<<

Archaeological Variability and Interpretation in Global Perspective

Archaeological Variability and Interpretation in Global Perspective

edited by Alan P. Sullivan III and Deborah I. Olszewski

University Press of Colorado Boulder © 2016 by University Press of Colorado

Published by University Press of Colorado 5589 Arapahoe Avenue, Suite 206C Boulder, Colorado 80303

All rights reserved Printed in the of America

The University Press of Colorado is a proud member of Association of American University Presses.

The University Press of Colorado is a cooperative publishing enterprise supported, in part, by Adams State University, Colorado State University, Fort Lewis College, Metropolitan State University of Denver, Regis University, University of Colorado, University of Northern Colorado, Utah State University, and Western State Colorado University.

∞ This paper meets the requirements of the ANSI/NISO Z39.48-1992 (Permanence of Paper).

ISBN: 978-1-60732-493-5 (cloth) ISBN: 978-1-60732-494-2 (ebook)

Library of Congress Cataloging-in-Publication Data

Names: Sullivan, Alan P., editor. | Olszewski, Deborah, editor. Title: Archaeological variability and interpretation in global perspective / edited by Alan P. Sullivan III and Deborah I. Olszewski. Description: Boulder : University Press of Colorado, [2016] | Includes bibliographical references. Identifiers: LCCN 2016000562| ISBN 9781607324935 (cloth) | ISBN 9781607324942 (ebook) Subjects: LCSH: Archaeology—Methodology. Classification: LCC CC75 .A6545 2016 | DDC 930.1—dc23 LC record available at http://lccn.loc.gov/2016000562 Front cover design by Irfan Ibrahim Contents

Acknowledgments ix

Chapter One Working with Archaeological Variability in the Twenty-First Century—Thinking about Materiality, Epistemology, and Ontology Alan P. Sullivan III and Deborah I. Olszewski 3

Section I. Advances in Interpreting Regional Archaeological Records

Chapter two A Lithic Perspective on Ecological Dynamics in the Upper of Western Eurasia C. Michael Barton and Julien Riel-Salvatore 25

Chapter Three The Significance of “Persistent Places” in Shaping Regional Settlement History: The Case of the Mimbres Mogollon Barbara J. Roth 53

Chapter Four Reductive Technology and the Epipaleolithic of the Middle East and North Africa Deborah I. Olszewski 71 Chapter Five Context and Complexity on the Arid Margins of : Assessing Human Reponses to an Unpredictable Environment Simon J. Holdaway, Justin I. Shiner, Patricia C. Fanning, and Matthew J. Douglass 99

Chapter six Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability for Mobility-Based Models of Technological Organization Alan P. Sullivan III 125

Section II. Venerable Sites Revisited

Chapter Seven Timelessness and the Legacy of Archaeological Cartography Sissel Schroeder and Lynne Goldstein 153

Chapter Eight Sherd Cross-Joins, Ceramic Use-Wear, and Depositional History: Rethinking the Sociopolitical Aftermath of a Collapsed Bronze Age Cistern at Myrtos-Pyrgos, Crete Emilia Oddo and Gerald Cadogan 175

Chapter Nine Estimating the Population Size of Casas Grandes: Empirical Issues and Theoretical Consequences David R. Wilcox 191

Chapter Ten Biface Production at Tabun: Manufacture, Maintenance, and Morphological Variability Gary O. Rollefson 213

vi contents Section III. Cross-Cultural, Conceptual, and Experimental Perspectives

Capth er eleven Celebrating the Dead and Recrafting Social Identity: Placing Prehistoric Mortuary Practices in Broader Social Context Brian F. Byrd and Jeffrey Rosenthal 233

Chapter Twelve Flint from the Ancestors: Ritualized Use of Stone Tools in the Prehistoric Southwest John C. Whittaker and Kathryn A. Kamp 267

C hAPTer Thirteen Form, Function, and Mental Templates in Paleolithic Archaeology Philip G. Chase 291

Chapter fourteen The Role of ControlledE xperiments in Understanding Variation in Flake Production Zeljko Rezek, Sam Lin, and Harold L. Dibble 307

List of Contributors 321 Index 325

contents vii

Acknowledgments

The editors are beholden to Jessica d’Arbonne, acquisi- tions editor, whose encouraging words and unflagging commitment to the volume enabled us to move for- ward in the face of the many obstacles that, at various times, seemed insurmountable. We are grateful to the C. P. Taft Research Center, University of Cincinnati, for awarding a publication grant to Alan P. Sullivan to cover a portion of the vol- ume’s indexing costs. Proceeds from this volume will be directed to the Native American Scholarships Fund of the Society for American Archaeology in the name of Arthur J. Jelinek, Emeritus Professor, School of Anthropology, University of .

Archaeological Variability and Interpretation in Global Perspective

1

One inclusive view of archaeology is that the field is Working with concerned with providing theoretically informed nar- Archaeological Variability in ratives of the cultural past that arise from unbiased the Twenty-First Century engagements with the archaeological record. To achieve this lofty objective, archaeologists routinely examine Thinking about Materiality, their assumptions about the interpretation of archaeo- Epistemology, and Ontology logical variability (e.g., Schroeder 2013), as well as ideas regarding the creation, organization, and analysis of problem-specific data (e.g., Jackson 2014). This wide- Alan P. Sullivan III and spread, and accelerating, practice of critical reflection Deborah I. Olszewski promotes disciplinary renewal, which in turn enables the development of robust methods and contributes to insights about how to conduct archaeological studies of human behavior and evolution in ways that are not constrained by disciplinary privilege (Lyman 2007) or political partiality (Leone and Potter 1992). But these are relatively recent developments (Fagan 2005) and contrast sharply with simplistic late-nine- teenth to mid-twentieth century conceptualizations of the emergence and interpretation of archaeologi- cal variability (Longacre 2010; Meltzer 1985). Looking back, this period of “innocence” (Clarke 1973), easily appreciated with a casual examination of Man the Tool- Maker (Oakley 1949), Ancient Man in North America (Wormington 1957), World Prehistory: An Outline (Clark 1961), or The Old Stone Age (Bordes 1968), evokes a time when accounts of human prehistory were largely uncontroversial and comfortably familiar. Everyone is aware, of course, that this state of affairs was upended DOI: 10.5876/9781607324942.c001

3 more than half a century ago when Lewis R. Binford (1962) observed that archaeologists conduct their investigations with an incomplete understanding of the archaeological record—its properties, sources of variability, and inferen- tial potential. Since then, archaeologists have labored, and continue to struggle, in hope of understanding the factors that influence the formation and con- tent diversity of the archaeological record (e.g., Barton and Riel-Salvatore 2014; Bar-Yosef et al. 2005; Jelinek 2013; Lucas 2012; March et al. 2014; Schiffer 1987; Shott 1998; Sullivan 2008; van der Veen 2007; Weiner 2010). Now largely unbound from its former conceptual constraints (Trigger 1991), archaeology today is populated by handfuls of theoretical approaches and interpretive par- adigms, all intended to enlighten investigations of the world’s extraordinarily diverse archaeological records (e.g., Bintliff and Pearce 2011; Hodder 2012; Preucel 2006; Rathje et al. 2013; Schiffer 2012; Wallace 2011). In fact, hardly a week goes by without the archaeological community receiv- ing word that a stunning new discovery has shattered what were considered settled matters in human prehistory and evolution, or that new methods now challenge archaeologists to rethink how best to study the remains of the cul- tural past. For instance, consider this sample of recent dispatches from the field: • Chronostratigraphic and artifactual evidence from has pushed the origins of the archaeological record to 3.3 mya (Harmand et al. 2015). • Geoarchaeological and paleohydrological data show a strong connection between the timing and magnitude of Mississippi River flood events and the rhythm of cultural dynamics at Cahokia (ad 600–1350), west-central Illinois, which is one the largest pre-Columbian settlements in North America (Munoz et al. 2015). • Micromorphological analysis of sediments combined with the distributional analysis of burned flints from Tabun , , indicate that mid-­ Pleistocene hominins learned to control fire and use it habitually far earlier than previously thought (Shimelmitz et al. 2014). • Correlation of distinctive growth patterns of wood recovered from seven Chaco Canyon Great Houses (northern New ) with those of har- vesting locales in distant (> 75 km) mountain ranges, revealed a previously unsuspected source, as well as a shift in the ranges that supplied construction timber for Chaco’s massive ancestral Puebloan structures (ca. ad 850–1140; Guiterman et al. 2016). These tightly controlled studies, among numerous others (see Harrison- Buck 2014), attest to the necessity of determining how the phenomena that archaeologists seek to understand arose and came to express the properties

