without the input of archaeologists relies upon the allied assump- Meeting pasts halfway: a consideration of the ontology of material evidence 325 tion that materials are mute and inert without a human presence. Inert materials once received the imprint of human cultural activity; it therefore remains for archaeologists to recognise the significance of these materials and culturally animate them anew. In this characterisation, materials that have been culturally worked upon in the past can act as a form of evidence for that past activity. This characterisation of materials can occur only when we perceive materials as distinct from culture or society. It can also occur only when we place the interpreting subject at the centre of our analyses. Both viewpoints are the outcome of a dualism that has its roots in the philosophy of Descartes and Kant and a suite of Enlightenment thinkers. The work of Descartes and Kant both positioned the thinking subject as the foundation or centrepiece of their philosophies as an alternative to notions of God or Truth as foundational concepts; Descartes’s formulation cogito ergo sum, ‘I think therefore I am’, is therefore a foundational statement par excellence. Likewise, numerous Enlightenment thinkers, including Hume, Voltaire and Montesquieu, wished to establish a ‘science of man’ (Porter 2001: 11–21) that placed human individuality at centre stage. In Enlightenment thought humans were imagined to be the fulcrums of a material universe upon which they acted, interpreted, manipulated and created. Archaeology – as a product of the Enlightenment and Modernity (Thomas 2004) – has inherited this perspec- tive. As Thomas (2004) points out, an important consequence of Enlightenment perspectives is that the distinctions imagined to pertain between materials and culture, or nature and society, provide the very conditions of possibility for archaeological study; due to this archaeologists are able to conceptualise materials as categories of evidence for past human cultural or social activity. However, these conditions of possibility for archaeology come at the expense of treating materials as inert and mute:

As part of the structure of modern thought, archaeology seeks clarity, objectivity, and a reduction to law-like or mathematical terms. It demands precision, unambiguous resolution, universality and the transcendence of local conditions. All of this is achieved by declaring the world to be object- like and free of meaning . . . (Thomas 2004: 247)

As Thomas so eloquently demonstrates, materials are rendered object-like, without meaning, and inert due to archaeologists’ characterisation of them and interaction with them. In many ways this resonates with the analysis of the relationship between people and materials in the work of the philosopher of science Karen Barad. In her book Meeting the Universe Halfway, Barad (2007) analyses particle physics, and especially Niels Bohr’s discovery and description of the atom. Barad’s particular interest is the way in which knowledge and matter are mutually performed. She develops Bohr’s argument that ‘theoretical concepts are defined by the circumstances required for their measurement’ (see Barad 1999: 3) to argue that there are no unambiguous methods of differentiating between an ‘object’ and the 326 Andrew Meirion Jones

‘agencies of observation’ used to examine and define that object. Observations of the world do not simply represent the world; they help to bring the world into being. Julian Thomas appears to argue as much in the quote above. Thomas especially emphasises the way in which the intellectual legacy of archaeology produces a particular conceptualisation of materials. On a practical level, Gavin Lucas (2012: 215–216) also points out that archaeological interactions with the site or artefact help to produce the object of study. Lucas uses the example of the German archaeologist Gerhard Bersu, interned in wartime Britain. Bersu’s larger open-area excavation of British Iron Age sites helped post-war British archaeolo- gists recognise the validity of post-built structures as opposed to pit dwellings; the intervention of new methods of observation produced a new kind of archaeological record. In Barad’s analysis materials are constituted by human observation and inter- action, a process she describes as ‘agential realism’ (Barad 1999), which emphasises the constitutive power of interaction and the mutability of materials. In a clear definitional statement, Barad (2007: 170) points out that her conceptualisation of ‘agential realism’ entails a reconceptualisation of materiality, in which matter is agentive and intra-active. It is this intra-activity that is particularly important for my argument. I follow Barad’s argument that

Matter is a dynamic intra-active becoming that never sits still – an ongoing reconfiguring that exceeds any linear conception of dynamics in which effect follows cause end-on-end, and in which the global is a straightforward emanation outward of the local. Matter’s dynamism is generative not merely in the sense of bringing new things into the world but in the sense of bringing forth new worlds, of engaging in an ongoing reconfiguring of the world. Bodies do not simply take their places in the world. They are not simply situated in, located in, particular environments. Rather ‘environments’ and ‘bodies’ are intra-actively co-constituted. Bodies (‘human’, ‘environmental’ or otherwise) are integral ‘parts’ of, or dynamic reconfigurings of, what is. (Barad 2007: 170)

The paragraph above highlights a series of key points. Intra-activity means that bodies are relationally configured; matter is dynamic and undergoing constant change; intra-activity is also generative, and through intra-activity new worlds, new possibilities, are realised. These ideas seem to parallel some of the points Gavin Lucas discussed regarding archaeological fieldwork (see Lucas 2012). The notion of intra-activity seems particularly helpful in an archaeological context in which the character of archaeo- logical data is shaped by fieldwork and intervention. To consider the validity of this approach I will analyse the behaviour of some materials that I have encountered in a recent archaeological project. One material – rock – was encountered as part of a field project relating to rock art; the other material – quartz – was encountered in post-excavation laboratory analysis. Meeting pasts halfway: a consideration of the ontology of material evidence 327

Living rocks, part 1: rock art and the prehistory of Kilmartin, Argyll, Scotland The Kilmartin region, in Argyll on the west coast of Scotland, is an important prehistoric landscape (Butter 1994). The landscape includes numerous early Neolithic chambered tombs, a late Neolithic henge, several stone rows of probable middle-late Bronze Age date and one of the most extensive early Bronze Age cairn cemeteries in Britain. In Britain, rock art is a largely northern and western phenomenon (Bradley 1997), and, with some 133 individual rock art sites, the Kilmartin landscape is the most significant rock art landscape in Britain. The Kilmartin rock art project was directed by the author, and ran from 2002 to 2009 (Jones et al. 2011). Rock art is notoriously difficult to date (Whitley 2005: 53–78), and, through excavation around a number of rock art panels, the field project aimed to provide some of the first dates for rock art in the British Isles, and also understand the position of the rock art in the development of the local prehistoric landscape. One of the reasons rock art is difficult to date is that it does not behave in the same way as other forms of archaeological material: it is not discretely stratified, but outcrops and impinges on the landscape. In this situation both relative and absolute dating methods can be problematic. The first season of work began by analysing rock art in the wider landscape. Taking in most of the major sites, we examined the relationship between the carvings and the rock surfaces on which they were executed. One of the striking observations to emerge from this work was the close relationship between the texture of the rock surface, its geological cleavage planes, joints and fissures, and the positioning of the rock art engraved on its surface (see Figure 18.1). Fieldwork in subsequent seasons demonstrated that cracked and fissured surfaces were

FIGURE 18.1 Ormaig at night, illustrating the relationship between rock art and cracks and fissures. (Photograph: Andrew Cochrane and Aaron Watson; copyright: author.) 328 Andrew Meirion Jones deliberately chosen for carving. I have discussed this on a number of occasions elsewhere, and have described the rocks as occupying an ‘ancestral’ relationship to their carvers (Jones 2005, 2006; Jones and Tipping 2011). I have also noted the animated and living character of the rocks (Jones 2011, 2012a). Here, I am not concerned with pinning down the precise character of the relationship between rocks and rock art; instead I simply wish to note that the rocks impinged upon and intra-acted with their past human carvers. The rocks also impinged upon, and intra-acted with, other aspects of the project. In the third season we began excavation at two rock art sites at Torbhlaren in the Kilmichael Glen region (see Jones et al. 2011). At the smaller of these two sites – Tiger rock – excavation close to the eastern face of the rock revealed a clay-and- stone built platform covered in quartz debris (see Figure 18.2). Excavation beneath this platform revealed a small post-built structure dating to the Late Neolithic. It was possible then to determine a sequence of activity beginning with the construction of the post-built structure. This site was burned down prior to the construction of the clay-and-stone built platform, which extended the length of the eastern face of the rock. This platform was covered in quartz debris, which upon post-excavation analysis proved to consist of hammerstones used for the production of rock art. In addition to this, a large fissure on the rock surface also contained stratified quartz debris that likewise proved to consist of the remains of hammerstones. This also produced late Neolithic radiocarbon dates. The larger of the two rock art sites – Lion rock – had no built structures associated with it. Instead, a large rocky natural platform ran along its western flank

FIGURE 18.2 Excavating the clay platform on the eastern face of Tiger Rock, Torbhlaren. (Photograph: Aaron Watson; copyright: author.) Meeting pasts halfway: a consideration of the ontology of material evidence 329

(see Figure 18.3). A large fissure running along the inside length of this rock platform contained stratified deposits of quartz debris, also proving to be the remains of hammerstones used in rock art production. These produced Late Bronze Age radiocarbon dates. In addition, several fissures on the upper surface were excavated and produced a variety of materials, including quartz, flint pebbles and small flakes of pitchstone – an obsidian-like volcanic glass mainly exploited during the Scottish Neolithic – as well as beads of glass and amber (Ballin 2009). These latter fissures produced a number of dates ranging from the Iron Age to high Medieval periods. The rocks on which the rock art motifs were carved at Torbhlaren impinged upon activities in a number of ways. Most obviously, a variety of archaeological artefacts were deposited in geological joints and cleavage planes, the cracks and fissures on the rock surface. In addition, archaeological features were closely associated with the rocks. At Tiger rock, the post-built structure and platform were located along the shelved and domed features of the eastern edge of the rock. At Lion rock, a natural shelf occurred along the western edge of the rock, whose fissure received deposits of hammerstone debris. In each case these features were in close proximity to the concentration of carved motifs on the rock surface. Finally, the major archaeological artefacts consisted of the debris of quartz hammerstones that had shattered on contact with the rock during episodes of carving in both the Late Neolithic and Late Bronze Age. The rocks continued to receive deposits of artefacts from the Neolithic to the high Medieval period, though these later deposits were not obviously related to rock-carving activities.

FIGURE 18.3 A linear fissure on the upper surface of Lion Rock, Torbhlaren. (Photograph: Aaron Watson; copyright: author.) 330 Andrew Meirion Jones

Our fieldwork in the Kilmartin region, and in particular at the Torbhlaren site, involved a series of encounters with the properties of rock. Rather than archae- ologists intervening on past material artefacts and thereby shaping the archaeological record, our fieldwork felt much more like an intra-active dialogue with another entity: rock. The properties of the rock surface appeared to affect which rocks, and which surfaces, were carved. The shape and form of the rocks affected what kinds of activities took place in relation to them. Did activities take place on the rock? Were they situated in crevices on the rock? How were artificial platforms built in relation to the rock surface? In each of these cases the character of the rock had to be reckoned with. Our fieldwork and excavations, while being interventions that created a new kind of archaeological record, also took place alongside, and with the presence and personality, of the rocks. In a more classically scientific sense, the geological formation and lithological characteristics of the rocks had an important influence on how we conducted fieldwork and excavated. Our interventions did not solely shape the archaeological record; the record was shaped both by us and by the character and form of the rocks. The rocks influenced us, just as they had influenced people in the past. Rock art research involves a particularly tactile and immediate form of archaeological research, as researchers crawl over rocks, touch them to sense the presence of difficult-to-see motifs, lie on them to record them; rock art researchers are especially sensitive to the changing form and character of rocks as they examine them. Intra-activity appears to lie at the heart of rock art research. If the members of the project had not intra-acted with the rocks of the Kilmartin region we would have applied conventional archaeological methods and techniques to them, and would have overlooked many of their significant features, including the deposits on their surfaces, and the deposits and built structures around their perimeters. We could not have arrived at the interpretations that we did, had we not intra- acted with the rocks of Kilmartin and in particular those at Torbhlaren. By intra- acting with them we learnt how the geological character of the rocks influenced the choice of which rocks were to be carved. Intra-acting with the rocks allowed us to recognise that we were dealing with upstanding three-dimensional solids as opposed to stratified two dimensional layers; close intra-action allowed us to understand where to excavate and what to record, which rock surfaces were significant, which cracks and fissures were important and which to ignore. This enabled me in my field report to consider how the forms of the rocks influenced the activities conducted around them: how some rocks provided a natural platform for carving, while in other cases artificial platforms augmented the forms of the rocks. It also considered the deposition of lithic raw materials and other artefacts around the rocks and asked why these materials had been discarded in this location; this triggered further questions concerning the significance and character of different kinds of rock during a certain period of prehistory. While the date of the sites were a focus for analysis, the resulting spread of dates from the Neolithic to the high Medieval period were also seen to be of value, as I recognised that the rock and its carvings continued to exert an influence on the inhabitants of the Meeting pasts halfway: a consideration of the ontology of material evidence 331 landscape from prehistory to the present day. Without a process of intra-action we would have approached the rocks using conventional methods and missed much that was significant. Instead we approached the rocks as entities whose changing character influenced a series of discrete and intimate deposits both on and around the rock surface; through an intra-active engagement we understood the multiple and complex facets of these remarkable archaeological sites.

Living rocks, part 2: post-excavation analysis of quartz artefacts from the Kilmartin project The bulk of artefacts recovered at the Kilmartin rock art sites were of quartz and struck flint and pitchstone. In addition, two beads (one of amber, the other of glass) were recovered during post-excavation flotation of samples of charcoal for radio- carbon dating. Here I want to focus on the analysis of the quartz. In total, 10,155 pieces of worked and unworked stone, weighing some 186.139kg, were recovered from the two rock art sites at Torbhlaren. The assemblage was investigated by lithics specialist Hugo Lamdin-Whymark. The following account derives from discussion with Hugo during the process of post- excavation analysis. As discussed above, there have been few excavations around rock art sites in Britain. Because of this, the character of artefacts associated with rock art sites was completely unknown. This accounts for the large quantities of stone recovered from the sites, as we erred on the side of caution during excavation, and retained everything that could conceivably be interpreted as a stone tool. The problem was exacerbated by the fact that the majority of finds were of quartz, and there has been remarkably little work undertaken on the character of quartz technology in British prehistory. Quartz cannot be worked in the same way as flint or obsidian. The post- excavation analysis involved becoming familiar with quartz, its properties and characteristics. In order for Hugo to sort between the worked and unworked stone from the two excavated sites he first needed to understand what he was analysing. He initially began by analysing the quartz assemblage as if it were flint, assuming that the majority of quartz finds from the site derived from percussion flaking for tool production. Very quickly, he realised that this analytical strategy was problematic: there seemed to be no evidence of tools being removed from the site for use (as would normally occur with the debitage from lithic production), and the typical patterns of working did not seem to be evident amongst the assemblage. In order to understand the quartz assemblage, a different approach was required. Statistical analysis of the size of pebbles of quartz (including quartz, fine-grained quartzite and coarse grained ‘Crinan Grit’ quartzite) indicated that 99.9 per cent of the assemblage by count consisted of pieces of quartz less than 75mm in size, with another group less than 10mm in size (Lamdin-Whymark 2011). These are hand- sized pieces, and Hugo wondered if these could have been utilised as hammer- stones. He embarked on a series of replication experiments, using local slabs of epidiorite (the native rock of the Kilmartin region) to reproduce rock art motifs 332 Andrew Meirion Jones using quartz chunks found during a geological survey of the region. In this way he made a series of remarkable discoveries. Analysis of wear on stones used in these replication experiments indicated that much of the wear observed on artefacts in the archaeological assemblage certainly derived from hammering: these were hammerstones utilised for rock art production (see Figure 18.4). However, importantly, he also observed that many raw materials shattered in a peculiar pattern when used for rock art production, shattering in a series of segments rather like those of an orange. These were also found in the archaeological assemblage, and accounted for the peculiar groups of knapped lithics erroneously assumed to be the result of tool production strategies; the reason there was no removal of lithics from the site is because these groups of material were the result of rock art making, not lithic production. He had therefore identified these groups of material as the residue of rock art production activities. It was through a process of intra-action with the unknown and unexpected character of quartz that interpretations of the Kilmartin assemblage were advanced. Intra-action took place between Hugo, quartz chunks and slabs of epidiorite rock. The quartz revealed its properties through a complex process of intra-action, then. ‘Off the shelf ’ theoretical and methodological assumptions derived from flint analysis were not equal to the task of interpreting the quartz from Kilmartin.

FIGURE 18.4 A selection of quartz hammerstones from Tiger Rock, Torbhlaren. (Photograph: Aaron Watson; copyright: author.) Meeting pasts halfway: a consideration of the ontology of material evidence 333

Instead, Hugo had to meet the quartz on its own terms if he was to understand it. It was a process of becoming familiar with the properties of quartz through practical experimentation that allowed him to comprehend the assemblage. From this he was able to then analyse the lithic assemblage as a whole, sifting between worked and unworked lithics, and advancing a coherent interpretation about the activities in which lithics were involved at the two sites in Kilmartin.

Meeting the past halfway The two case studies (relating to excavation and post-excavation) discussed above seem to support Karen Barad’s subtle argument that things ‘exist only as a result of, and as part of, the world’s ongoing intra-activity, its dynamic and contingent differentiation into specific relationalities’ (Barad 2007: 353). Above, I argued that intra-action is a relational process. Barad argues that engaging with, and meeting, the matter of the world is critical to understanding how the world comes to be composed. Engagement is a two-way process. Scientists meet the fabric of the universe, just as electrons, quarks and positrons meet scientists. Crucially, this is not a meeting of active human subjects manipulating an inert material world; it is a meeting of entities with equal powers to act, and equal, but differing, properties and qualities. I believe this argument is important, and the remainder of this chapter will develop this point with regard to archaeology. The two case studies above highlighted an important point: it is unhelpful to conceptualise materials as inert and mute. The rocks we surveyed and excavated around in Argyll impinged upon us, while the quartz imposed itself upon laboratory analyses. This accords well with the findings of scholars in disciplines as diverse as anthropology, geography, cultural theory and science studies. These scholars point out that in fact when we look closely at matter, whether at the atomic level (Coole and Frost 2010: 11–15) or at the macro scale (Ingold 2007; Henare et al. 2007), materials appear to be vibrant and changeable. It is possible to imagine matter as a lively materiality that is self-transformative (Barrett 2010; Whatmore 2008) and already infused with the capacity for agency and significance that are typically accorded humans in idealist and subjectivist accounts. Rather than treating materials as inert ‘stuff ’ that is manipulated by humans, we are instead able to consider how materials change and evolve, and think about the fecundity and dynamism of matter. We need to distinguish these accounts of agency from the ideas of ‘material agency’ familiar from archaeological and anthropological theory. Perhaps the best known of these is Alfred Gell’s notion of artworks as agents (Gell 1998). What is clear in Gell’s analysis is that materials, objects or artworks achieve a derived form of agency; this is an agency supplied to them by humans by virtue of standing in an indexical relationship to humans (Gosden 2001: 164); artworks exist in a nexus of relations, and due to this are able to stand in for, or act as indexes of, human relations and intentions. Analytically artworks might ‘look like’ humans, but their agency is deficient of the intentionality of humans. While this approach to material 334 Andrew Meirion Jones agency wishes to consider the agentive potentiality of objects, things or materials, in fact it preserves familiar ontological dualisms (Holbraad 2009). To reduce the agency of matter to that of human agency is unhelpful. In the opening section of this paper I highlighted the inheritance of the dualist thinking of the Enlightenment in twentieth-century archaeological thought. If we are to take seriously Karen Barad, the scholars noted above and the results of my archaeological case studies, we are adopting a quite different philosophical standpoint: a standpoint that assumes a monist perspective, rather than a dualist perspective. Monism is associated with the philosophy of Leibniz and Spinoza. Spinoza’s philosophy is an attempt to develop the underlying rationalist motivations of Descartes’s work, and to follow through these motivations more consistently than Descartes (Della Rocca 2008: 87). Rather than there being several things in the world – ‘mind’ and ‘matter’, as Descartes supposes – Spinoza argues for the view that there is fundamentally only one thing in the world; concomitantly, you and the table, for example, are merely modes of states or properties of this single substance (Della Rocca 2008: 87). A monistic universe is therefore composed of intersecting affective relations. Humans take their place in this universe alongside other things, but they are not situated centre stage (Braidotti 2013). It is particularly important that we do not regard these intersecting relations as characteristic of a peculiarly human form of agency, as we saw with Gell’s formulation of artworks as a form of ‘secondary agency’. Instead, we need to recognise the multiple forms that agency may take, and that these forms of agency are intersecting and also intra- actively generative (see Hodder 2012), whether those agencies take the form of companion species like dogs and cats, domestic animals like cattle and sheep or the walking sticks and wheelchairs that Donna Haraway’s father, Frank Haraway, used to propel him to a career in sports journalism (Haraway 2008: 165–173). If we take these discussions on board, how do we conceptualise materials as classes of archaeological evidence? To do this we need to distinguish between two important archaeological concepts: context and assemblage. In the dualistic view of materials, to treat materials as evidence is to imagine that materials are not the authors of their own activities; rather, they are residues of activity directed by other social or cultural forces. Concomitantly, the meaning or significance of material evidence bears only an indirect relationship to materials, as they are not the authors of the activities whose evidence we are interested in interpreting. In this formulation the past is ‘dead and gone’ and all that remains is material evidence that requires evaluation by archaeologists. One of the ways we evaluate the significance of material evidence is by placing materials in context. Contexts allow archaeologists to differentiate the meaning and significance of materials, according to their contextual relationships and associations (for fuller discussion of the archaeological concept of context, see Hodder 1992). Alternatively, if we instead think about materials in a monistic sense then we accept that materials are authors of their own activities; we do not have to assume that materials can only be understood because of their mediation by human cultural or social activities. We can still study human social and cultural activities, but Meeting pasts halfway: a consideration of the ontology of material evidence 335 we are instead examining how these activities intersect with the activities and potentialities of materials. A more useful way of thinking about this relationship archaeologically is to consider materials as components of archaeological assemblages (for discussion of archaeological assemblages, see Lucas 2012). To draw out the differences between context and assemblage, an analysis of context implies analysis of the displacements of meaning that occur when things are placed in context, while an analysis of assemblage implies comprehending how things are composed or assembled. While context is an analysis of arrangement and organisation, the idea of assemblage relates to an analysis of arranging, organising and fitting together of heterogeneous components (for further discussion, see Jones et al. 2013). Contextual analysis places materials in a passive role, as materials are mainly evaluated according to their contextual associations and arrangements. Analysis of material assemblages considers how materials actively compose, make up and hold together their associations, and how these associations affect compositions (see Fowler 2013).

From material evidence to material intra-action If we consider that materials are not inert, but are vibrant and full of potential, there are important consequences for how we understand human engagement with materials in the past. Chantal Conneller (2011) has provided the most sophisticated account of this. Conneller examines a range of Palaeolithic technologies, from the production of Aurignacian basket-shaped beads of ivory and stone to the fashioning of decorated implements (so-called contours découpés) from the hyoid bones of horses in the Magdalenian. In each case she provides a sophisticated discussion of how materials are known and how they exhibit different qualities in different circumstances. Rather than materials containing a finite list of inherent properties that are actively elicited by early humans, instead properties and qualities are the product of interaction among material, technology and human. This interaction between people and materials is also highlighted in the work of Matt Edgeworth (2012). Edgeworth is concerned to point out that excavation concerns a direct encounter with materials, and that following the composition of stratigraphy, the layers, cuts and edges of archaeological features during excavation generate meaning. For Edgeworth (2012: 77), as archaeological material emerges during excavation, it has the power to re-shape both action and thought. The work of Conneller on archaeological materials and that of Edgeworth on excavations are therefore united by their recognition of the significant role materials play in human interactions. Both employ the term ‘interaction’ in their accounts, but these accounts equally make sense if we consider them as ‘intra-actions’ of the type discussed by Barad. Again, if we accept that materials are vibrant and dynamic, this alters how we conceptualise the organisation of the discipline intellectually. Traditionally, archaeological scientists merely characterise materials and provide an account of materials to archaeological theorists, who interpret the significance of these 336 Andrew Meirion Jones materials (see Jones 2002 for an analysis of this relationship). Instead, if we argue that these materials are vibrant and dynamic, then the role of archaeological scientists changes completely as they offer an account of the properties, qualities and capacities of materials. Importantly we are no longer concerned with how these materials can be interpreted; instead we are interested in how these materials intra- acted with past peoples. Following Conneller (2011), we are concerned instead with how the properties and qualities of materials are the product of intra-actions among materials, technologies and humans. This therefore places archaeological scientists at centre stage in the interpretation process. Archaeological scientists are no longer simply providing ‘data’ on the characteristics of materials. Instead they are explicitly concerned with how the properties and qualities of materials relate to past human lives.

Conclusion In a monistic understanding, material evidence consists of the residue of past human and material interactions. Materials are not inert, nor are they ‘dead and gone’; they continue to exert an influence in the present. The task of the archaeologist involves evaluating how materials intra-acted with human lives in the past and how materials enabled the propagation of these past human lives. As a consequence, the qualities and properties of materials are evaluated as a means of establishing how they were used and how they intra-acted with people in the past. There are important implications for how we consider social or cultural analyses in archaeology. If we are to examine human and material interactions in the past it is no longer appropriate simply to explain the character of material on the basis of its assumed cultural or social context. John Barrett argued some years ago that both processual and post-processual accounts operated with the assumption that material evidence represents past events or processes (Barrett 1994: 155–172). Processual archaeologists sought to understand certain ideal types of social formations (bands, tribes, chiefdoms, states) that are argued to have left their material signatures in the archaeological record (for an incisive analysis of these assumptions, see Pauketat 2007). Meanwhile, post-processual archaeologists attempted to make artefacts meaningful or comprehensible by situating them in contexts that are products of a universal structure (Barrett 1994: 167). Barrett goes on to note that in this interpretative project, ‘artefacts mean nothing. It is only when they are interpreted through practice that they become invested with meanings and may then act as props for the strategies of social life’ (Barrett 1994: 168–169). Although Barrett discusses post-processual archaeology here, he could equally be talking about how artefacts are treated in processual approaches; in both cases, artefacts are simply inert props, which illustrate broader social or cultural processes. If we instead consider materials as active components of past social or cultural worlds, then the task of social or cultural analysis shifts to understanding how materials interacted with, or intersected with, past peoples. Rather than attempting to explain materials on the basis of their cultural or social context (for Meeting pasts halfway: a consideration of the ontology of material evidence 337 similar arguments in anthropology, see also Henare et al. 2007; Ingold 2013) we are instead considering how past social or cultural lives were conducted with materials. We observe a shift, then, from an epistemological concern with how we can interpret from past material evidence to an ontological appreciation of what past materials were, how they behaved in intra-actions with past peoples and how they continue to intra-act with the archaeologists who study them.