4 Sullivan & Olszewski that are implicated in addressing different problems K( arkanas et al. 2015:1– 2). They illustrate, as well, the significance of a key attribute of twenty-first-­ century archaeological inquiry—the cultural past is “constantly being recreated” (Shanks 2007:591). The consequentiality of this idea is not that archaeologists are compulsive revisionists but that the consideration of new evidence, which arises commonly from new survey and excavation discoveries (e.g., Watson et al. 2015) and the application of advanced theoretically inspired methods (e.g., Caruana et al. 2014), invariably shifts our understandings of the nature of the cultural past, as the studies in this volume demonstrate. It is not surprising, therefore, that the field has sustained a recent surge in dialogues concerning the influence of the “ontological turn” in anthropology (e.g., Bessire and Bond 2014; Pedersen 2012; Swenson 2015), particularly with respect to discussions regard- ing the likelihood of alternative past cultural “worlds,” their discoverability, and “the problem of confirmation” (Alberti 2014; Jackson 2016). Nevertheless, these modern conversations dovetail seamlessly with David L. Clarke’s (1973) call for critical self-examination of archaeology’s epistemological foundations and illustrate the complex, shifting relationships between knowledge claims and what constitutes evidence in support of them (Wylie 2011). In continuing to explore the disciplinary consequences of these develop- ments, the studies in this volume employ a variety of theoretical approaches and assess their suitability for addressing persistent problems in the field. For instance, a well-established theoretical subfield in archaeology, analytic the- ory (Clarke 1968; Schiffer 1988), which even today is still broadly concerned with considerations of “typological revision” (Fowles 2011:898 [in Alberti et al. 2011]) and artifact classification (Zedeño 2009), is squarely aligned with inves- tigations of Paleolithic assemblage variability (chapters 4, 10, 13, and 14). In these respects, the authors of these studies are not only investigating ways to develop “impartial methodologies” (Shanks 2007:589), but are arguing as well that certain units of analysis are more advantageous than others for understand- ing aspects of assemblage variation that register the evolutionary significance of different artifact designs (e.g., Shea 2013). Similarly, aspects of ecological and evolutionary theory are entailed in sev- eral studies in this volume that focus on resolving how lithic assemblage vari- ability expresses regional-scale survival strategies when the objects of analysis are retouched artifacts (chapter 2), unretouched artifacts (chapter 5), or both retouched and unretouched artifacts (chapter 6). These diverse analyses are instructive because they reveal how different external theoretical frameworks— behavioral ecology (Codding and Bird 2015) in chapter 2, evolutionary theory (Cannon and Broughton 2010) in chapter 5, and niche-construction theory

Wn orki g with Archaeological Variability in the Twenty-First Century 5 (Smith 2011) in chapter 6—enable thoughtful assessments of the principle that variability among lithic assemblages is attributable to problem-solving strategies that humans develop in response to living among geographically heterogeneous and seasonally dynamic resource distributions (Holdaway and Douglass 2012:123). Moreover, the critical rethinking reflected in these chap- ters illustrates how the adoption of either an obligate (chapters 2 and 5) or facultative (chapter 6) ecological paradigm—that is, resource availability or potential productivity either is restricted to a fixed set of conditions (obligate) or it is not (facultative)—has profound consequences for inferring the effects of land-use on lithic artifact production, use, and discard. For the past decade, archaeologists worldwide have been exploring the degree to which reconsiderations of the relations between humans and their handiwork (material culture, material remains, “the material,” or things; e.g., Hodder 2012; Joyce 2012; Pearson 2004; Walker and Schiffer 2006; Wallace 2011) create opportunities to interrogate archaeological variability in ways that enable previously inaccessible or understudied aspects of the cultural past, and their connections to the modern world, to be revealed (e.g., Shanks 2012). For instance, artifacts that formerly had been marginalized or ignored in archaeological investigations, because of their low frequency or oddness, now have been reinterpreted in terms of theories of materiality (e.g., a “theory of bundling” [Pauketat 2013:35]; see also Zedeño 2009) as objects specifically designed to mediate uncertainty, risk, or danger (chapter 12). Also, aspects of what broadly can be considered agency theory (Barrett 2012; Varien and Potter 2008) are featured here in (1) a cross-cultural archaeological study that explores how social inequality and reproduction are registered in burial accompani- ments (chapter 11) and (2) several studies that investigate how regional politi- cal dynamics and “centers” come to be expressed archaeologically in architec- tural remains (chapters 3, 7, and 9) and ceramics (chapter 8).

Organization of the Volume Having situated the contributions to this volume theoretically, we now discuss how the investigation of problems involving classes of archaeologi- cal phenomena, using cases drawn from a range of biogeographical settings worldwide (figure 1.1), is facilitated when diverse perspectives and modes of inquiry are brought to bear on their resolution. For instance, the volume’s first section—Advances in Interpreting Regional Archaeological Records—con- sists of five chapters that explore how the analysis of spatial distributions of artifacts, assemblages, and sites at different spatial scales provides new insights

6 Sullivan & Olszewski Figure 1.1. Geographical distribution of the archaeological studies discussed in this volume. Numbers in circles refer to chapters.

regarding mobility strategies, interaction patterns, the organization of tech- nology, and the factors that promote a “sense of place.” This section begins with a study by C. Michael Barton and Julien Riel-Salvatore (chapter 2). According to the authors, the undeniable importance of stone tools to human survival makes variability among these artifacts a key factor in understanding the effects of climate change on hunter-gatherer adaptations (see also Barton and Riel-Salvatore 2014). In support of this proposition, their analysis of 167 assemblages from 31 localities in western Eurasia indicates a fundamental shift in land-use patterns that arose in response to environmental changes during the Upper Pleistocene, which ultimately bestowed a competitive advantage to and ensured the success of anatomically modern humans (AMH), in contrast to their contemporaneous neighbors, the Neanderthals. In chapter 3, Barbara J. Roth tackles an enduring problem in archaeology, understanding the emergence and socioeconomic consequences of “persistent places,” with a fresh analysis of the heavily studied Mimbres River valley of southwestern New Mexico. Earlier surveys there had documented a number of pithouse and pueblo sites, and many of them were clustered in particular locales on the landscape (e.g., Pool 2013). Subsequent excavations revealed long- term use of some of these areas, with many occupations dating to the Early Pithouse period (beginning ca. ad 200). Employing the concept of “persistent places,” Roth shows how protracted occupation of specific areas consequentially shaped Mimbres economic and social interactions (cf. Hegmon and Nelson 2007). Her study will be of interest to those archaeologists who are concerned

Wn orki g with Archaeological Variability in the Twenty-First Century 7 with understanding how the emergent archaeological record itself affects the trajectory of regional cultural change as people live on and react to a landscape that becomes increasingly crowded by large numbers of abandoned settlements. The interpretive utility of time-honored stone-tool typologies has come under intense scrutiny lately because of the expansion of alternative perspec- tives for interpreting the “meaning” of assemblage variability (e.g., Bisson 2000; Monnier 2006; Shea 2013; see also Shott 2008). Building on these exchanges, Deborah I. Olszewski (chapter 4) posits that variation among formal tool types that postdate the Middle Paleolithic is indicative of reduction sequences rather than the expression of different “cultures” (see also, e.g., Dibble 1995; Will et al. 2015). By its nature, reductive technology constricts available options in stone-tool production and retouching episodes, thereby trumping the effects of prevailing cultural differences. The merits of this reductive approach are illustrated with an analysis of Middle Eastern (Nebekian and Zarzian) and North African (Iberomaurusian) Epipaleolithic assemblages. As Olszewski demonstrates, because lithic reduction strategies frequently respond to and are contingent on local situations, we should expect to encounter multiple tech- nological convergences in the absence of cultural connections or continuity. Archaeologists working in other regions of the world and with material that originated during different time periods should appreciate immediately the significance of this study because it illustrates an approach for deconstraining interpretations that have been tied too closely to single referent (unicausal) explanations for assemblage variability (e.g., “cultural” or “mental template” differences [see chapter 13, this volume]). In chapter 5, Simon J. Holdaway, Justin I. Shiner, Patricia C. Fanning, and Matthew J. Douglass discuss the importance of considering context, raw material access, occupation duration, technology, and artifact reuse at the landscape scale of analysis (cf. Surovell 2009). With the results of a dozen years of research on surface scatters of stone artifacts in western New South Wales, Australia, the authors highlight the inferential significance of regional patterns of lithic assemblage variability using measures such as cortex ratio (Dibble et al. 2005; Douglass and Holdaway 2011) as well as the patterning provided by radiocarbon dates from heat-retainer hearths. They conclude that their “contextual analysis” approach enables a reconsideration of the nature of Aboriginal society at the margins of the Australian arid zone during the mid to late Holocene, that is, their mobile lifestyle emphasized the centrality of provisioning people rather than places. For many years, mobility-based models of technological organization have been used to explain changes in lithic artifact production and patterns