References Ballin, T. (2009) Archaeological Pitchstone in Northern Britain: Characterization and Interpretation of an Important Prehistoric Source, British Archaeological Reports, British Series 476, Oxford: Archaeopress. Barad, K. (1999) ‘Agential realism: feminist interventions in understanding scientific practices’, in M. Biagioli (ed.) The Science Studies Reader, London: Routledge, 1–11. Barad, K. (2007) Meeting the Universe Halfway: Quantum Physics and the Entanglement of Matter and Meaning, Durham, NC: Duke University Press. Barrett, J. (1994) Fragments from Antiquity: An Archaeology of Social Life in Britain, 2900–1200 BC, Oxford: Blackwell. Barrett, J. (2010) Vibrant Matter: A Political Ecology of Things, Durham, NC: Duke University Press. Binford, L. R. (1983) Working at Archaeology, New York: Academic Press. Bradley, R. (1997) Rock Art and the Prehistory of Atlantic Europe: Signing the Land, London: Routledge. Braidotti, R. (2013) The Posthuman, Cambridge: Polity Press. Butter, R. (1994) Kilmartin: Scotland’s Richest Prehistoric Landscape, Kilmartin: Kilmartin House Trust. Clarke, D. L. (1973) ‘Archaeology: the loss of innocence’, Antiquity, 47(185): 6–18. Conneller, C. (2011) An Archaeology of Materials: Substantial Transformations in Early Prehistoric Europe, London: Routledge. Coole, D. and Frost, S. (2010) ‘Introducing the new materialisms’, in D. Coole and S. Frost (eds) New Materialisms. Ontology, Agency and Politics, Durham, NC: Duke University Press, 1–43. Della Rocca, M. (2008) Spinoza, London: Routledge. Edgeworth, M. (2012) ‘Follow the cut, follow the rhythm, follow the material’, Norwegian Archaeological Review, 45(1): 76–92. Fowler, C. (2013) The Emergent Past. A Relational Realist Archaeology of Early Bronze Age Mortuary Practices, Oxford: Oxford University Press. Gell, A. (1998) Art and Agency: An Anthropological Theory, Oxford: Clarendon. Gosden, C. (2001) ‘Making sense: archaeology and aesthetics’, World Archaeology, 33(2): 163–167. Haraway, D. (2008) When Species Meet, Minneapolis, MI: University of Minnesota Press. Henare, A., Holbraad, M. and Wastell, S. (2007) Thinking through Things: Theorising Artefacts Ethnographically, London: Routledge. Hodder, I. (1992) Theory and Practice in Archaeology, London: Routledge. Hodder, I. (2012) Entangled: An Archaeology of the Relationships between Humans and Things, Malden, MA: Wiley-Blackwell. Holbraad, M. (2009) ‘Ontology, ethnography, archaeology: an afterword on the ontography of things’, Cambridge Archaeological Journal, 19(3): 431–441. Ingold, T. (2007) ‘Materials against materiality’, Archaeological Dialogues, 14: 1–16. 338 Andrew Meirion Jones

Ingold, T. (2013) Making: Anthropology, Archaeology, Art and Architecture, London: Routledge. Jones, A. (2002) Archaeological Theory and Scientific Practice, Cambridge: Cambridge University Press. Jones, A. (2005) ‘Between a rock and a hard place: rock art and mimesis in Neolithic and Bronze Age Scotland’, in V. Cummings and A. Pannett (eds) Set in Stone: New Approaches to Neolithic Monuments in Scotland, Oxford: Oxbow, 107–117. Jones, A. (2006) ‘Animated images: images, agency and landscape in Kilmartin, Argyll, Scotland’, Journal of Material Culture, 11: 211–225. Jones, A. (2011) ‘Coda’, in A. M. Jones, D. Freedman, B. O’Connor, H. Lamdin- Whymark, R. Tipping and A. Watson (eds) An Animate Landscape: Rock Art and the Prehistory of Kilmartin, Argyll, Scotland, Oxford: Windgather Press, 324–333. Jones, A. (2012a) ‘Living rocks: animacy, performance and the rock art of the Kilmartin region, Argyll, Scotland’, in A. Cochrane and A. M. Jones (eds) Visualising the Neolithic, Oxford: Oxbow, 79–88. Jones, A. (2012b) Prehistoric Materialities: Becoming Material in Prehistoric Britain and Ireland, Oxford: Oxford University Press. Jones, A. M. and Tipping, R. (2011) ‘From geology to microtopography’, in A. M. Jones, D. Freedman, B. O’Connor, H. Lamdin-Whymark, R. Tipping and A. Watson (eds) An Animate Landscape: Rock Art and the Prehistory of Kilmartin, Argyll, Scotland, Oxford: Windgather Press, 12–35. Jones, A. M., Freedman, D., O’Connor, B., Lamdin-Whymark, H., Tipping, R. and Watson, A. (eds) (2011) An Animate Landscape: Rock Art and the Prehistory of Kilmartin, Argyll, Scotland, Oxford: Windgather Press. Jones, A. M., Alberti, B., Fowler, C. and Lucas, G. (2013) ‘Archaeology after Interpretation’, in B. Alberti, A. M. Jones and J. Pollard (eds) Archaeology after Interpretation, Walnut Creek, CA: Left Coast Press, 15–35. Lamdin-Whymark, H. (2011) ‘Lithics, landscape and performance’, in A. M. Jones, D. Freedman, B. O’Connor, H. Lamdin-Whymark, R. Tipping and A. Watson (eds) An Animate Landscape: Rock Art and the Prehistory of Kilmartin, Argyll, Scotland, Oxford: Windgather Press, 176–201. Lucas, G. (2012) Understanding the Archaeological Record, Cambridge: Cambridge University Press. Pauketat, T. (2007) Chiefdoms and Other Archaeological Delusions, New York: Altamira Press. Porter, R. (2001) The Enlightenment, Basingstoke: Palgrave. Randall-McIver, D. (1933) ‘Archaeology as a science’, Antiquity, 7: 5–20. Schnapp, A. (1993) The Discovery of the Past, trans. I. A. Kinnes and G. Varndell, London: British Museum Press. Shanks, M. and Hodder, I. (1995) ‘Processual, postprocessual and interpretative archaeologies’, in I. Hodder, M. Shanks, A. Alexandri, V. Buchli, J. Carman, J. Last and G. Lucas (eds) Interpreting Archaeology: Finding Meaning in the Past, London: Routledge, 3–29. Shanks, M. and Tilley, C. (1987) Reconstructing Archaeology, Cambridge: Cambridge University Press. Thomas, J. (2004) Archaeology and Modernity, London: Routledge. Tilley, C. (1993) ‘Interpretation and a poetics of the past’, in C. Tilley (ed.) Interpretative Archaeology, Oxford: Berg, 1–30. Watson, P. J., Le Blanc, S. A. and Redman, C. L. (1971) Explanation in Archaeology, New York: Columbia University Press. Whatmore, S. (2008) Hybrid Geographies: Natures, Cultures, Spaces, London: Sage. Whitley, D. S. (2005) Introduction to Rock Art Research, Walnut Creek, CA: Left Coast Press. 19 MATTER AND FACTS

Material culture and the history of science

Simon Werrett

Historians of science interested in material culture, among whom I would count myself, often gripe about the tendency of other historians to present science as if it was really all about theory, a history of ideas and great geniuses, when really material culture and practice are essential to scientific practice. I assumed, given that material culture is the focus, that archaeologists would share this gripe. So I was surprised to read on the first page of Colin Renfrew and Paul Bahn’s textbook Archaeology Essentials: Theories, Methods, and Practice, ‘The history of archeology is . . . in the first instance a history of ideas, of theory, of ways of looking at the past’ (Renfrew and Bahn 2007: 13). For a discipline so powerfully evocative of material culture to a novice like me, it was a surprise to find that objects did not figure in this defining statement. Historians of science today would strongly disagree with such an interpretation. They would want to say that ideas and theory are not the only or even principal features of any science, especially not archaeology! This is not meant as a rant against archaeology textbooks, but more as an illustration of what the history of science might bring to the table for archaeologists. I think the approach of historians of science offers a careful, detailed and ultimately empirical understanding of the history of archaeology, one that avoids grand narratives of continual progress and the evolution of a properly scientific atti- tude in the discipline. Some studies, such as those of Knossos and the Tomb of Agamemnon by Cathy Gere, have already demonstrated the richness of integrating the two disciplines (Gere 2007, 2011). At the same time, it is abundantly clear that the history of science has plenty to learn from archaeology. In recent times, historians have become increasingly interested in the history of the material dimensions of scientific practice and in the use of material culture as a source for historical inquiry. But if historians have embraced material culture as evidence, they have only done so with a relatively limited disciplinary repertoire, which archaeology could greatly enrich. This chapter charts how the history of science has engaged with material 340 Simon Werrett culture, and considers how archaeological perspectives might expand this approach for both the history of science and the history of archaeology. The chapter begins with a brief exploration of how historians of science shifted from a focus on intellectual history to an interest in the material culture of science in the last decades of the twentieth century. Subsequent sections explore areas of material culture that are of particular interest to historians of science, and how these have revealed new insights into the nature of scientific evidence and the changing conditions by which evidence is identified, judged and assessed. Historians have explored the role of the human body in making knowledge and have used re-enactments and re-stagings of past experiments to reveal this. They have investigated how instruments and models came to be regarded (or disregarded) as legitimate sources of scientific evidence, and they have opened up studies of the more general roles material culture might play in the sciences. The chapter then turns to an exploration of archaeological studies of scientific material culture and identifies some of the ways archaeology might contribute to extending the use of materials as evidence in the history of science.

From the history of ideas to embodiments of knowledge Historians of science have always been interested in material culture, but encouraged a more idealist image of science in the middle decades of the twentieth century, which the discipline is now trying to move away from. Some of the earliest histories of science considered the material conditions of scientific production. In the eighteenth century, the chemist Joseph Priestley made experiment the object of historical inquiry, cataloguing past experiments in electricity and optics to avoid future research repeating past studies (Christie 1990: 9–10). In the nineteenth century, historians of science included regular references to scientific apparatus in their works, while museum curators exhibited on the history of instruments (Whewell 1837; Anon. 1876). While curators remained interested in scientific instruments, some historians of science took a more idealist turn in the twentieth century. For some, the Cold War politicized questions about the place of the material in science. In the revolutionary turmoil of the 1920s, the Russian physicist Boris Hessen proposed that scientific theories, in this case Isaac Newton’s theory of gravitation, should be understood as products of the economic and social conditions prevailing in a particular time and place, reflecting the Marxist view that culture is a superstructural product of the conditions of production (Hessen 1931). Russian Marxists inspired admirers and critics in the West. Critics countered with an ‘internalist’ history of ideas in which scientific theories and ideas predominated. Alexander Koyré, a former White soldier in the Russian revolution and a leading light in this more intellectual history, dismissed the Marxist view: ‘I do not see what the scientia activa has ever had to do with the development of the calculus, nor the rise of the bourgeoisie with that of the Copernican, or the Keplerian, astronomy’ (Koyré 1968: 6). Ultimately, the progress of ideas was the real key to the history of science. Matter and facts: material culture and the history of science 341

This emphasis on the history of ideas endured in the history of science through the 1970s. Change came as historians looked to new resources in sociology and anthropology, which revived more contextual and social understandings of science, recalling those of the Marxists earlier in the century. These inspired more interest in material culture and the study of science as a physical, embodied practice. One resource for this was the ‘sociology of scientific knowledge’ (SSK) arising in the 1970s, which emphasized the ways theories and artifacts were socially constructed by the needs and expectations of communities of interacting users and makers (Law 2010). Some strands of SSK lent particular importance to material culture and practice as evidence. While historians’ evidence base often remained exclusively textual, sociologists went to laboratories and observed scientists as a practice (Latour 1987; Pickering 1992). Historians inspired by SSK followed suit by paying more attention to the physical contexts of past science, wedding this to sociological interpretations of scientific change. For example, in Leviathan and the Air-Pump, Steven Shapin and Simon Schaffer (1985) explored the rise of experimental science through the career of a seventeenth-century scientific instrument, the air pump of Robert Boyle, and controversies over its value to science. By the late 1980s, material culture had become a more significant focus for historians of science. In this period, an interest in the context of scientific activity grew into the ‘spatial turn,’ exploring the physical locations and conditions of knowledge-making in laboratories, museums, the field and other sites (Shapin and Ophir 1991; Livingstone 2003). The spatial turn stressed that ideas and theories should be understood as always being embodied in material objects, from human bodies to books, papers, instruments, and other ‘immutable mobiles,’ as sociologist Bruno Latour called them (Law 2010: 181–182). From this point of view, science is less a collection of abstract notions and concepts emerging from the minds of individuals, than a concrete, collective practice of producing, assessing and circulating material objects designated as carriers of natural knowledge. By the close of the twentieth century, then, many historians came to see science as an embodied practice involving primarily the manipulation of things.