8 Sullivan & Olszewski of regional abandonment in the prehispanic American Southwest (e.g., Torres 2000). After reviewing the history of debitage analysis in Southwest archaeology, Alan P. Sullivan III (chapter 6) evaluates the Expedient Core and Adaptive Diversity hypotheses (Parry and Kelly 1987; Upham 1984) that employ such models. With debitage and tool assemblage data from five types of archaeological sites (dating between the eighth and sixteenth centuries ad) in the Upper Basin area of northern Arizona, the author concludes that, in contrast to the expectations of both hypotheses, significant bifacial tool manu- facture occurred during this period at both perennial home bases and contem- poraneous short-term, task-specific workplaces. This finding implies that the applicability of mobility-based models of technological organization, which were inspired by hunter-gatherer ethnoarchaeology or ethnography, may be more restricted than originally thought (Kelly 1992; McCall 2012). More gen- erally, his study illustrates that normative categories of human behavior (e.g., logistic foragers) and their alleged archaeological consequences (e.g., limited activity sites) are imperfectly connected, at best, particularly in cases where lithic technology was designed to acquire and process resources that material- ized in anthropogenic ecosystems (Haws 2012:72–73; Smith 2011). The second section of the volume—Venerable Sites Revisited—illustrates how new theoretical perspectives and methods promote the reinterpretation of important archaeological sites, such as Myrtos-Pyrgos (Crete), Aztalan (Wisconsin, USA), (Israel), Casa Grande (Arizona, USA), and Casas Grandes (Chihuahua, Mexico), that have figured prominently in accounts of New World and Old World prehistory for decades. In chapter 7, Sissel Schroeder and Lynne Goldstein explore the extent to which nineteenth-century surveyors’ maps of pre-European archaeological sites across eastern North America developed under the presumption that the region’s occupational history was short, which served to establish a time- less view of the past that persists today in many of the interpretations of the physical layout of mound sites in this region (Meltzer 1985). Mississippian mound sites (ca. eleventh to seventeenth centuries ad), in particular, continue to be rendered as though their attributes—platform mounds, palisades, one or more plazas, and structures—were all built, used, and abandoned nearly simul- taneously. Interpretations of site organization at the palisaded mound site of Aztalan serve as an example of the hegemony of flat or timeless archaeologi- cal cartography. Drawing on theoretical frameworks that feature time per- spectivism (Bailey 2007) and comparisons with other palisaded Mississippian sites across the Southeast, the authors offer an alternative to the timeless view of Aztalan’s site structure, one that focuses on understanding the complex

Wn orki g with Archaeological Variability in the Twenty-First Century 9 processes and archaeological consequences that are associated with the devel- opment of coalescent communities (see also chapter 11, this volume). Eschewing the investigation of features, such as heroic temples and iconic palaces (e.g., Galaty and Parkinson 2007) that (until recently) epitomized classical archaeology, Emilia Oddo and Gerald Cadogan (chapter 8) instead reassess the pottery assemblage that accumulated in Cistern 2 at the Bronze Age site of Myrtos-Pyrgos, Crete, after it collapsed and was repurposed as a “dump” during the Neopalatial period (1750–1450 bc). Pivotal to their analysis are considerations of the stratigraphic relations among ceramic cross-joins, as well as observations regarding wear on the sherds’ surfaces and breakage points (cf. Tenwolde 1992). Integration of both sets of observations reveals that, fol- lowing the cistern’s disuse as a water-holding facility, the ceramic assemblage accumulated in it during a single depositional cycle. In addition, the strati- graphic concentration of sherds with little wear suggests that they originated from household-debris clearing or related “feasting” activities that immediately predated the destruction of Myrtos-Pyrgos by fire. These new understandings provide a basis for the authors to assess the interpretive sufficiency of models of regional sociopolitical complexity on ancient Crete (e.g., Knappett 2009). In chapter 9, David R. Wilcox considers how the detailed analysis of post- occupational architectural characteristics of two large adobe sites in the south- ern American Southwest—Casa Grande (Coolidge, AZ) and Casas Grandes (or, Paquimé, Chihuahua, Mexico)—has consequences for estimating the populations of these late prehistoric centers, which have been implicated in models of Southwest prehistory for more than a century (Fowler and Cordell 2005). Here, the author focuses his discussion on a recent study by Whalen et al. (2010) that argues the population of Paquimé did not exceed about 2,500 people—widely regarded as a threshold for the emergence of social complex- ity (Kosse 1996). Drawing on his study of post-abandonment architectural disintegration and fill processes at Casa Grande, on observations made by early visitors (Obregón and Bartlett) to Casas Grandes, and on a detailed examination of how the main building at Paquimé deteriorated, the author infers a somewhat higher population estimate for the site, at least 3,000 peo- ple, and explores its theoretical and regional political implications. Next, Gary O. Rollefson (chapter 10) contrasts his earlier analysis of bifaces from the reexcavation of Tabun Cave (1967–1972), Israel—which produced one of the most precisely controlled collections of Lower and Middle Paleolithic artifacts in the Levant ( Jelinek 1982)—with that based on a new model of biface production. In his study, the author describes how experiments involv- ing biface and cleaver manufacture have shown that the technological features

10 Sullivan & Olszewski of these implements often are suppressed in interpretations that favor mor- phology (e.g., Quintero et al. 2007). This protocol results in a bias toward clas- sifying these artifacts as various types of bifaces rather than as cleavers, which are butchery or cutting implements. The implication of such “misclassifica- tion” is that the prevalence of cutting/butchering activities is underreported in assemblage interpretations. As it turns out, cleavers are far more common in the Tabun assemblage than the author’s original analysis indicated and, importantly, occur in much higher frequencies at a number of other sites in the Levant. Such typological confusion becomes problematic, moreover, for those researchers who use the frequencies of bifaces in evolutionary models that pos- tulate a correlation between the expansion of hominin cognitive abilities and technological differentiation (see Nowell 2010 for a review of these issues). The final section of the volume—Cross-Cultural, Conceptual, andE xperi­ mental Perspectives—includes four studies that examine the theoretical con- structs that archaeologists often use to interpret assemblage variability, and explore the possibility that archaeological phenomena, widely separated in time and space, share features that can be used to enhance inferences about the socioeconomic factors that influence mortuary practices, the meaning of “exotic” artifacts, and the causes of variation in artifact form. Drawing from two classic examples of complex hunter-gatherer societies (in the Near East and the San Francisco Bay area in northern California), Brian F. Byrd and Jeffrey Rosenthal (chapter 11) highlight how changes in socioeconomic strategies (e.g., resource intensification) were correlated with an elaboration of mortuary practices (Bandy and Fox 2010). As they show, age-grade-related mortuary practices functioned as stabilizing forces to inte- grate communities, which underscores two significant points. First, mortuary practices among transegalitarian groups can vary independently of political complexity. Second, because mortuary practices can change dramatically over relatively short periods of time, the search for broad cross-cultural trends requires diachronic investigations (e.g., Silverman and Small 2002), which robust archaeological studies, such as theirs, provide. In chapter 12, John C. Whittaker and Kathryn A. Kamp present an analysis of a class of stone tools, found occasionally at prehistoric sites in the American Southwest, comprising artifacts that are unusual in terms of form, material, context, and indications of use (cf. Mills 2004). The authors observe that, because these artifacts are ambiguous and relatively uncommon, the patterned behaviors they represent are rarely considered seriously in archaeological stud- ies. However, they present new evidence to support their claim that many of these atypical lithic items can be interpreted as powerful ritual objects (e.g.,

Wn orki g with Archaeological Variability in the Twenty-First Century 11 Brown and Walker 2008; cf. Alberti and Bray 2009:339–340), a suggestion bol- stered by ethnographic examples of stone tools being used as offerings, sym- bols of social status, protection against malevolent beings, lightning, and other dangers, as well as gifts from the ancestors. Philip G. Chase (chapter 13) discusses the epistemological usefulness of the “mental template” concept that archaeologists have routinely employed to interpret the degree to which assemblage variation reflects the evolution of human cognitive ability, symbolism, and language (cf. Wadley’s [2013] concept of “cognitive complexity”). As he shows, the term itself is hopelessly ambigu- ous and is defined differently by different analysts. For these and other reasons, Chase concludes that “mental template” should be dropped entirely from the literature as a descriptive and explanatory taxon. No longer constrained by the mental-template concept, the author argues that Paleolithic archaeologists, in particular, now can profitably turn their attention to developing testable behavioral models of interassemblage variability (e.g., Shipton et al. 2013). In chapter 14, Zeljko Rezek, Sam Lin, and Harold L. Dibble make the case for the role of highly controlled experiments in understanding how flakes form (see also Rezek et al. 2011). The authors propose that properly designed experi- ments enable the study of the effects of particular independent variables—such as the angle and force of the detaching blow—on flake form that are not appar- ent in artifact replication studies. In addition, the authors review some of the limitations of earlier controlled experiments and describe a new experimen- tal design that eliminates much of the artificiality inherent to this approach, thereby contributing to a synthetic model of assemblage formation dynamics.

Final Thoughts In closing, we would like to emphasize that each chapter in the volume endorses the proposition that, because archaeological research proceeds at dif- ferent spatial and temporal scales, and engages different theoretical frame- works and methodological protocols (e.g., Robb and Pauketat 2013), determi- nation of the origins and histories of archaeological phenomena is essential in evaluating their relevance for resolving significant problems in world archae- ology (Wylie 2008). With this perspective, orthodoxy is challenged, research- worthy controversies are defined, and strong inferences about the evolutionary pathways of humankind are thoughtfully developed and impartially evalu- ated (Bauer 2013). Mindful of the challenges, opportunities, and responsibili- ties that come with the investigation of the archaeological record, we think that archaeologists who aspire to learn about the ontological diversity of past

12 Sullivan & Olszewski cultural worlds ought to be encouraged by the possibility that their investiga- tions will be enriched by epistemological frameworks that, as the following chapters exemplify, focus on understanding how Earth’s archaeological phe- nomena came to be as they are today.

References Alberti, Benjamin. 2014. “Archaeology, Risk, and the Alter-Politics of Materiality.” Cultural Anthropology Online. Accessed May 31, 2015. http://culanth.org/field sights/476-archaeology-risk-and-the-alter-politics-of-materiality. Alberti, Benjamin, and Tamara L. Bray. 2009. “Animating Archaeology: Of Subjects, Objects, and Alternative Ontologies.” Cambridge Archaeological Journal 19(3):337 –343. http://dx.doi.org/10.1017/S0959774309000523. Alberti, Benjamin, Severin Fowles, Martin Holbraad, Yvonne Marshall, and Christopher Witmore. 2011. “‘Worlds Otherwise:’ Archaeology, Anthropology, and Ontological Difference.” Current Anthropology 52(6):896–912. http://dx.doi.org /10.1086/662027. Bailey, G. N. 2007. “Time Perspectives, Palimpsests, and the Archaeology of Time.” Journal of Anthropological Archaeology 26(2):198–223. http://dx.doi.org/10.1016/j.jaa .2006.08.002. Bandy, Matthew S., and Jake R. Fox. 2010. “Becoming Villagers: The Evolution of Early Village Societies.” In Becoming Villagers: Comparing Early Village Societies, ed. Matthew S. Bandy and Jake R. Fox, 1–16. Tucson: University of Arizona Press. Barrett, John C. 2012. “Agency: A Revisionist Account.” In Archaeological Theory Today, 2nd ed., ed. Ian Hodder, 146–166. Cambridge: Polity Press. Barton, C. Michael, and Julien Riel-Salvatore. 2014. “The Formation of Lithic Assemblages.” Journal of Archaeological Science 46:334–352. http://dx.doi.org/10.1016 /j.jas.2014.03.031. Bar-Yosef, Ofer, Anna Belfer-Cohen, Paul Goldberg, Steven L. Kuhn, Liliane Meignen, Bernard Vandermeersch, and Stephan Weiner. 2005. “Archaeological Background to Hayonim Cave and Meged Rockshelter.” In The Faunas of Hayonim Cave, Israel: A 200,000-Year Record of Paleolithic Diet, Demography, and Society, ed. Mary C. Stiner, 17–38. Cambridge, MA: Peabody Museum of Archaeology and Ethnology, Harvard University. Bauer, Alexander A. 2013. “Multivocality and ‘Wikiality:’ The Epistemology and Ethics of a Pragmatic Archaeology.” In Appropriating the Past: Philosophical Perspectives on the Practice of Archaeology, ed. Geoffrey Scarre and Robin Coningham, 176–194. Cambridge: Cambridge University Press.

Wn orki g with Archaeological Variability in the Twenty-First Century 13 Bessire, Lucas, and David Bond. 2014. “Ontological Anthropology and the Deferral of Critique.” American Ethnologist 41(3):440–456. http://dx.doi.org/10.1111/amet .12083. Binford, Lewis R. 1962. “Archaeology as Anthropology.” American Antiquity 28(2):217–226. http://dx.doi.org/10.2307/278380. Bintliff, John, and Mark Pearce. 2011. “Introduction.” In The Death of Archaeological Theory? ed. John Bintliff and Mark Pearce, 1–6. Oxford: Oxbow Books. Bisson, M. S. 2000. “Nineteenth Century Tools for Twenty-First Century Archaeology? Why the Middle Paleolithic Typology of François Bordes Must Be Replaced.” Journal of Archaeological Method and Theory 7(1):1–48. http://dx.doi.org /10.1023/A:1009578011590. Bordes, François. 1968. The Old Stone Age. London: World University Library. Brown, Linda A., and William H. Walker. 2008. “Prologue: Archaeology, Animism, and Non-Human Agents.” Journal of Archaeological Method and Theory 15(4):297 –299. http://dx.doi.org/10.1007/s10816-008-9056-6. Cannon, Michael D., and Jack M. Broughton. 2010. “Evolutionary Ecology and Archaeology: An Introduction.” In Evolutionary Ecology and Archaeology: Applications to Problems in Human Evolution and Prehistory, ed. Jack M. Broughton and Michael D. Cannon, 1–12. Salt Lake City: University of Utah Press. Caruana, Matthew V., Susana Carvalho, David R. Braun, Darya Presnyakova, Michael Haslam, Will Archer, Rene Bobe, and John W. K. Harris. 2014. “Quantifying Traces of Tool Use: A Novel Morphometric Analysis of Damage Patterns on Percussive Tools.” PLoS One 9(11):1–18. http://dx.doi.org/10.1371/jour nal.pone.0113856. Clark, Grahame. 1961. World Prehistory: An Outline. Cambridge: Cambridge University Press. Clarke, David L. 1968. Analytical Archaeology. London: Methuen. Clarke, David L. 1973. “Archaeology: The Loss of Innocence.” Antiquity 47(185):6–18. http://dx.doi.org/10.1017/S0003598X0003461X. Codding, Brian F., and Douglas W. Bird. 2015. “Behavioral Ecology and the Future of Archaeological Science.” Journal of Archaeological Science 56:9–20. http://dx.doi .org/10.1016/j.jas.2015.02.027. Dibble, Harold L. 1995. “Middle Paleolithic Scraper Reduction: Background, Clarification, and Review of the Evidence to Date.” Journal of Archaeological Method and Theory 2(4):299–368. http://dx.doi.org/10.1007/BF02229003. Dibble, Harold L., Utsav A. Schurmans, Radu P. Iovita, and Michael V. McLaughlin. 2005. “The Measurement and Interpretation of Cortex in Lithic Assemblages.” American Antiquity 70(3):545–560. http://dx.doi.org/10.2307/40035313.