Bodies of evidence One of the most important arenas for science is the human body, and much recent history of science has been concerned with the role of the body in scientific practice. Bodies are stores of information, gestures and skills learned over many hundreds or thousands of hours, and ingrained into nerves, muscles and bones. Historically, epistemic judgments have been radically shaped by the capacities and judgments of the human body. As Shapin and Christopher Lawrence have argued, the bodies of knowledge-makers are intimately tied to their identities, and hence their reliability as truth-tellers, while ‘The bodily identity of the truth-seeker undoubtedly varies from culture to culture and from time to time within a culture’ (Lawrence and Shapin 1998: 10; Schaffer 1994). The credibility of past evidence hinges on the credibility of those who assess it, and their credit hinges, in part, on 342 Simon Werrett their self-representation through bodily form, acts and appearances. The most obvious instance in which embodiment has been linked to epistemic authority is the case of women, who for centuries were excluded from science on the grounds that their bodies disabled them from being reliable truth-tellers in comparison to men (Schiebinger 1989). Male credibility also varied according to bodily forms and gestures. Asceticism, signaled through bodily acts of physical restraint, was long taken to mark a willed disengagement from the world and was associated with the reliable knower, in the figure of the hermit, the chaste or temperate scholar or the self-effacing seeker of objectivity (Shapin 1998; Daston and Galison 2007: 191–252). Identifying the embodied nature of evidential assessments has been accompanied by innovative ways of making history more performative. Historians of science now increasingly share with archaeology an interest in re-stagings and re-enactments of past practices as a source of evidence for enriching historical understanding. The extent of these practices has been much greater in archaeology than in the history of science. While archaeologists routinely use ethnoarchaeology and ethnographic analogy to see how past techniques might have been achieved, historians of science have only relatively recently begun to engage in similar investigations. The epistemic status of this work differs in the history of science compared to archaeology. While archaeologists have typically framed re-enactments as scientific experiments, occurring within science as a means of increasing knowledge of the past, historians have for the most part framed their re-enactments as historical inquiries into the nature of science itself. Re-enactments in the history of science have a long history (Chang 2011). In the 1960s, Thomas B. Settle of Cornell University investigated Galileo’s theories of motion by re-staging his experiments (Settle 1961). In the 1990s, Peter Heering used replications to study Coulomb’s inverse-square law (Heering 2006; Heering and Wittje 2011). Since the early 1990s, H. Otto Sibum has re-staged numerous physics experiments from the past, including those undertaken by the Manchester physicist James Joule in the mid- nineteenth century to calculate the mechanical equivalent of heat (Sibum 1995). Working from a replica of Joule’s apparatus and performing careful re-stagings of his experiments, Sibum revealed that the successful practice of Joule’s experiment required skills that were not widely distributed among contemporary physicists, but that Joule had obtained through his experience as a brewer (his family owned and ran a brewery in Manchester). Sibum has highlighted the essential role of ‘gestural knowledge’ in science, or ‘the complex of skills and forms of mastery developed in . . . real-time performances’ in experimentation, and the diverse communities that such knowledge might come from (Sibum 1995: 76). He also reveals an important tension in innovative experiments between the highly individualized skills that may be necessary to make new experiments succeed (in this case, Joule’s skills as a brewer) and the need for experiments to be sufficiently deskilled to allow them to be replicated by anyone in the scientific community. For those who could not replicate Joule’s embodied skills, his experiment did not count as evidence of a mechanical equivalent of heat. Matter and facts: material culture and the history of science 343

More recently, historians of chemistry and alchemy have turned to re-stagings to make sense of complex chemical and alchemical processes from the past. Jenny Rampling, Lawrence Principe and William Newman, for example, have explored chemical processes and reactions that may have lay behind the recipes of renaissance and early modern alchemical texts (Principe 1987, 2012). Once a historian is able to witness alchemical processes re-created in the laboratory, they argue, the intentional obscurities of alchemical writing may be more easily judged and deciphered. Alternatively, Hasok Chang has used the re-staging of past experiments to recover lost knowledge and to explore alternative directions for scientific experiments that earlier investigators may not have taken, an approach Chang calls ‘complementary science’ (Chang 2004). The emphasis on the material in histori- ography has thus translated into a growing practical engagement with material evidence as a source for historical research, beyond the traditional emphasis on using texts.

Instruments and models This historical work shows that in the sciences, evidence needs to be both engaged with the body to be credible and appropriately distanced from it. The body is necessary to make evidence, but if evidence is too embodied in the scientist, or the scientist’s body is considered problematic in some way, then the evidence cannot count. This is also apparent in the use of instruments. While historians agonized over the relative merits of intellectual and material history in the late twentieth century, curators working on the history of scientific instruments were already very familiar with issues of material culture and questions of practice and evidence (Van Helden and Hankins 1994; Taub 2011). As interest in material culture grew, so this work became more integrated into the history of science. This work demonstrated that, like the body itself, the integrity of instruments plays a critical role in assessments of evidence. What counts as evidence depends on different notions of what count as valid instruments, and how those instruments interact with the bodies of their users. In some cases, these have changed radically over time. In the seventeenth century, instruments, such as the telescope, microscope and air pump, were adopted in the sciences because they were viewed as extensions to the senses that would allow humans to restore the extraordinary perceptive powers of Adam and Eve in the Garden of Eden before the Fall (Harrison 2002). Throughout the seventeenth century, debates raged about the validity of instruments as sources of scientific knowledge. Critics rejected the status of microscopic evidence as scientific on the grounds that simply magnifying objects taught nothing about human values, the essential feature of any investigation claiming to be a science (Wilson 1988). The legitimacy of evidence obtained by instruments is closely related to judgments about the capacities of the human body. In their history of objectivity, Lorraine Daston and Peter Galison (2007) have shown how instruments came to be seen in the nineteenth century as a way to remove the body, understood as a 344 Simon Werrett repository of weaknesses, personal prejudices and idiosyncrasies, from experimental inquiry. But, even this ‘mechanical objectivity’ did not endure as a credible approach in the sciences (Daston and Galison 2007). In recognition that the human senses could never be entirely removed from acts of observation, some scientists sought to exclude representations and observations entirely from science, in favor of logic and mathematics. Others settled on ‘trained judgment’ or an appropriately disciplined bodily engagement in acts of observation and experimentation, bringing a human expertise into combination with instrumental renderings of natural phenomena. Evidential assessments thus depended on changing ideas of the appropriate interrelation of the body, instruments and objects, and the relative credibility of their different configurations. How these configurations are judged will depend on changing local social contexts and interests. One consequence of the rise of historical studies interested in materials and practice has been to expand the definition of what counts as scientific material culture, well beyond the traditional focus on scientific instruments. Models, for example, have been the subject of many recent studies. Archaeologists have long employed scale models of buildings, ships, cities, sites and excavations to make better sense of them for professionals and the public. Historians dismissed models for a long time as mere teaching aids or museum displays, but as Soraya de Chadarevian and Nick Hopwood have shown recently, scientific models have moved constantly between contexts of research, display and education and are often crucial to all three (Chadarevian and Hopwood 2004; Maerker 2011; Bud 2013). Like scientific instruments, the status of models as legitimate sources of evidence has been widely debated in history (Schaffer 2004a). Christopher Evans has drawn attention to the fragility of archaeological models as a form of evidence (Evans 2004). A large watercolor of a ‘druidical temple’ at Mont St Helier, Jersey, belonging to nineteenth-century architect Sir John Soane, was likely made not from the original site but by copying a model held in the Society of Antiquaries. Evans asks how many other early images of sites of ancient remains were made from models rather than from life? Rusticated textures on stone may be more to do with the cork used to make models than textures on the original stones. Models often existed ambiguously between objective representations and entertaining spectacles in the nineteenth century, deploying lighting effects, magic lantern projections and paintings to give atmosphere and presence to the representation of a site. Models remained prominent in archaeological conference presentations into the twentieth century, shifting from representations of ancient buildings or stone to models of excavations and then reconstructions. Models mediated archaeology to the public, but may also have served to legitimate concepts of evidence in archaeological practice. Pitt Rivers had more than a hundred wooden models of his excavations constructed for public education, and Evans supposes their unprecedented attention to detail reflected, and helped promote, Pitt Rivers’ demand for the use of a broader range of evidence in archaeology. The history of instruments and models demonstrates how the legitimacy of different forms of material culture as evidence in a science may take much time and Matter and facts: material culture and the history of science 345 effort to establish. Lorraine Daston has proposed that a feature of scientific objects is that they are partly defined by the duration and difficulty of constituting them as legitimate objects of inquiry. Daston contrasts the solid ‘quotidian’ objects of common experience, which ‘possess the self-evidence of a slap in the face’ with ‘scientific objects’: typically unfamiliar entities, such as atoms, genes or forces, which take a long time to capture and may or may not prove enduring as things (Daston 2000: 2). More often perhaps, objects may be simultaneously quotidian and scientific, physically robust but fragile in their status as evidence of a theory or interpretation.

Nature, artifice and culture Historical inquiry thus reveals how the apparently straightforward status of materials and objects as evidence may not always have been so obvious. Another place where this becomes apparent is the boundary between the natural and the artificial, of crucial significance in archaeology. It is a boundary that historians have shown to be changing constantly in history (Newman and Bensaud-Vincent 2007). A striking case is provided by a cameo examined by the German monk Albertus Magnus in the thirteenth century, depicting the carefully sculpted heads of two Romans in plumed helmets. As Lorraine Daston has pointed out, while the cameo appears to us as an obviously human artifact, Albertus took it to be a natural object, a product of the personified deity Nature at play (Daston 1998). To see it in this way entailed belonging to a Christian, European culture in which the natural world was the production of a deified Nature acting as vice-regent to the creator God. Cameos, crystals, fossils and other objects were identified as exceptional ‘jokes,’ which Nature created when not producing the normal realm of things (Findlen 1990). Different times and cultures thus draw the boundary of artifice and nature differently. Today, changing boundaries are signaled by debates over objects that seem to lie ambiguously between the natural and artificial, such as genetically engineered organisms, artificial intelligence or, in the case of archaeology, the ‘geofacts’ at Pedra Furada in South America (Meltzer 1995). Recognizing the very different ways that material objects have appeared to scientific inquiry in the past was the goal of a 2004 collection of essays edited by Lorraine Daston (2004) entitled Things That Talk, which expanded historical studies from a focus on the instruments of scientific inquiry to a broader exploration of certain classes of ‘things’ in general. Daston and her collaborators sought in this work to reveal how objects could be the source of multiple meanings and cultural significances, emerging in particular historical moments as evocative and fascinating artifacts. Daston is especially interested in things that bridge the divide between those prompting a singular, unproblematic ‘positivist’ understanding and those deemed open to endless interpretation: between ‘evidence’ and ‘idol,’ as she puts it. Such things, with examples including photographs, bubbles, Rorschach tests and glass flowers, are shown to have evoked new debates and ideas to the degree that their ‘thingness’ or materiality did not fit into preconceived boundaries and 346 Simon Werrett sensibilities. Bubbles, for example, were transformed in the late nineteenth century from transient things into fixed commodities, objects of knowledge and icons of visual culture. They ‘talked’ by evoking not only the facts of their constitution but also the renderings of artists, entrepreneurs and engineers (Schaffer 2004b). Things That Talk was representative of a new direction in the history of science, away from the study of the instruments of science and toward the study of science as one among a variety of artistic, commercial and technical material cultures. Another work that has been important for developing this approach is the 2010 collection Materials and Expertise in Early Modern Science, edited by Ursula Klein and Emma Spary (Klein and Spary 2010; Klein and Lefèvre 2007). Like Things That Talk, Klein and Spary’s collection evokes the diverse meanings given to substances in history, and stresses their location at the intersection of multiple discourses and practices. The authors examine how substances such as dyes, ceramics, milk, ink, spa water and gunpowder formed the focus of new interactions between govern- ment, the market, consumers and the sciences in the sixteenth to eighteenth centuries, constituting new hybrids of expertise and reshaping knowledge. The approach of this work and Daston’s Things That Talk has not been without its critics. The historians and curators Thomas Söderqvist and Adam Bencard have criticized Daston for characterizing material objects as ‘talking,’ which, even if not meant literally, shifts the focus from the physical to the semiotic. This does objects’ materiality a disservice, they argue, by studying it as nothing more than a special version of a text (Söderqvist and Bencard 2010). Nevertheless, the material turn that Daston, Klein, Spary and others have initiated represents a significant shift in the way historians of science have related to material culture.