14 Sullivan & Olszewski Douglass, Matthew J., and Simon J. Holdaway. 2011. “Quantifying Stone Raw Material Size Distributions: An Archaeological Application for Investigating Cortex Proportions and Stone Artefact Curation in Lithic Assemblages.” In Changing Perspectives in Australian Archaeology, Part IV, ed. Jim Specht and Robin Torrence. Technical Reports of the Australian Museum, Online, 23. Sydney: Australian Museum. Fagan, Brian M. 2005. A Brief History of Archaeology: Classical Times to the Twenty- First Century. Saddle River, NJ: Pearson/Prentice Hall. Fowler, Don D., and Linda S. Cordell. 2005. “Introduction.” In Southwest Archaeology in the Twentieth Century, ed. Linda S. Cordell and Don D. Fowler, 1–15. Salt Lake City: University of Utah Press. Galaty, Michael L., and William A. Parkinson, eds. 2007. Rethinking Mycenaean Palaces II. Monograph 60. University of California. Los Angeles: Cotsen Institute of Archaeology. Guiterman, Christopher H., Thomas W. Swetnam, and Jeffrey S. Dean. 2016. “Eleventh-Century Shift in Timber Procurement Areas for the Great Houses of Chaco Canyon.” Proceedings of the National Academy of Sciences of the United States of America 113 (5): 1186–90. http://dx.doi.org/10.1073/pnas.1514272112. Corrected at http://dx.doi.org/10.1073/pnas.1525340113 Harmand, Sonia, Jason E. Lewis, Craig S. Feibel, Christopher J. Lepre, Sandrine Prat, Arnaud Lenoble, Xavier Boës, Rhonda L. Quinn, Michel Brenet, Adrian Arroyo, et al. 2015. “3.3-Million-Year-Old Stone Tools from Lomekwi 3, West Turkana, Kenya.” Nature 521(7552):310–315. http://dx.doi.org/10.1038/nature14464. Harrison-Buck, Eleanor. 2014. “Anthropological Archaeology in 2013: The Search for Truth(s).” American Anthropologist 116(2):338–351. http://dx.doi.org/10.1111/aman .12101. Haws, Jonathan A. 2012. “Paleolithic Socionatural Relationships during MIS 3 and 2 in Central .” Quaternary International 264:61–77. http://dx.doi.org/10.1016 /j.quaint.2011.10.003. Hegmon, Michelle, and Margaret C. Nelson. 2007. “In Sync, but Barely in Touch: Relations between the Mimbres Region and the Hohokam Regional System.” In Hinterlands and Regional Dynamics in the Ancient Southwest, ed. Alan P. Sullivan III and James M. Bayman, 70–96. Tucson: University of Arizona Press. Hodder, Ian. 2012. Entangled: An Archaeology of the Relationships between Humans and Things. Malden, MA: Wiley-Blackwell. http://dx.doi.org/10.1002/9781118241912. Holdaway, Simon, and Matthew J. Douglass. 2012. “A Twenty-First Century Archaeology of Stone Artifacts.” Journal of Archaeological Method and Theory 19(1):101–131. http://dx.doi.org/10.1007/s10816-011-9103-6.

Wn orki g with Archaeological Variability in the Twenty-First Century 15 Jackson, Sarah E. 2014. “Domestication and Liberation: How We Relate to Our Data, and What It Means for Understanding the Maya.” Reviews in Anthropology 43:111–134. Jackson, Sarah E. 2016. “Envisioning Artifacts: a Classic Maya View of the Archaeological Record.” Journal of Archaeological Method and Theory 23 (forthcoming). Jelinek, Arthur J. 1982. “The Tabun Cave and Paleolithic Man in the Levant.” Science 216(4553):1369–1375. http://dx.doi.org/10.1126/science.216.4553.1369. Jelinek, Arthur J. 2013. Neanderthal Lithic Industries at . Tucson: University of Arizona Press. Joyce, Rosemary A. 2012. “Life with Things: Archaeology and Materiality.” In Archaeology and Anthropology: Past, Present, and Future, ed. David Shankland, 119–132. ASA Monographs 48. London: Berg. Karkanas, Panagiotis, Kyle S. Brown, Erich C. Fisher, Zenobia Jacobs, and Curtis W. Marean. 2015. “Interpreting Human Behavior from Depositional Rates and Combustion Features through the Study of Sedimentary Microfacies at Site 5-6, .” Journal of Human Evolution 85:1–21. http:// dx.doi.org/10.1016/j.jhevol.2015.04.006. Kelly, Robert L. 1992. “Mobility/Sedentism: Concepts, Archaeological Measures, and Effects.” Annual Review of Anthropology 21(1):43–66. http://dx.doi.org/10.1146 /annurev.an.21.100192.000355. Knappett, Carl. 2009. “Scaling Up: from Household to State in Bronze Age Crete.” In Inside the City in the Greek World: Studies of Urbanism from the Bronze Age to the Hellenistic Period, ed. Sara Owen and Laura Preston, 14–26. Oakville, CT: Oxbow Books. Kosse, Krisztina. 1996. “Middle Range Societies from a Scalar Perspective.” In Interpreting Southwestern Diversity: Underlying Principles and Overarching Patterns, ed. Paul R. Fish and J. Jefferson Reid, 87–96. Anthropological Research Papers No. 48. Tempe: Arizona State University. Leone, Mark P., and Parker B. Potter, Jr. 1992. “Legitimation and the Classification of Archaeological Sites.” American Antiquity 57(1):137–145. http://dx.doi.org/10.2307 /2694840. Longacre, William A. 2010. “Archaeology as Anthropology Revisited.” Journal of Archaeological Method and Theory 17(2):81–100. http://dx.doi.org/10.1007/s10816 -010-9080-1. Lucas, Gavin. 2012. Understanding the Archaeological Record. Cambridge: Cambridge University Press.

16 Sullivan & Olszewski Lyman, R. Lee. 2007. “Archaeology’s Quest for a Seat at the High Table of Anthropology.” Journal of Anthropological Archaeology 26(2):133–149. http://dx.doi .org/10.1016/j.jaa.2006.11.003. March, Ramiro Javier, Alexandre Lucquin, Delphine Joly, Juan Carlos Ferreri, and Mohamad Muhieddine. 2014. “Processes of Formation and Alteration of Archaeological Fire Structures: Complexity Viewed in the Light of Experimental Approaches.” Journal of Archaeological Method and Theory 21(1):1–45. http://dx.doi .org/10.1007/s10816-012-9134-7. McCall, Grant S. 2012. “Ethnoarchaeology and the Organization of Lithic Technology.” Journal of Archaeological Research 20(2):157–203. http://dx.doi.org/10 .1007/s10814-011-9056-z. Meltzer, David J. 1985. “North American Archaeology and Archaeologists, 1879–1934.” American Antiquity 50(2):249–260. http://dx.doi.org/10.2307/280483. Mills, Barbara J. 2004. “The Establishment and Defeat of Hierarchy: Inalienable Possessions and the History of Collective Prestige Structures in the Pueblo Southwest.” American Anthropologist 106 (2): 238–51. http://dx.doi.org/10.1525 /aa.2004.106.2.238. Monnier, G. F. 2006. “The Lower/Middle Paleolithic Periodization in Western Europe.” Current Anthropology 47(5):709–744. http://dx.doi.org/10.1086/506280. Munoz, Samuel E., Kristine E. Gruley, Ashtin Massie, David A. Fike, Sissel Schroeder, and John W. Williams. 2015. “Cahokia’s Emergence and Decline Coincided with Shifts of Flood Frequency on the Mississippi River.” Proceedings of the National Academy of Sciences of the United States of America 112(20):6319–6324. http://dx.doi.org/10.1073/pnas.1501904112. Nowell, April. 2010. “Defining Behavioral Modernity in the Context of Neandertal and Anatomically Modern Human Populations.” Annual Review of Anthropology 39(1):437–452. http://dx.doi.org/10.1146/annurev.anthro.012809.105113. Oakley, Kenneth P. 1949 [1964]. Man the Tool-Maker. Chicago: University of Chicago Press. Parry, William J., and Robert L. Kelly. 1987. “Expedient Core Technology and Sedentism.” In The Organization of Core Technology, ed. Jay K. Johnson and Carol A. Morrow, 285–305. Boulder: Westview Press. Pauketat, Timothy R. 2013. “Bundles of/in/as Time.” In Big Histories, Human Lives: Tackling Problems of Scale in Archaeology, ed. John Robb and Timothy R. Pauketat, 35–56. Santa Fe: School for Advanced Research Press. Pearson, Mike P. 2004. “Earth, Wood, and Fire: Materiality and Stonehenge.” In Soils, Stones, and Symbols, ed. Nicole Boivin and Mary Ann Owoc, 71–91. London: UCL Press.