Archaeology and material histories of science Lest it appear that historians have had a lot to say about material culture, it is worth stressing that a great deal of work in the discipline remains focused on the history of ideas, while most of those who have engaged with material culture (including myself ) have tended to do so by reading about things rather than engaging with them directly. This is certainly not true of all studies. Museologically inspired work on scientific instruments, for example, has a long tradition of working with objects (Bennett 1987; Bud 2013). But actual physical and material objects and their properties have not been used as evidence in the history of science as much as might be wished. I will argue in this final section that archaeology offers great resources for historians to expand their engagements with materials. Archaeology itself contributed much to the history of science in the twentieth century, particularly in excavations of chemical and alchemical sites in the post- war era. These include the castle of Oberstockstall in Austria, the Louvre in Paris and the Old Ashmolean in Oxford (Osten 1998; Rouaze 1989; Bennett et al. 2000; Moorhouse 1972; Anderson 2000; Martinón-Torres 2007). This has revealed information that complements historical research and has offered new insights that textual studies alone could not provide. Scholars such as Robert Anderson have Matter and facts: material culture and the history of science 347 pointed out the value of archaeology in this regard. While drawings and archi- tectural plans can show the situation and arrangement of laboratory spaces, exca- vations reveal the range and nature of instruments, apparatus and substances in use inside them (Klein 2008: 775–777). Excavated instruments may also speak of practice in ways that texts cannot. Looking at material objects directly avoids the problem that textual descriptions may be inaccurate because they were copied and distorted over many years. Texts might also be deliberately obscure, as in the case of alchemy. Texts often lack details of structure and process that a direct experience of apparatus makes evident (Anderson 2000: 24). Objects, unlike texts, can reveal their location of origin, showing how materials circulated around Europe (or the world) at different times. The locations of objects also make apparent the kinds of places where they were in use (as Anderson notes, alchemical vessels have typically appeared in excavations of glass-houses, castles and monasteries, which might not have otherwise been obviously scientific sites) (Anderson 2000: 18). Archaeology reveals further useful information, depending on what it fails to find. If certain objects are scarce it may be because they were not as extensively utilized as historians might have thought, based on the textual record. The lack of metal objects in excavations of alchemical apparatus, despite their evident frequency in alchemical texts, suggests that metals were melted down and recycled, whereas glass was cheap enough to throw away and ceramics could not be used again if substances fused into them (Anderson 2000: 25). The relationship between the history of science and archaeology will change as historians and archaeologists reconsider what count as scientific sites. Anderson points out that an adequate history of the chemical laboratory should include not only alchemical or chemical research laboratories but also places of chemical production, testing and teaching (Anderson 2013: 671–672). Studies of excavations showing, for example, evidence of glass production for pharmaceutical vessels in Fustat, Cairo, then become relevant to a history of chemical laboratories, taking that history back much further than is normally the case (Scanlon and Pinder- Wilson 2001; Bacharach 2002). Archaeology may also help to manifest the degree to which past scientific spaces were adapted from existing buildings, changing conceptions of what counts as a scientific space. Kitchens, bedrooms and cellars in homes, in addition to chapels and barracks have all been adapted to undertake experimental research. Excavations might provide evidence of other adapted spaces (Anderson 2013: 673; Werrett 2013). Archaeology has also helped to reveal the facility and boundaries of past methods of natural inquiry. Recently, Marcos Martinón-Torres from University College London has led a team examining early modern laboratory remains in Europe and America. This work has made manifest the practicality of alchemy and the specific techniques and instruments that it entailed. Martinón-Torres, whose work represents a thoroughly interdisciplinary combination of history of science and archaeology, has shown that alchemy was not just an esoteric search for gold but incorporated mining, assaying and metallurgy more generally. He suggests that studies of remains show that alchemists worked with materials previously thought 348 Simon Werrett to have only been synthesized in the twentieth century (Pinkowski 2004). Martinón-Torres has also addressed the question of how archaeology and the history of science might be better integrated. He has stressed the value of archaeometric techniques for examining processes of manufacturing and using laboratory instruments to reveal international networks of production, distribution and collaboration. He notes that makers’ marks can reveal unexpected connections, for example showing combinations of alchemical and masonic imagery on early modern objects (Martinón-Torres 2010, 2011). Medical archaeology also provides insights normally overlooked by historians working only with texts. Patricia Baker has pointed to the way archaeology enriches the geography of ancient Roman medicine. While texts may make practices appear homogeneous across the ancient Roman empire, excavations provide evidence of regional and local differences. Furthermore, considering contexts where Roman medical instruments have been found discloses much about their non-functional uses. Medical instruments found in rivers and graves suggest their use as votive offerings, samples of property or as mediating objects between life and death (Baker and Carr 2002; Baker 2004a, 2004b). Daniel Antoine notes that medical archaeology also offers a diverse array of resources for understanding the history of epidemics. Archaeological approaches may make possible the identification of disease hosts and vectors, reconstructions of the natural and urban environments and climatic conditions in which diseases spread between individuals, communities or populations. Archaeology has also identified and dated burial sites of epidemic victims, and sheds light on the material practices and culture around burial or the disposal of bodies (Antoine 2008).

Conclusion These brief examples indicate the potential richness of archaeological approaches for developing further the history of science’s engagements with material culture. This chapter has explored how material culture became a focus in the history of science, what historians have been able to learn about the status of material culture as evidence and how archaeology offers new resources and perspectives for taking these studies further. Historians of science have taken a growing interest in material culture in recent decades. They have shown how changing assessments and constructions of the human body and the material culture of buildings, instruments, models and other apparatus have been deeply implicated in defining what counts as scientific evidence and how it should be judged. Material evidence is always assessed within a complex of understandings of the nature of materials themselves (as natural or artificial, for example), about the status of materials as valid forms of evidence and about the bodies of investigators and their credibility in making natural knowledge. These criteria have been constantly changing, so that what might appear as unassailable evidence in one context may soon become irrelevant to scientific knowledge in another. Judgments on all these matters are culturally and historically situated and variable. Matter and facts: material culture and the history of science 349

If the history of science has highlighted the historicity and fragility of material evidence, it nevertheless has much to learn if it is fully to appreciate the role of the material in the sciences. Against all that fragility is the hard materiality of things themselves. If, recalling the introduction to this chapter, Renfrew and Bahn (2007) overplay the significance of ideas in archaeology, then historians of science have surely underplayed the value of archaeology for their enterprise. As historians venture into a more general and broad appreciation of the material in science, archaeology offers them the opportunity to learn beyond the text, revealing new insights into the diverse uses, distributions, circulations and values given to materials and artifacts in the past. If things can talk, archaeology may help historians to listen more closely to what things have to say.

Acknowledgments Thanks to the editors and all the participants of 2010 Reading conference ‘Material Culture as Evidence’, to Jenny Rampling, Hasok Chang and to the members of the Ad Hoc History of Chemistry Reading Group at UCL for their comments and suggestions for this essay.

References Anderson, R. G. W. (2000) ‘The archaeology of chemistry,’ in F. L. Holmes and T. H. Levere (eds) Instruments and Experimentation in the History of Chemistry, Cambridge, MA: MIT Press. Anderson, R. G. W. (2013) ‘Chemistry laboratories, and how they might be studied,’ Studies in History and Philosophy of Science, 44: 669–675. Anon. (1876) ‘Loan collection of scientific apparatus,’ The Athenaeum, 2538 (17 June): 832–833. Antoine, D. (2008) ‘The archaeology of “plague,”’ Medical History. Supplement, 27: 101–114. Bacharach, J. L. (ed.) (2002) Fustat Finds: Beads, Coins, Medical Instruments, Textiles, and Other Artifacts from the Awad Collection, Cairo: American University in Cairo Press. Baker, P. A. (2004a) ‘Roman medical instruments: archaeological interpretations of their possible “non-functional” uses,’ Social History of Medicine, 17: 3–21. Baker, P. A. (2004b) Medical Care for the Roman Army on the Rhine, Danube and British Frontiers from the First Through Third Centuries AD, British Archaeological Reports International Series 1286, Oxford: Hadrian Books. Baker, P. A. and Carr, G. (eds) (2002) Practitioners, Practices and Patients: New Approaches to Medical Archaeology and Anthropology, Oxford: Oxbow Books. Bennett, J. A. (1987) The Divided Circle: A History of Instruments for Astronomy, Navigation and Surveying, Oxford: Phaidon Christie’s. Bennett, J. A., Johnston, S. A. and Simcock, A. V. (2000) Solomon’s House in Oxford: New Finds from the First Museum, Oxford: Museum of the History of Science. Bud, R. (2013) ‘Life, DNA, and the model,’ British Journal for the History of Science, 46: 311–334. Chadarevian S. de, and Hopwood, N. (eds) (2004) Models: The Third Dimension of Science, Stanford, CA: Stanford University Press. Chang, H. (2004) Inventing Temperature: Measurement and Scientific Progress, Oxford Studies in the Philosophy of Science, New York: Oxford University Press. 350 Simon Werrett

Chang, H. (2011) ‘How historical experiments can improve scientific knowledge and science education: the cases of boiling water and electrochemistry,’ Science and Education, 20: 317–341. Christie, J. R. R. (1990) ‘The development of the historiography of science,’ in R. C. Olby, G. N. Cantor, J. R. R. Christie and M. J. S. Hodge (eds) Companion to the History of Modern Science, London: Routledge. Daston, L. (1998) ‘Nature by design,’ in C. A. Jones and P. Galison (eds) Picturing Science, Producing Art, New York and London: Routledge, 232–253. Daston, L. (ed.) (2000) Biographies of Scientific Objects, Chicago, IL: University of Chicago Press. Daston, L. (ed.) (2004) Things That Talk: Object Lessons from Art and Science, New York: Zone Books. Daston, L. and Galison, P. (2007) Objectivity, New York: Zone Books. Evans, C. (2004) ‘Modelling monuments and excavations,’ in S. de Chadarevian and N. Hopwood (eds) Models: The Third Dimension of Science, Stanford, CA: Stanford University Press, 109–137. Findlen, P. (1990) ‘Jokes of nature and jokes of knowledge: the playfulness of scientific discourse in Early Modern Europe,’ Renaissance Quarterly, 43: 292–331. Gere, C. (2007) The Tomb of Agamemnon: Mycenae and the Search for a Hero, London: Profile. Gere, C. (2011) Knossos and the Prophets of Modernism, Chicago, IL: University of Chicago Press. Harrison, P. (2002) ‘Original sin and the problem of knowledge in Early Modern Europe,’ Journal of the History of Ideas, 63: 239–259. Heering, P. (2006) ‘Regular twists: replicating Coulomb’s wire-torsion experiments,’ Physics in Perspective, 8: 52–63. Heering, P. and Wittje, R. (eds) (2011) Learning by Doing: Experiments and Instruments in the History of Science Teaching, Stuttgart: Franz Steiner Verlag. Hessen, B. (1931) ‘The social and economic roots of Newton’s Principia,’ in N. I. Bukharin (ed.) Science at the Crossroads, Papers Presented to the International Congress of the History of Science and Technology Held in London from June 29th to July 3rd, 1931, by the Delegates of the USSR, London: Kniga Ltd, 149–212. Klein, U. (2008) ‘The laboratory challenge: some revisions of the standard view of early modern experimentation,’ Isis, 99: 769–782. Klein, U. and Spary, E. (eds) (2010) Materials and Expertise in Early Modern Science, Chicago, IL: University of Chicago Press. Klein, U. and Lefèvre, W. (2007) Materials in Eighteenth-Century Science: A Historical Ontology, Cambridge, MA: MIT Press. Koyré, A. (1968) Newtonian Studies, Chicago, IL: University of Chicago Press. Latour, B. (1987) Science in Action, Milton Keynes: Open University Press. Law, J. (2010) ‘The materials of STS,’ in D. Hicks and M. C. Beaudry (eds) The Oxford Handbook of Material Culture Studies, Oxford: Oxford University Press, 173–188. Lawrence, C. and Shapin, S. (1998) ‘Introduction: the body of knowledge,’ in C. Lawrence and S. Shapin (eds) Science Incarnate: Historical Embodiments of Natural Knowledge, Chicago, IL: The University of Chicago Press, 1–20. Livingstone, D. N. (2003) Putting Science in Its Place: Geographies of Scientific Knowledge, Chicago, IL: University of Chicago Press. Maerker, A. (2011) Model Experts: Wax Anatomies and Enlightenment in Florence and Vienna, 1775–1815, Manchester: Manchester University Press. Martinón-Torres, M. (2007) ‘The tools of the chymist: archaeological and scientific analyses of early modern laboratories,’ in L. M. Principe (ed.) Chymists and Chymistry: Studies in the History of Alchemy and Early Chemistry, Sagamore Beach, MA: Science History Publications and Chemical Heritage Foundation, 149–163. Matter and facts: material culture and the history of science 351

Martinón-Torres, M. (2010) ‘Of marks, prints, pots and Becherovka: freemasons’ branding in early modern Europe?,’ in A. Bevan and D. Wengrow (eds) Cultures of Commodity Branding, Walnut Creek, CA: Left Coast Press, 213–233. Martinón-Torres, M. (2011) ‘Some recent developments in the historiography of alchemy,’ Ambix, 58: 215–237. Meltzer, D. (1995) ‘Stones of contention,’ New Scientist, 1983 (24 June): 31–35. Moorhouse, S. (1972) ‘Medieval distilling apparatus of glass and pottery,’ Medieval Archaeology, 16: 79–121. Newman, W. and Bensaud-Vincent, B. (eds) (2007) The Artificial and the Natural: An Evolving Polarity, Cambridge, MA: MIT Press. Osten, S. von (1998) Das Alchemistenlaboratorium Oberstockstall. Ein Fundkomplex des 16. Jahrhunderts aus Niederösterreich, Innsbruck: Universitätsverlag Wagner. Pickering, A. (ed.) (1992) Science as Practice and Culture, Chicago, IL: University of Chicago Press. Pinkowski, J. (2004) ‘The alchemist’s lab,’ Archaeology, 57: 26–31. Principe, L. (1987) ‘“Chemical translation” and the role of impurities in alchemy: examples from Basil Valentine’s triumph-wagen,’ Ambix, 34: 21–30. Principe, L. (2012) The Secrets of Alchemy, Chicago, IL: University of Chicago Press. Renfrew, C. and Bahn, P. (2007) Archaeology Essentials: Theories, Methods, and Practice, London: Thames and Hudson. Rouaze, I. (1989) ‘Un atelier de distillation du moyen age,’ Bulletin archeologique du comite des travaux historiques et scientifiques, 22: 159–271. Scanlon, G. T. and Pinder-Wilson, R. (2001) Fustat Glass of the Early Islamic Period, London: Altajir World of Islam Trust. Schaffer, S. (1994) ‘Self evidence,’ in J. Chandler, A. I. Davidson and H. Harootunian (eds) Questions of Evidence: Proof, Practice, and Persuasion across the Disciplines, Chicago, IL: University of Chicago Press, 56–91. Schaffer, S. (2004a) ‘Fish and ships: models in the Age of Reason,’ in S. de Chadarevian and N. Hopwood (eds) Models: The Third Dimension of Science, Stanford, CA: Stanford University Press, 71–105. Schaffer, S. (2004b) ‘A science whose business is bursting: soap bubbles as commodities in classical physics,’ in L. Daston (ed.) Things That Talk: Object Lessons from Art and Science, New York: Zone Books, 147–194. Schiebinger, L. (1989) The Mind Has No Sex? Women in the Origins of Modern Science, Cambridge, MA: Harvard University Press. Settle, T. B. (1961) ‘An experiment in the history of science,’ Science, 133: 19–23. Shapin, S. (1998) ‘The philosopher and the chicken: on the dietetics of disembodied knowledge,’ in C. Lawrence and S. Shapin (eds) Science Incarnate: Historical Embodiments of Natural Knowledge, Chicago, IL: The University of Chicago Press, 21–50. Shapin, S. and Ophir, A. (eds) (1991) ‘The place of knowledge: the spatial setting and its relation to the production of knowledge,’ special issue of Science in Context, 4: 3–218. Shapin, S. and Schaffer, S. (1985) Leviathan and the Air Pump: Hobbes, Boyle and the Experimental Life, Princeton, NJ: Princeton University Press. Sibum, H. O. (1995) ‘Reworking the mechanical equivalent of heat: instruments of precision and gestures of accuracy in early Victorian England,’ Studies in History and Philosophy of Science, 26: 73–106. Söderqvist, T. and Bencard, A. (2010) ‘Do things talk?,’ in S. Lehmann-Brauns, C. Sichau and H. Trischler (eds) The Exhibition as Product and Generator of Scholarship, Preprint no. 399, Berlin: Max Planck Institute, 93–102. Taub, L. (2011) ‘Introduction: reengaging with instruments,’ Isis, 102: 689–696. 352 Simon Werrett