Wn orki g with Archaeological Variability in the Twenty-First Century 17 Pedersen, Morten A. 2012. “Common Nonsense: A Review of Certain Recent Reviews of the ‘Ontological Turn.’” Anthropology of this Century 5. http://aotcpress .com/articles/common_nonsense/. Pool, Michael D. 2013. “Mimbres Mogollon Farming: Estimating Prehistoric Agricultural Production during the Classic Mimbres Period.” In Soils, Climate, and Society: Archaeological Investigations in Ancient America, ed. John D. Wingard and Sue E. Hayes, 85–107. Boulder: University Press of Colorado. Preucel, Robert W. 2006. Archaeological Semiotics. Malden, MA: Blackwell Publishing. http://dx.doi.org/10.1002/9780470754962. Quintero, L., P. Wilke, and G. Rollefson. 2007. “An Eastern Perspective on the Lower Paleolithic of the ‘Levantine Corridor.’.” Studies in the History and Archaeology of Jordan 9:157–166. Rathje, William L., Michael Shanks, and Christopher Witmore, eds. 2013. Archaeology in the Making: Conversations Through a Discipline. Oxon: Routledge. Rezek, Zeljko, Sam Lin, Radu Iovita, and Harold L. Dibble. 2011. “The Relative Effects of Core Surface Morphology on Flake Shape and Other Attributes.” Journal of Archaeological Science 38(6):1346–1359. http://dx.doi.org/10.1016/j.jas .2011.01.014. Robb, John, and Timothy R. Pauketat. 2013. “From Moments to Millennia: Theorizing Scale and Change in Human History.” In Big Histories, Human Lives: Tackling Problems of Scale in Archaeology, ed. John Robb and Timothy R. Pauketat, 3–33. Santa Fe: School for Advanced Research Press. Schiffer, Michael B. 1987. Formation Processes of the Archaeological Record. Albuquerque: University of New Mexico Press. Schiffer, Michael B. 1988. “The Structure of Archaeological Theory.” American Antiquity 5 (3):461–485. http://dx.doi.org/10.2307/281212. Schiffer, Michael B. 2012. Behavioral Archaeology: Principles and Practice. London: Equinox Publishing. Schroeder, Sissel. 2013. “How Can We Know? The Epistemological Foundation of Ecological Modeling in Archaeology.” In Soils, Climate, and Societies: Archaeological Investigations in Ancient America, ed. John D. Wingard and Sue E. Hayes, 205–223. Boulder: University Press of Colorado. Shanks, Michael. 2007. “Symmetrical Archaeology.” World Archaeology 39(4):589–596. http://dx.doi.org/10.1080/00438240701679676. Shanks, Michael. 2012. The Archaeological Imagination. Walnut Creek, CA: Left Coast Press. Shea, John J. 2013. “Lithic Modes A-I: A New Framework for Describing Global- Scale Variation in Stone Tool Technology Illustrated with Evidence from the East

18 Sullivan & Olszewski Mediterranean Levant.” Journal of Archaeological Method and Theory 20(1):151–186. http://dx.doi.org/10.1007/s10816-012-9128-5. Shimelmitz, Ron, Steven L. Kuhn, Arthur J. Jelinek, Avraham Ronen, Amy E. Clark, and Mina Weinstein-Evron. 2014. “‘Fire at Will:’ The Emergence of Habitual Fire Use 350,000 Years Ago.” Journal of Human Evolution 77:196–203. http://dx.doi.org /10.1016/j.jhevol.2014.07.005. Shipton, Ceri, Christopher Clarkson, Marco Antonio Bernal, Nicole Boivin, Clive Finlayson, Geraldine Finlayson, Darren Fa, Franciso Giles Pacheco, and Michael Petraglia. 2013. “Variation in Lithic Technological Strategies among the Neanderthals of .” PLoS One 8(6):1–13. http://dx.doi.org/10.1371/journal .pone.0065185. Shott, Michael. 1998. “Status and Role of Formation Theory in Contemporary Archaeological Practice.” Journal of Archaeological Research 6(4):299–329. http://dx .doi.org/10.1007/BF02446082. Shott, Michael. 2008. “Lower Paleolithic Industries, Time, and the Meaning of Assemblage Variation.” In Time in Archaeology: Time Perspectivism Revisited, ed. Simon Holdaway and LuAnn Wandsnider, 46–60. Salt Lake City: University of Utah Press. Silverman, Helaine, and David D. Small, eds. 2002. The Space and Place of Death. Archeological Papers No. 11. Arlington, VA: American Anthropological Association. Smith, Bruce D. 2011. “General Patterns of Niche Construction and the Management of ‘Wild’ Plant and Animal Resources by Small-Scale Pre-Industrial Societies.” Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences 366(1566):836–848. http://dx.doi.org/10.1098/rstb.2010.0253. Sullivan, Alan P., III. 2008. “An Archaeological View of the Archaeological Record.” In Archaeological Concepts for the Study of the Cultural Past, ed. Alan P. Sullivan, III, 7–23. Salt Lake City: University of Utah Press. Surovell, Todd A. 2009. Toward a Behavioral Ecology of Lithic Technology: Cases from Paleoindian Archaeology. Tucson: University of Arizona Press. Swenson, Edward. 2015. “The Materialities of Place Making in the Ancient Andes: a Critical Appraisal of the Ontological Turn in Archaeological Interpretation.” Journal of Archaeological Method and Theory 22 (3): 677–712. http://dx.doi.org/10 .1007/s10816-014-9202-2. Tenwolde, C. 1992. “Myrtos Revisited: The Role of Relative Function Ceramic Typologies in Bronze Age Settlement Analysis.” Oxford Journal of Archaeology 11(1):1–24. http://dx.doi.org/10.1111/j.1468-0092.1992.tb00253.x.

Wn orki g with Archaeological Variability in the Twenty-First Century 19 Torres, John A. 2000. “Changing Lithic Technology during the Basketmaker-Pueblo Transition: Evidence from the Anasazi Heartland.” In Foundations of Anasazi Culture: The Basketmaker-Pueblo Transition, ed. Paul F. Reed, 221–229. Salt Lake City: University of Utah Press. Trigger, Bruce G. 1991. “Distinguished Lecture in Archaeology: Constraint and Freedom—A New Synthesis for Archaeological Explanation.” American Anthropologist 93(3):551–659. http://dx.doi.org/10.1525/aa.1991.93.3.02a00010. Upham, Steadman. 1984. “Adaptive Diversity and Southwestern Abandonment.” Journal of Anthropological Research 40:235–256. van der Veen, Marijke. 2007. “Formation Processes of Desiccated and Carbonized Plant Remains—The Identification of Routine Practice.” Journal of Archaeological Science 34(6):968–990. http://dx.doi.org/10.1016/j.jas.2006.09.007. Varien, Mark D., and James M. Potter. 2008. “The Social Production of Communities: Structure, Agency, and Identity.” In The Social Construction of Communities: Agency, Structure, and Identity in the Prehispanic Southwest, ed. Mark D. Varien and James M. Potter, 1–18. Lanham, MD: AltaMira Press. Wadley, Lyn. 2013. “Recognizing Complex Cognition through Innovative Technology in Stone Age and Palaeolithic Sites.” Cambridge Archaeological Journal 23(02):163–183. http://dx.doi.org/10.1017/S0959774313000309. Walker, William, and Michael B. Schiffer. 2006. “The Materiality of Social Power: The Artifact-Acquisition Perspective.” Journal of Archaeological Method and Theory 13(2):67–88. http://dx.doi.org/10.1007/s10816-006-9002-4. Wallace, Sandra. 2011. Contradictions of Archaeological Theory: Engaging Critical Realism and Archaeological Theory. New York: Routledge. Watson, Adam S., Stephen Plog, Brendan J. Culleton, Patricia A. Gilman, Steven A. LeBlanc, Peter M. Whiteley, Santiago Claramunt, and Douglas J. Kennett. 2015. “Early Procurement of Scarlet Macaws and the Emergence of Social Complexity at Chaco Canyon, NM.” Proceedings of the National Academy of Sciences (Early Edition 1509825112). http://dx.doi.org/10.1073/pnas.1509825112. Weiner, Stephen. 2010. Microarchaeology: Beyond the Visible Archaeological Record. Cambridge: Cambridge University Press. http://dx.doi.org/10.1017/CBO9780 511811210. Whalen, Michael E., A. C. MacWilliams, and Todd Pitezel. 2010. “Reconsidering the Size and Structure of Casas Grandes, Chihuahua, Mexico.” American Antiquity 75(3):527–550. http://dx.doi.org/10.7183/0002-7316.75.3.527. Will, Manuel, Alex Mackay, and Natasha Phillips. 2015. “Implications of Nubian- Like Core Reduction Systems in Southern Africa for the Identification of Early

20 Sullivan & Olszewski Modern Human Dispersals.” PLoS One 10(6):e0131824. http://dx.doi.org/10.1371 /journal.pone.0131824. Wormington, H. M. 1957. Ancient Man in North America. Denver: Denver Museum of Natural History. Wylie, Alison. 2008. “Mapping Ignorance in Archaeology: The Advantages of Historical Hindsight.” In Agnotology (The Making and Unmaking of Ignorance), ed. Robert N. Proctor and Linda Schiebinger, 183–205. Stanford: Stanford University Press. Wylie, Alison. 2011. “Critical Distance: Stabilizing Evidential Claims in Archaeology.” In Evidence, Inference, and Enquiry, ed. Philip Dawid, William Twining, and Mimi Vasilaki, 371–394. Proceedings of the British Academy 171. Oxford: Oxford University Press. http://dx.doi.org/10.5871/bacad/9780197264843.003.0014. Zedeño, María Nieves. 2009. “Animating by Association: Index Objects and Relational Taxonomies.” Cambridge Archaeological Journal 19(3):407–417. http://dx .doi.org/10.1017/S0959774309000596.

Wn orki g with Archaeological Variability in the Twenty-First Century 21

Section I Advances in Interpreting Regional Archaeological Records

2

Neanderthals and their social and biological relation- A Lithic Perspective on ships to us have long been a subject of fascination span- Ecological Dynamics in ning the scientific and lay communities. The hominins the Upper Pleistocene generally classified as Neanderthal are found in western of Western Eurasia Eurasia, from the Near East to , and extend tem- porally from sometime in the late Middle Pleistocene through Oxygen Isotope Stage (OIS) 3 of the Upper C. Michael Barton and Pleistocene (Finlayson et al. 2006; Harvati 2007; Klein Julien Riel-Salvatore 2003). Within this temporal and geographic range, Neanderthals almost certainly lived as mobile hunter- gatherers, with small family-based social groups occu- pying territories of variable sizes and moving between residential camps, which they occupied for variable lengths of time in order to acquire, process, and con- sume the resources they needed to survive and repro- duce. Such a lifestyle leaves little in the way of material residues and most of what is left rapidly decomposes to invisibility from the standpoint of current archaeo- logical data recovery methods. Even the bones of ani- mals consumed and hominins who died are preserved only in exceptional circumstances. Stone artifacts are the only behavioral residues of Neanderthal behav- ior that remain in any abundance, scattered across the landscape—and these are largely discarded trash (Barton 1991; Frison 1968). There are probably more known Neanderthal sites than of any other premod- ern human. However, of the hundreds of thousands or millions of camps at which these Pleistocene hominins carried out their daily activities over the course of 100 DOI: 10.5876/9781607324942.c002

25 to 200 kya and across some 2 million km2, archaeologists know of only a few hundred (Klein 2003; Serangeli and Bolus 2008)—and most of these are in rockshelters and that probably were visited only rarely by these hunter- gatherers but that offer the exceptional conditions of preservation needed in order to find more than lithics (Laville, Rigaud, and Sackett 1980). Moreover, there is a general tendency for Paleolithic archaeologists to focus on one or a few deeply stratified sites, requiring many seasons of painstaking excavations. Such sites, while offering a long, diachronic view, provide but tiny windows on past human ecological behaviors, the kind of information that is needed to understand the drivers of human evolutionary dynamics. In this chapter, we summarize continental-scale data on human ecology and land-use that builds on over a decade of prior work identifying robust proxy measures of Upper Pleistocene human land-use and mobility strategies, supported both by archaeological analysis and agent-based models. We begin with an overview of the theoretical framework for using Paleolithic lithic assemblages to study past human ecology. Subsequently, we present a syn- thesis of human land-use and ecology, in response to environmental change during the Upper Pleistocene. We conclude with a discussion of implications of this biogeographical change for human evolution.

Middle-Range Lithic Theory While a number of complex classification schemes have been devised to characterize the diversity of lithic artifacts found in the archaeological record, in fact much of the macroscopic variability noted by archaeologists results from a few fundamental characteristics of workable stone and the tech- niques needed to make it useable: stone is heavy, stone was necessary for all Paleolithic hunter/gatherers, stone tools have short use lives, and stone tech- nology is messy.

Stone Is Heavy Most lithics are made from rocks that are composed primarily of silicon 3 dioxide (SiO2). With a density of 2.2 g/cm a 20-cm-diameter piece of flint weighs 9.2 kg. The equal volume of water weighs 4.2 kg; a small child weighs around 7 kg. Modern dietary recommendations suggest that 9.2 kg of pro- tein would meet the nutritional requirements of an average woman for 200 days and an average man for 164 days (Otten et al. 2006). In other words, for a mobile prehistoric forager on foot, stone for making tools had a much

26 B arTON & Riel-Salvatore higher energy cost to transport than many things of more direct importance to human survival. Given the high cost/return ratio of toolstone, foragers should have tried to minimize the amount of stone they needed to carry on a regular basis so that they could carry other items with higher and more direct energy payoffs.

Stone Was Necessary If stone had such an apparently low direct return relative to the cost needed to transport it, why then do we find it associated with all past humans so ubiq- uitously? In fact, until very recently, stone was a critical resource for human survival (Kuhn 1992; McCall 2012). People began to make recognizable flaked- stone artifacts over two—possibly three—million years ago and may have used stone for much longer (Harmand et al. 2015; cf. McPherron et al. 2010; Kimbel et al. 1996; Semaw et al. 1997); by at least a million years ago we do not find traces of humans without also finding stone artifacts. Stone tools had become an essential component of being human (cf. Nowell and Davidson 2010). They were the fundamental technology needed to process resources and to make other technological items upon which humans increasingly came to depend for their survival. This meant that, in spite of its high transportation costs, to be without usable stone could be as fatal as to be without food—and in fact could result in a lack of food. Humans had to stockpile usable stone at places they visited regularly and/or carry it with them on their increasingly lengthy foraging trips (Kuhn 1992). The constant need for useable stone along with its high transport costs put humans under strong and continuous selection pres- sure to devise ways to get increasing usability out of decreasing mass.

Stone Tools Are Short-Lived Many years ago, a student on a field project sported a t-shirt with the phrase “Love is fleeting, stone tools are forever.” While a catchy turn of phrase (for archaeologists at least), it is a misleading portrayal of stone tools. Stone far outlasts the most enduring affection, but stone tools commonly have a use-life much shorter than even the most casual flirtation. As observed in many rep- licative and ethnoarchaeological studies, stone tool edges dull rapidly, retouch is not particularly effective at rejuvenating a worn edge, and retouch rapidly reduces the size of the tool to the point that it becomes unusable (Andrefsky 2009; Barton 1990; Dibble 1987; Frison 1968; Gould et al. 1971; White and Thomas 1972). The short use-life of lithic artifacts is particularly problematic

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 27 given their critical role for prehistoric human ecology and the cost of ensuring that sufficient stone was available when needed. Again, this characteristic of stone artifacts must have kept human technological behaviors under strong selection pressure to devise ways to extend short use-life of available stone.