Van Helden, A. and Hankins, T. L. (1994) ‘Introduction: instruments in the history of science,’ Osiris, 9: 1–6. Werrett, S. (2013) ‘Recycling in early modern science,’ British Journal for the History of Science, 46: 627–646. Whewell, W. (1837) History of the Inductive Sciences from the Earliest to the Present Times, 3 vols, London: John W. Parker. Wilson, C. (1988) ‘Visual surface and visual symbol: the microscope and the occult in early modern science,’ Journal of the History of Ideas, 49: 85–108 INDEX

14C: see radiocarbon dating angle: of incidence 181–182, 184–185, 187 abductive reasoning 257, 273, 277 Anglo-Saxon xi, 48, 227–230, 237 accelerator mass spectrometry (aka AMS) anisotropic costs 178 131, 135, 137–139, 166 Apache 298 accuracy 9, 96, 117, 121, 135, 146, 150, arable plots 243–244, 246, 248–249, 252 166–168, 193, 203, 223, 230, 297 archaeobotany 68, 244, 246–249, 251–252 Adena 164–165 archaeological record: 6, 10, 24, 36, 44, 45, admissibility 273, 284 49, 52, 53, 55, 58, 74, 83, 95, 98, 104, agency 6, 10, 18, 53, 125, 244, 318, 326, 116, 137, 165, 173, 191, 192, 193, 198, 333–334 200, 205, 238, 241, 287, 288, 298, 299, agential realism 10, 326 312, 313–322, 323, 326, 330, 336; agriculture 24, 165–166, 239, 243–244 fragment 7, 15, 46, 313–317, 321–322, Akrotiri x, 145–149 243–346; survival 7, 15, 313–317, alloy 114–118, 121–124, 168, 170 321–322 ambiguity 14, 93, 98, 104–105, 193, 195, archaeological science xvii, xix, 43, 56, 57, 282, 312, 315, 344 121, 125, 126, 154, 157, 160, 162, 163, Amheida (aka Trimithis) 102–106 166, 169, 171, 172, 187, 188, 252, 253 analogy, ethnographic 24, 26, 35, 43, 52, archaeometallurgy 13, 115, 117, 125, 170 55, 117, 121, 163, 238, 288 archaeometry vi, 9, 113, 123, 125, 348 analysis 7–9, 16, 64, 66–72, 79–82, 85, archives 2–3, 23, 36–37, 69, 79, 89–90, 114, 118–119, 129, 134, 151, 160–163, 132, 135, 170, 189, 193, 233, 294, 315 196, 215, 256, 315, 333–335; ard cultivation 243, 245, 248, 251 correspondence 215, 218, 226–228, 230, argument 216–217, 256–265, 269, 236; chemical 13, 15, 71, 114–117, 271–281, 285; evidential vi, 7, 9, 12–14, 159–160, 167–168; discriminant 16–17, 72, 92, 260, 263, 265, 288, 302 249–251; fast track 69; isotope 9, 116, argumentation 193, 273 160–165, 252; post-excavation 48, 64, art of enquiry 196 80–81, 87, 326, 328, 331–333; reflective artefact(s) xviii, 19, 23, 24, 26, 27, 31, 33, 16, 49, 60, 73; scales of 174; specialist 35, 42, 47, 49, 53, 55, 121, 123, 124, 59; statistical 123, 167, 332; strength of 127, 194, 205, 218, 222, 262, 318, 329, 268, 278, 283–284; see also material, 330, 331, 332, 336, 337; classification spatial, and typological analysis 215, 218, 225, 227–228, 230; diagnostic angle: of depression 187 29; re-enactment 51; visitor 51 angle: of elevation 181–182, 185, 187 artisan revolution 195; literacy 194 354 Index asceticism 342 chart method (Wigmore) xii, 14, 259–260, assemblage 10, 13, 66, 67, 69, 72, 77, 97, 263–264, 267–268, 271–276, 279, 117–119, 121–122, 125, 218–219, 282–283 318–319, 321, 331–335; chemical analysis, see analysis archaeobotanical 244, 246, 247, 249, chemistry 7, 13, 24, 44, 114, 117, 124–126, 251; closed 225; grave 227–230 164, 168–169, 343; analytical 114, Assyriology 271, 278–279, 281 123–124; archaeological 123, 127 Avebury 33, 35, 113 Childe, V. G. 95, 129–130 chronology xi, 8, 40, 119, 125–126, Babylonian 279, 281–282 129–130, 134–135, 140–143, 146, background assumptions: 8, 10–14, 16, 148–151, 166, 214, 216–219, 226–229, 25, 44–47, 52, 55–56, 98, 117, 151, 233, 244, 258, 262, 264, 269, 316; 236, 273; generalization 262, 264, 268, archaeological 129, 134, 151, 217, 230; 275 Bayesian 140, 143, 146, 151, 219, 230, Balkesbury house 49 232, 236; ceramic 114, 167; chrono- balsamroot 293, 299, 301 typology 214–215, 222–223, 236; Barad, K. 10, 325–326, 333–335 conventional 13, 114, 149–150; Barker, P. 11, 23–24, 31, 39–40 dendrochronology 47, 129–130, 133, barrows xi, xii, 47, 53, 55, 173–174, 166, 214–215, 238; historical 129, 176–178, 180–182, 237 134–135, 146; models 14, 140, 143, 146, Bayes’ Theorem 140, 217, 272 150, 151, 218–219, 222, 227, 230, 232, Bayesian modeling 10, 14, 140–141, 235–236; radiocarbon 134, 236–237; 143, 146, 151, 219, 223, 227, 230, resolution 10, 141, 147, 148, 149, 158, 232–233, 236, 238; statistics xvii, 215, 238; scale: see timescale(s) 217–218, 236 classification: archaeological 94, 161, 273; behavioural archaeology 44, 55 artefact 96, 102, 117, 167, 222, 315 Binford, L. R. 5–6, 44–45, 55, 95, 288, cluster analysis 123 312, 316–317, 324 collaboration (collaborative process) 9, 46, bioarchaeology 298, 300 63–64, 66–69, 72, 76, 80, 82–83, 160, bootstrapping 11–14, 99, 104, 173 169, 189, 195, 201, 206, 311, 348 British Columbia xix, 293, 295, 297–298 Collingwood, R. G. 25–26 Bronze Age, 8, 33–34, 47, 50, 116, 118, comparanda 14–15, 101–103 120–124, 135, 139–140, 145, 147–149, composite plan ix, 81 222–223, 225–226, 327, 329; Early xi, computer: animation 200; applications in 33, 118, 120–122, 135, 139, 222–223, archaeology 79; technologies 79 225–227; Later 8, 139–140, 145, 147, conflict, 75–76, 79, 81, 135, 192, 217, 244, 148–149, 327, 329 269, 291; resolution 75, 140, 273, 325 Butser Experimental Farm 48–53, 56 congruence 291–292, 295, 299, 302 conjunctions 68, 98, 268 cables, of evidence: see evidence, lines of Conneller, C. 335–336 calibration x, 129–142, 144, 147–149, 151, context 46, 66–77, 80, 102, 106, 116, 126, 167, 216–217, 219, 230; curve 131–133, 132, 138–139, 140–142, 165, 174, 215, 135–140, 147, 149, 230; IntCal13 x, 220, 249, 252, 262–264, 315, 317, 322, 131–133, 136–139, 141, 148 327, 334–336; facsimile record ix, Calleva Attrebatum 82 85–86, 88–89; recording sheets 66, carbon cycle 130; isotopes 164–165 83–84; single context digital plan ix, 59, Carpathian Basin 133 61–62, 66, 70–71, 74, 80–81 Carver, M. 23, 315 contradiction 264, 288, 302 case theory 259 copper: source 116, 118–119, 121; type xv, Çatalhöyük v, ix, 8, 46, 59–66, 68–70, 118–119, 122, 125 73–77, 191, 197–206 correspondence analysis xi, 215, 218, 225, ceramic petrography 162–163 227–230, 236 Chaco Canyon: Great Houses 163 Coulomb, inverse square law 342 Chang, H. 7, 13, 342–343, 349 Cox, G. 201–204, 206 Charcoal Outlier model 143–144 craft: production 76, 196, 198; skills 19 Index 355 craftmanship aesthetic xi, 197 diffusionist paradigm 132–33 craftwork/craftworkers 193 digital: diaries 64, 72, 74, 198; information credibility 9, 190, 278, 283, 302, 341–342, 173–174, 184, 186; media 89–90, 191, 344, 348 196, 199–200, 205–206; models 200; Cree 296–297, 299 pens 82–84, 87–89; publication 262; Criado-Boado, F. 173, 176–178, 180, 182, representation 173; simulation 9, 14, 184, 187 173–175, 178–181, 184, 186–187; Croftmoraig (aka Croft Moraig) ix, 25, 31, technologies 16, 59, 73; visualization vi, 33–38, 40 189, 191–192, 194, 196–198, 201; cultivation: extensive vs. intensive 243–246, working environment 194 249, 251; mobile vs. permanent 44; discoveries 12–13, 55, 94, 96–98, 106, 165, models of 244; transient 44 293, 324, 332 ‘Curve of Knowns’ 128–129 discriminant analysis 249–251 Cycladic, Early period x, 141–145 DNA, ancient 70, 160, 170, 300 dreamings xii, 295–296 Daly, R. 295 Darwin, C. 25–26, 52–53 earthwork ix, 14, 27, 29, 31, 33, 47–48, Daston, L. 1, 3, 10–11, 342–346 53 data: Big Data 167, 169; chemical 170; ecology 4–5, 10, 174, 243–244, 246–251, cleaning 14, 94, 97, 104; digital 173; 291–292, 294, 301, 319; functional entry 72, 74, 82; geographic(al) 122; 247–248, 250–251; weed 246–247, isotopic 115–116, 123–124, 160, 248–249 162–163, 165; metadata 74, 85–90; Edgeworth, M. 335 mixed 119; old (legacy) 8, 15, 26, Elevation Model 174 29–31, 35–36, 76, 96, 98–102, 170, 315; elimination testing 15, 162 open access 106; partial 95; patterning embodiment 201, 340, 342 27, 97, 98, 161, 288–289, 315; quality empirical grounding 7, 43, 64. 94, 104; 140, 146; recording v, ix, 8, 11, 21, 24, research 11, 23, 31, 45, 55, 277, 288 39, 45, 59–61, 63–64, 66, 70, 72–74, 76, empiricist 159, 257, 274, 287 79–81, 83–85, 87–91, 113, 194, 198, Enlightenment 114, 325, 334 279; resolution-scale 141, 174, 184, 215, Environmental Impact Assessment 291 317; shared access 64, 74; stretching 97, epistemic: credibility 9, 342; genealogy 99; typological 92, 94–99, 105–106, 191; ideals 16; ladder (Hawkes) 313; 121, 222, 318; visualisation 9, 11, 74, optimism vs. pessimism 3, 6–8, 18, 256, 190–194, 196–202, 205–206 312–313; productivity 9, 15, 190, 192, database: access 64; archaeological packages 198, 200; status 7, 195, 313, 342; 79–82; computerised 65, 74, 79–83, value/virtues 10, 16, 196, 199 85–87, 117; design 74, 79–83, 85, 87, epistemological uncertainty 7, 173, 194, 89–91; functionality 74, 79–91; Intrasis 214, 239, 257 79; integrated 59, 74, 79–80, 85; epistemology 194–195, 199–200, 214, 239, publication 88–89, 105; relational 59, 290–300, 302, 304, 313–315, 337; 74; site-wide 74 negative 7, 313; Western 12, 36, 288, dataset x, 74–76, 89, 117, 130, 132, 290–291, 295, 299, 301–303, 304 135–137, 141, 148, 193, 201–202, 233, ethnoarchaeology 288, 342 248–252 ethnogenesis 297–298 dating, see chronology ethnographic: analogy 13, 244, 342; Debert, Nova Scotia 292 knowledge 303; record 42, 162–163, Decay Table 121 245; research 250, 287; sources 287, deductivism 6, 13, 257, 273, 277–278 301 Delgamuukw 295, 297, 299–300 ethnohistoric sources 288, 302 dendrochronology, see chronology Evans, C. 344 depositional episodes 92 evidence, archaeological vi, vii, 6–7, 18, Descartes, R. 325, 334 23–24, 51, 92, 94, 99, 104, 133, 140, dialectical process 46, 260, 274 161, 169, 175, 213, 236, 256, 264, diaries, digital, video 64, 72, 74, 198 266–267, 279, 287–288, 300–302, 312, 356 Index

320, 334; chemical 118, 168; contrary/ farming vi, 10, 13–14, 59, 243–252; non-congruent 16, 257, 295, 299; practice vi, 10, 14, 243–247; scenarios direct/indirect 14, 186, 244–245; 13, 243 historical 11, 98, 135, 281–283, 341; field system 28–29 lack of 60, 71, 74, 92, 102, 161, 314, fieldwork v, 8, 18, 21, 23–27, 29, 31, 60, 221; Law of 271–273, 276–277, 284, 69, 194, 321, 326–327, 330; strategies 297, 299; legitimate sources of 340, 344; 23–40, 60 lines of (independent, multiple, new) 3, find records 81, 90 9–10, 14, 16, 42, 45–46, 162, 168, 185, First Nations 46, 288, 397 214–218, 250–251, 255, 260, 265, 272, Flag Fen 50 297, 300–302; marshalling 256, flexibility 61, 169 259–260, 264, 266, 274, 283; material floodplain cultivation 243, 246, 249, 251 2, 4–6, 8, 10–12, 14, 17, 94, 98–99, flora 243, 246–248 268, 297, 324–325, 334–337, 340, fluidity 74, 105, 206 343–344, 346–348; ‘mixed masses of’ forensic archaeology 255 10, 259, 272–273; papyrological 98, formation processes 44, 47, 53, 55, 318 101; scientific 9, 15, 82, 159–169, Four Corners 297 166, 214–216, 220, 340, 348; see also fragment, fragmentary 7, 15, 46, 243–246, argument, ethnographic, 313–317, 321–322 experimental functional plant ecology 247 evidential paradox 93–94 Fussell’s Lodge long barrow xi, xii, evolutionary archaeology 161–162 233–235, 237 evolutionary theory 31, 40, 161 Evvia, Greece 248–251 Galicia xiii, 173, 176, 184 excavation ix, 2–3, 8, 15, 18, 23–27, Gell, A. 3, 18, 333 29–31, 33, 35–37, 39–40, 47–51, 53, Geller, M. 268, 271, 278–284 59–61, 63–72, 74–77, 79–85, 87–91, 97, generations 237 99–104, 106, 113–114, 121, 150, 163, geophysical 160 170, 176, 198, 200–201, 204, 215, 219, Gero, J. M. 14, 94, 97–99, 104 257, 259, 262–264, 263–264, 266, gestural knowledge 14, 195, 342 268–269, 270, 298, 311–312, 323, GIS (Geographical Information Systems) 326–328, 330–331, 333, 335, 344, 59, 117, 174–176, 186 346–348; open area 326 Gitksan 297 experience 25, 34–35, 65–66, 72, 76, 88, Glooscap 292, 299, 301 160, 170, 194 graphics 194 experiential 161, 174, 195–196, 200, 213–214, 217, 343, 347 Hambach Forest 248, 250 expertise 4–5, 9–10, 12, 16, 56, 65, 76–77, hammerstones xii, 328–329, 331–332 160, 190–192, 194, 97, 200–205, 275, Harris ‘Law of Superposition’ 264 346 Harris matrix ix, xii, 61–62, 174, experiment(s) v, ix, xi, 3, 8, 10–11, 13–14, 263–264 23–24, 27, 39, 40, 42–50, 52–53, 55–56, Hellenistic period 280 60, 194, 196, 198–199, 203, 248, heritage xii, 46, 187, 192, 291, 295, 299, 250–251, 267, 288, 331–333, 340–344, 303 347 hermeneutics 239, 315, 324 experimental archaeology v, 14, 42–43, historical ecology 4, 18 45–47, 49, 51, 53, 55, 74, 169; historiography 3, 15, 25, 343 Earthwork Project 47–48, 56 history, text-based 1, 2, 4, 18,132, 194, 347 expertise vi, 4–5, 9–10, 12, 15–17, 56, 65, history of science vii, 1, 11, 339–343, 76, 111, 160, 190, 192, 194, 196–197, 346–349 199, 205, 275, 344, 346 Hodder, I. 6–8, 11, 42–46, 56, 59–60, 63–68, 72, 75–76, 96, 134, 198–200, fact(s) vii, 3, 6, 44, 98, 259, 261–262, 274, 204, 215, 238, 313, 324, 334; 12-point 277, 284, 302, 315, 339, 346; material plan ix, 63–64, 75 277, 284; questions of 3, 6, 259 Holocene 130, 135–137, 246 Index 357

Hopewell 164–165 isotopes 9, 160, 165, 251; ‘fingerprint’ horizon index xi, 181–183 115–116; lead x, 116, 123–125, 170; horticulture 245 nitrogen 165; oxygen 160, 163, 165; Howick xi, 219–222 ratio analysis x, 124–125, 163; stable Hunting Shrine xi, 202–204 carbon 163–165, 251; strontium 160, husbandry, crop xv, 243, 245, 251 163, 165; values 123–124 hypothesis formation 277; testing 6, 14–15, 43–44, 95, 257, 277, 283, 301 Joint Information Systems Committee 80, hypothetico-deductive 257 83 Jordanova, L. 193, 195–197 Iamblichus argument 279, 281–282 Judd, N. 163 illustration, archaeological 15, 73, 189–190, juries 258–259, 262, 276, 283–284 193, 199 illustrator 9, 12, 72–73, 189, 190, 192, 196, Kant, I. 313, 325 199, 201, 204 Karanis x, 92–93, 97–106 Indigenous knowledge: see knowledge Kellis 102 Indigenous peoples 287–289, 292, 295, kernel density estimation 123 300, 303 Killackey, K. 201, 204–206 inductive reasoning 95, 243, 250, 257, 273, Kilmartin 327, 330–333 277, 315 knowledge: archaeological 88, 199, 217, inductively-coupled plasma mass 287; Indigenous 289–290, 295, 301, spectrometry (ICP–MS) 115, 167–168 303; local 287–289, 294, 300–303; inference 6, 12, 14–15, 52, 124, 243, systems 287, 290–291, 295, 299–302; 249–252, 255–257, 259–260, 262, traditional vii, 46, 287–291, 293–295, 264–267, 269, 272–273, 275, 282, 288, 300–303; Western 290–291, 299, 303 290, 302, 312; see also abductive, Kristiansen, K. 159 deductive, inductive reasoning Kuhnian paradigm 287 inference to the best explanation (ITBE) 15, 243, 249–252 Lakota 289, 300 information technologies 12, 79–91, Lamdin-Whymark, H. 331 173–188 landcape archaeology vi, 173, 175, 316 inorganic 115, 162 land-use planning (mapping, patterns, and institutionalization, of practice 189, 205 practices) 4, 5, 247, 288, 291, 294, instruments: medical 348; scientific 297–298, 301 340–341, 343–344, 346 ‘Last Wedge’ 276, 278–279, 281 intellectual property 301 law, English 258, 277 interaction 3, 44, 63–64, 91, 205, 214–215, Law of Evidence, see evidence 325–326, 335–336, 346 lead, see isotopes interdisciplinary 1, 8, 190, 374 legacy data, see data, old (legacy) Internet archaeology 88 legal analysis 260–264, 266–268, 271–276, interpretation: v, vii, 11, 42, 59, 63, 72, 74, 280–283; context 255–259, 262, 269, 105, 162–163, 202, 216, 257, 264, 300, 279, 284; standard of proof 257–258, 311–313, 315, 317, 322, 324; at the 275, 277; theory 10 ‘trowel’s edge’ 8, 64, 67, 72; Lejre, Denmark 47–48, 50, 55 collaborative 66, 72; static 123; Levant 139, 149 multivocal 45–46, 56, 59 Libby model x, 128–129, 131–133 intersection xi, 179–180, 200, 287–288, LiDAR data 174 290–291, 299, 301–302, 334, 346 lifetime 237 inter-subjectivity 213 Lipton, P. 243, 250–151 intra-act, intra-activity 326, 328, 330–333, lo xwi (Clam Gardens) xii, 292–193, 335–337 299–301, 304 Intrasis (database package) 79 logic, see abductive, deductive, inductive intra-site landscaping 65 reasoning; inference of proof 272–273 investigations: criminal 255, 257, 284; Longbridge Deverill 52 police 276–278, 284 luminescence dating 160, 214 358 Index maize agriculture 164–165 131–133, 140–151, 161, 166, 168, 170, mamoasx , xi, 173–186 173–174, 178, 186, 190, 192, 195, mariculture 292 200–203, 205, 217–227, 229–233, Markov Chain Monte Carlo 218 235–238, 243–251, 275, 289, 302, 315, Martinón-Torres, M. 169, 346–348 340, 343–344, 348; 3D xi, 201–203; Marxism 2, 95, 340–341 digital 200; scaffolding 173; strength of material/material culture i, vii, 1–2, 4–6, 146, 219; virtual reality 64, 192, 10, 17–18, 26–27, 44, 94, 133–134, 146, 199–200 148, 192, 252, 268, 287, 302, 309, 312, Modified Wigmorean Analysis vii, xii, 14, 322, 324, 339–341, 343–344, 346, 259, 261–264, 266–274, 276, 278–279, 348–349; analysis 4–5, 66–70, 333–335; 280, 283–284 culture, see material culture; dynamism Moel-y-Gaer ix, 49–54 201; traces i, 1–10, 29, 31, 42, 48–49, moral ethic 196 213, 313, 321–322, 324; see also Morgan, C. 190, 200–201, 206 evidence and assemblage Mount Gabriel mines 121 material culture studies 1, 5–6, 10, 18 movement: patterns, paths xi, 3, 119, 134, material science 4, 14, 126 173, 175, 177–181, 206, 219, 221 material agency 18, 333 multiple working hypotheses vi, 15–16, 46, materiality 1, 3, 6, 98, 125, 215, 237–238, 211, 301 273, 277, 284, 326, 333, 345–346, 349 multivocality, see interpretation measurement 43, 95, 113, 115, 128–133, Murrieta Flores, P. 178 135, 165–166, 214, 218, 223, 247–248, museums, open air 43, 55 325; repeated 129; scientific 166 media vi, 18, 89–90, 169, 189–191, 196, Nataawaau 296 199–200; visual vi, 189–191, 205 narratives 12, 60, 63, 70, 74, 113, 161–163, mediation 205–206, 317, 334 168–169, 193, 199, 221, 230, 233, 236, medical archaeology 348 238–239, 262, 264, 267, 273, 290, 300, Mediterranean 114, 129, 133–135, 321, 339 139–140, 143, 146, 148, 238, 244–245 Native science 290 megalithic architecture 31, 34 natural sciences vi, 109, 113, 159–160 Melbourn, Cambridgeshire 229–230 Navajo 297–299 Mellaart, J. 60, 135, 198, 202–204 negotiation 67, 177, 181, 197–198, metadata 74, 85–90 205–206, 302 metallurgy 122, 124–125, 169, 194, 347; Neolithic vi, xi, 10, 33, 35, 44, 59–60, 70, see also archaeometallury 128, 130, 173–174, 191, 198, 215, metalwork x, 27, 31, 116–117, 123–126, 230–233, 237, 243–244, 248–250, 269, 165, 225 327–330 methodology v, 2, 5–6, 14, 43–44, 60, neutrality 218 63–67, 69, 71–72, 74–76, 79, 81, 84, neutron activation analysis 115, 167 87–91, 101–102, 106, 115–117, 192, Ngatatjara 288 198–199, 215, 217–218, 225, 235, 248, Nlaka’pamux 295–296 251–252, 259, 289, 316–319, 322, 332; Nunamiut 288 data collection 67–72, 79; excavation 60–74, 76; recording v, 61, 66, 72–76, object(s) i, vii, 1–6, 10–11, 13, 18, 26–27, 79–81, 87–91; reflexive ix, 60, 63–67, 63, 68, 74, 86–87, 94, 101–102, 72, 74–76; sampling 59, 64–65, 69–71, 113–121, 123–125, 133, 138, 159, 117, 168, 314, 316 164, 170, 187, 195–197, 201, method-driven archeology 159 205–206, 214, 225, 291, 309, 311, 314, Mi’kmaw 290, 292 316, 318–321, 324–326, 333–334, micromorphology 48–49, 71–72 339–348 middle-range theory 12, 44, 55 object studies 1, 3, 5–6 Middle Woodland 164–165 objectivity 11, 16, 23, 24, 25, 27, 29, 31, models vi, x–xi, xiii, xv, 5, 7, 10–11, 33, 39, 45–48, 55, 63, 67, 76, 95, 96, 13–15, 24, 26–27, 34, 39, 47, 52, 61, 215, 267, 276–277, 279, 288, 302, 76, 103, 116, 121–122, 124–125, 315–316, 324, 325, 342–344 Index 359 observation 288, 290, 299, 301 precision 9, 13, 117, 121, 131, 135–136, ontological indeterminacy 214, 239 138–139, 141, 146, 150, 166–168, 176, ontology vii, 290, 300, 302, 320–321, 324 230, 236, 275, 283, 325 open source software 80 prehistory 4, 10, 24–25, 29, 31, 33–34, 36, optical petrography 160, 162, 167 40, 53, 95, 113, 128–133, 138, 151, oral history 292, 297–300, 303 162–163, 165–166, 170, 189, 198, tradition 10, 12, 288–292, 297 215–216, 237–238, 256, 298, 327, organic residue analysis 160, 163, 165 330–331 outlier x, 140, 143–144, 146–148, 219 pre-understandings, see background Overton Down Experimental Earthwork assumptions 52 principles of proof 262, 271; see also OxCal software 141, 143, 146–147, 151, evidence, Law of 218, 220, 224, 229, 239 priority tours ix, 64, 67–69 oxidation 117, 121 probability theory 217 probandum 258, 260–261, 264, 266, 268, palaeobotany 146, 165 272, 274, 276–280, 282, 284 papyrology 98, 101 probative force 273, 284 past, real 256, 302 processual archaeology, processualism 5–6, petrographic analysis 163, 168 16, 31, 134, 159, 162, 256, 268, 288, phenomenology 161, 200 312, 336 philosophy i, 17, 25–26, 42–45, 192, 195, proof 257–258, 262, 268, 271–273, 277, 257, 291, 325, 334 287, 292, 296; free 273; logic of philosophy of science 7, 162, 349, 371 272–273; principles of 262, 271; photography ix, 32, 37–39, 48–50, 54, 61, standard of 258 68, 73, 75, 80–82, 88, 98, 190, 192, provenance 115, 123, 168 197–199, 201, 262, 293, 295, 299, 312, publication 63, 96, 104, 263, 266 319, 327–329, 332, 345 Pueblo Bonito 163 phytosociology 246–48, 250 pueblos 162–163 Piggott, S. 24–26, 29, 31, 33–37, 40, 130, pyrotechnologies 160 191 pithouses xii, 298–299 quantitative methods 118, 217 Pitt-Rivers, A.H.L.F. 8, 11, 24–29, 31–32, quartz xii, 13, 33, 168, 184, 326, 328–329, 39, 113, 344 331–333 Plateau Pithouse Tradition 298 pluralism 8, 15–16, 46, 256; see also radiocarbon dating vi, x, 9, 33, 36, 40, interpretation, multivocal 128–129, 135, 140, 160, 164, 166–167, positivism, positivist 3, 16, 95, 345 170, 215–217, 219, 221–222, 225–227, post-excavation tools 48, 80–83, 85, 87, 229–231, 233, 235, 259, 269, 328, 329, 89, 326, 328, 331, 333 331; calibration curve x, 131–133, post-processual archaeology 31, 42, 256, 135–140, 147, 149, 230; chronology 268, 312, 336 134, 216, 236; dating methods 133, 138, practice i, v–vi, xi, 1–2, 4–16, 24–26, 29, 216, 327; SHCal13 x, 139; timescale 44, 59, 61, 63–64, 66–67, 70–72, 83–85, 130–131, 133–135, 140, 151 93–96, 99, 101, 104, 106, 116, 121, 123, radiocarbon revolutions: first, second, third 128, 150, 174, 189–196, 199, 201, 215, 10, 13, 128, 131–135, 140, 150 219, 225, 231–233, 237–238, 243–244, radiometric dating 166–167 268–269, 273, 277, 279, 283, 288–292, reasonable doubt 258, 276–277 302, 318–321, 336, 339, 341–344, recording v, ix, 8, 11, 21, 24, 39, 45, 346–348; archaeological 4, 7, 10, 11, 83, 59–61, 63–64, 66, 70, 72–74, 76, 79–81, 167, 170, 191, 199, 205, 257, 259, 264, 83–85, 87–91, 113, 194, 198, 279; 315, 364; artisanal 195; best/good 7, methodologies v, 61, 66, 79, 81, 87–91; 190, 206; data management 8; embodied sheets 61, 66, 72, 83–85, 87 341; feminist 125; guidelines for 7; recycling 116–118, 121, 123–124, 168 institutionalization 189, 205; pragmatics reflexivity 9, 16, 45–46, 59, 63–76, 202, of 16; working 8, 84 207, 227 360 Index relativism 6–7, 161–162, 213, 256–257, Scotland 24, 31, 35, 80, 231–232, 327 302 Scottish Urban Archaeological Trust 80 relevance 11, 146, 190, 193, 198, 248, 268, seasonality xv, 244–246, 248 272–273, 277–278, 284 Secwepemc, Cultural Education Society reliability 249, 275, 300, 341 xii, xv, 293–294, 298–299, 301 remote sensing 160 ‘seemaultaneity’ 213, 236 Renfrew, C. x, 128, 130–131, 133–135, selectivity 300–301 140–145, 262, 339, 349 sensitivity analysis 14, 233–236 replication xii, 43–47, 99, 194, 201, 299, seriation 214–215, 225–227, 229–230, 331–332, 342 236 representation(s) v, 11, 59, 173–175, 186, Settle, T. B. 342 191–195, 205, 342, 344; see also digital, Shepard, A. 162–163 visual, spatial Sibum, H. O. 4–5, 18, 195, 197, 342 research design 44, 47–48, 114 Sierra de Barbanza 173 resource management 291–292, 297 Sierra Morena Occidental 178 Reynolds, P. J. 44, 47–53, 56 Silchester Town Life Project 80, 82–84, ring ditch house 35–36 87–88, 91 robustness 192 simulation 9, 14, 173–175, 178–181, 184, rock art xii, 10, 288, 300, 326–332 186–187 Ross Island mine 119, 121 single context, see context roundhouse ix, 30, 45, 49–55 site seminars 75 Rutter Gap 143, 145 skill 9, 11, 23, 45, 60, 63, 65, 67, 73, 83, 88, 160, 164, 190–191, 193–197, samples 36, 61, 69–71, 82, 113, 123, 200–201, 205–206, 214, 274, 281, 290, 128–129, 131–132, 135, 137–139, 341–342 141–143, 146, 148, 164, 166–168, 170, Smith, G. (illustrator) 193 216, 218–219, 221–223, 225, 231, 233, social geography 252 236, 248–249, 252, 331, 348 social science 3–4, 320 sampling strategies 64, 69–71 sociology 2–3, 9, 341 scaffolding: 11–14, 99, 104, 173; see also Sociology of Scientific Knowledge (SSK) bootstrapping 341 Schiehallion ix, 34, 37–39 software interface 89–90 Schnapp, A. 324 Soknopaiou Nesos 99, 101, 106 science i, vi–vii, 1, 3–4, 7, 9, 11–12, 14–15, South Lodge Camp ix, 25–32, 36, 39 17, 42–46, 96, 113–114, 121, 125–126, Southwestern USA 162–163, 166–167 129–130, 134, 159–164, 166–169, 192, Spary, E. 346 194–196, 272, 275, 287–288, 290–291, spatial analysis 62–63, 81, 90, 96, 178, 186, 315, 320, 325, 339–344, 346–349; 215, 230, 233; associations 61–63, 81, archaeological applications vi, 12, 111; 95–96, 174, 214–216; representation 62, historical 44, 45; in archaeology 15, 174, 186 42–43, 113, 121, 125, 160, 162–163, Spaulding, A. 95 166–167, 169; studies i, 313, 333, 337; St Fagans ix, 49–53, 55–56 Western 291, 301–303 stable isotope analysis, see isotope analysis science wars 288 standpoint theory 10, 15–16, 97, 274, scientific (unscientific) vi, 2, 6–7, 9, 13, 276–277, 279–280, 282, 284, 313, 334 15–16, 24, 42–43, 45–46, 48, 56, 59, 82, statistical analysis 95, 115, 167, 225, 331; 95–97, 113–114, 125, 159–162, methods, in archaeology 218, 235 166–170, 191, 194–196, 203, 214–216, Stein River Valley xii, 295–296 220, 223, 259, 274, 288–289, 295, 298, stock enclosure 29 300–301, 324, 330, 339–348; stock of beliefs 262, 275 experiment 24, 46, 203, 342–343; stone circles ix, 25, 31, 33–36, 38–40 measurement 166; method vi, 42, 113, stratigraphic matrix ix, 79, 81–82, 88; 159, 167, 170, 274, 295; see also relationships 63, 66, 86, 263, 269; chronology, scientific; evidence, sequence 61, 65, 82, 218–219, 230, 233, scientific 236, 264 Index 361 stratigraphy ix, xii, 27, 61, 63, 65–67, triangulation vi, 3, 9, 45–46, 56, 211 70–71, 79–82, 85–86, 88, 90, 140, Trimithis x, 102–106 218–219, 230–231, 233, 236, 263–264, typological analysis 92, 94–97, 101, 106, 269 222–223; debate 95–96, 98, 106 Sumerian language 271, 276, 278 typology/typologies vi, 5, 8, 12, 14, survey(s) 3, 8, 29, 135, 169, 190, 194, 244, 92–99, 102, 104–106, 114, 116–119, 247–250, 292, 332–333 121, 125–126, 133, 141, 150, 176, Swogger, J. 189, 199, 206 214–215, 222–223, 225, 227, 236, 318; as pragmatic tool 96; domestic 92; taphonomy 92, 102–103, 105, 216–219, tyranny of vi, 92–96, 102, 105 223, 231, 236, 318 Taylor, W. 98 uniformitarianism 13, 43–45 technology 1, 3, 84, 123, 125, 131, 133, unobtrusive measures 3, 9 150, 196, 200, 274, 320, 331, 335 Tell Edfu 150 van der Merwe, N. 164–165 Tell el-Dab’a 150 van Minnen, P. 101 texts 2, 18, 97, 102, 113, 343, 347–348 Vázquez-Villoch, V. 173, 176–178, 180, textual record 4, 347 182, 184 Thayer, J. B. 272–273 video diaries, see digital, diaries theory 6–7, 10, 12, 23, 31, 40, 42–46, 52, virtual reality/research environments 64, 55, 63, 85, 95–96, 134, 161, 184, 217, 80, 83, 192, 199–200 259, 276, 280, 282, 284, 313–318, visual vi, xi, 64, 66, 70, 72, 97–98, 177, 339–340, 345; archaeological ii, xviii, 184, 186, 189–201, 203–206, 214, 216, xix, 20, 169, 302, 306, 323, 333, 338; 346; focus 97–98; intelligence 195; building 95–96; evolutionary, see media vi, 189–190, 205; representation evolutionary theory; formation 315–318; 191, 194, 205 legal, see legal theory; middle-range, visualisation(s): 9, 11, 14, 72, 74, 90, see middle-range theory; testing 44–45; 189–196, 196–202, 205–206; see also wars 7 data and digital visualization; media, Thera/Santorini volcano x, 114, 139, visual 145–147, 149–150 Vogel, J. 164–165 things i, ii, xi, xx, 1–4, 6, 10, 11, 17–20, Vulture Shrine 204 45, 58, 109, 114, 134, 140, 153, 193, 196, 205, 210, 214–215, 231, 232, 237, wall paintings 60, 101 240, 242, 307, 311, 312, 316–322, Wareham Experimental Earthwork 333–335, 337, 341, 345–346, 349, (Dorset) 53 350–351 weed assemblages 246, 249, 251 tillage 245, 248 West Stow farmhouse 55 timescale(s) 42–43, 47–48, 128, 130–131, Wet’suwet’en 297 133–135, 140, 151, 237–239 Wheeler, M. 26, 191 timescapes 13, 230, 236 Wigmore chart method, see Modified Torbhlaren xii, 328–332 Wigmorean Analysis trace elements 116–119, 121, 124 Wigmore, J. H. 10, 14, 16, 259–260, 262, traces, see material, traces 271 trading zone 8, 17 Wylie, A. 46, 162, 216–218 traditional knowledge, ecological 291, 301; see also knowledge, Indigenous xeroxing v, 8, 14, 42, 45, 50, 96 Traditional Land Use Studies 294 x-ray fluorescence 167 tree-ring data 129–131, 135–136, 138, 143, 166 Yak’s Quarter ix, 32 trial rules 271 York, A. 295–296, 299 trials 83–84, 255, 258, 277, 283–284 York Archaeological Trust 87–88, 90–91 eBooks from Taylor & Francis Helping you to choose the right eBooks for your Library

Add to your library's digital Choose from a range of subject collection today with packages or create your own! Taylor & Francis eBooks. Benefits for you We have over 50,000 eBooks ■ Free MARC records in the Humanities, Social ■ COUNTER-compliant usage statistics Sciences, Behavioural Sciences, ■ Flexible purchase and pricing options Built Environment and Law, ■ 70% approx of our eBooks are now from leading imprints, including DRM-free. Routledge, Focal Press and Benefits for your user Psychology Press. ■ Off-site, anytime access via Athens or referring URL ■ Print or copy pages or chapters Free Trials ■ Full content search Available ■ Bookmark, highlight and annotate text We offer free trials to qualifying academic, ■ Access to thousands of pages of corporate and government customers. quality research at the click of a button.

eCollections eFocus Choose from 20 different subject We have 16 cutting-edge interdisciplinary eCollections, including: collections, including:

Development Studies -t t 'A

Health Studies

I Law

^^Middl^as^tudie^^

For more information, pricing enquiries or to order a free trial, please contact your local sales team: UK/Rest of World: [email protected] USA/Canada/Latin America: [email protected] East/Southeast Asia: [email protected] India: [email protected] www.tandfebooks.com 3