Lithic Technology Is Messy When archaeologists began to carry out quantitative replicative experiments of lithic technology, one result of flint-knapping was immediately apparent: making any artifact rapidly and easily produces a very large number of flakes, fragments, and pieces of debris, only some of which are useful for many tasks (e.g., Amick and Mauldin 1989; Magne and Pokotylo 1981; Shott 1994). Even assuming that most sharp-edged flakes of reasonable size potentially served as tools, creating even flakes produces a large amount of debris. Given the strong constraints on lithic artifacts described above, the messi- ness of lithic reduction technology meant that people whose activities required them to move their residence regularly would be under considerable selection pressure to carry with them only those pieces of stone with the most usable edges and least amount of waste, rather than carrying entire nodules and the heavy, unusable waste that they produce (Kelly 1983; Kuhn 1994; Surovell 2009). They would only discard these pieces of stone into the archaeological record (i.e., the trash) after trying to extend their use-lives as much as possible through rejuvenating the tool edges. On the other hand, people whose activi- ties allowed them to reside in a single locality for longer and accumulate a stockpile of stone, could break rocks with abandon, replacing short-lived tools with fresh ones instead of resharpening them, and amassing great quantities of lithic debris in the process (Parry and Kelly 1987). Because such temporarily sedentary groups did not have to worry as much about portability and did not have to pay the cost of transporting waste stone, lithic debris could accumulate rapidly at residential localities.

Implications for the Archaeological Record These inherent characteristics of flaked-stone technology have important implications for the archaeological record created from the trash generated by lithic production and use (Shott 1998, 2008). Human foragers who exhibit high residential mobility should be expected to transport and discard rela- tively few stone artifacts because stone is heavy but important for survival. In order to extend the short use-lives of the relatively few stone artifacts they

28 B arTON & Riel-Salvatore carry, lithics discarded by residentially mobile foragers should also exhibit reg- ular evidence of edge rejuvenation. That is, a comparatively higher proportion of the discarded artifacts should be retouched. On the other hand, logistically organized hunter-gatherers (i.e., those practicing central place foraging) have the opportunity to accumulate stone at regularly visited base camps, which they can occupy for longer stays. Lithic refuse at such sites should be charac- terized by larger quantities of lithic debris and fewer retouched artifacts. At most Paleolithic sites that are recognized as such by archaeologists—that is, recognized because of the quantity of ancient refuse (mostly stone artifacts and animal bones)—single occupations by residentially or logistically orga- nized foragers are not easily discernable. When they are, they are very difficult to interpret due to the low number of clearly associated artifacts (Aldeias et al. 2012). In fact, most assemblages that we study are probably time-averaged palimpsests of the trash of repeated human occupations of a locale (Barton et al. 2004; Barton and Clark 1993; Riel-Salvatore and Barton 2004). Hence, we would expect to see variable mixtures of residues from the different mobility strategies discussed above. The more that refuse from transiently sedentary base camps is represented, the greater the overall density of lithic debris that should be found, but the lower the relative frequency of retouched pieces in the assemblage. Conversely, the more that residentially mobile foragers con- tributed their lithic trash to an assemblage, the more it should be character- ized by a lower overall density of discarded lithic materials per volume of exca- vated sediment and by a higher frequency of retouched pieces. If the largely inherent properties of lithic technology and selection pressures on human technological behaviors described above hold true, we would expect to see a strong negative relationship between retouch frequency and total lithic arti- fact volumetric density in assemblages. In fact, in a series of papers published over the past decade, we have shown this to be the case repeatedly in assem- blages across southern Europe (Barton 1998; Barton and Riel-Salvatore 2012; Barton et al. 2011; Riel-Salvatore and Barton 2004; Riel-Salvatore and Barton 2007), and others have further replicated these results (Clark 2008, 2015; Kuhn 2004; Kuhn and Clark 2015; Sandgathe 2006). Furthermore, a series of recently published experiments using agent-based models have lent further support to these studies (Barton and Riel-Salvatore 2014); in fact, so far, only in excep- tional cases have these expectations not been met (e.g., Conard and Will 2015). These complementary studies and lines of evidence lend strong support to our hypothesis that this relationship between retouch frequency and artifact density (which we have elsewhere termed the “whole assemblage behavioral indicator,” or WABI) serves as robust proxy for human land-use strategies. In

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 29 addition, recent comparisons between the results of WABI values and detailed morphological analysis of individual artifacts further support these conclu- sions (Riel-Salvatore 2007, 2009). Building on this work, we use this proxy record of human land-use to inves- tigate large-scale human ecological dynamics during the Upper Pleistocene across southern Europe and the Near East. It is clear from our work and that of others that, because retouch frequency and artifact volumetric density are predicted to be negatively highly correlated—and indeed often are—we only need to use one of these two measures as a proxy for past land-use strategies. As noted in our prior work (Riel-Salvatore and Barton 2004), lithic volumetric density can vary with deposition processes across space and time, and so does not provide a comparable measure for synthesizing large-scale behavioral pat- terns beyond a particular depositional context, while retouch frequency is a normalized measure that is not affected by such depositional processes. Hence, we focus here on the frequency of retouched artifacts within entire lithic assemblages as a proxy for human land-use strategies.

Ecological Dynamics for Western Eurasia The syntheses presented here are based on data from 167 assemblages, recov- ered from 31 localities distributed geographically from Gibraltar to southern Jordan (figure 2.1, table 2.1). Temporally, they span the entire Upper Pleistocene from the last Interglacial through the end of the Pleistocene. In terms of stan- dard typological assignment, these assemblages include Middle Paleolithic, Upper Paleolithic, and what are sometimes called “transitional” (e.g., Uluzzian) industries. While not comprehensive of all sites known across western Eurasia, of course, this sample still represents a large spatial, temporal, and industrial range and is sufficient to display robust patterning discussed below. Our prior work suggests a general trend through the Upper Pleistocene toward reduced variation in land-use strategies, as indicated by retouch fre- quency and an accompanying trend toward an emphasis on logistical mobil- ity, and suggests further that variation in land-use strategies is responding to environmental change (Barton 1998; Barton and Riel-Salvatore 2012; Barton et al. 2011; Riel-Salvatore and Barton 2004, 2007; Riel-Salvatore et al. 2008). While reliable radiometric dates are available for only a few of the assemblages (unfortunately, the norm for most of the Upper Pleistocene archaeological record), we are able to assign all 167 assemblages to an Oxygen Isotope Stage (OIS; figure 2.2). At this temporal resolution, the pattern toward reduced vari- ability is clear, as is the trend from variable mobility strategies toward a focus

30 B arTON & Riel-Salvatore Figure 2.1. Locations of sites that produced assemblages used in analyses. Map shows Pleistocene coastlines (at –100 m bmsl) in gray, with modern coastlines as black line.

primarily on logistical mobility by the Last Glacial Maximum (LGM) (fig- ures 2.2 and 2.3). In a recent, detailed analysis of assemblages and paleoenvironmental data from OIS 3 in , Riel-Salvatore suggests that the amount of environmental variation is as important as general environmental trends as a driver of human land-use changes during the Upper Pleistocene (Riel-Salvatore 2007:87–92; Riel-Salvatore 2010). That is, increased unpredictability over much shorter time frames and consequent risk of resource shortfalls in the period leading up to the LGM created a fitness landscape strongly favoring flexible logisti- cal mobility strategies. These allowed humans to concurrently target a wider range of potential resources, over a much larger spatial extent, than could resi- dentially mobile strategies that cycled sequentially through more predictable, seasonally variable resources. Indeed, studies of recent hunter-gathers show that logistical strategies are more common at high latitudes, with high spa- tiotemporal variance in resource distribution and abundance, while residential mobility predominates at low latitudes (Binford 1980; Grove 2010; Kelly 1995). We use a high-resolution proxy for global climatic change—18O/16O ratios from the GISP2 (Greenland Ice Sheet Project) ice core (Meese et al. 1997)— to represent climate change at large scales. A simple comparison of retouch frequency for dated assemblages and δ18O (figure 2.4) shows some degree of correspondence, but it is not visually compelling. However, the combination of environmental variance with the maximum range of 18O/16O values within

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 31 Table 2.1. Assemblages used in analyses. Number of Country Site Industry Assemblages Gibraltar Gorham’s Cave Upper Paleolithic 4 Gibraltar Gorham’s Cave Middle Paleolithic 6 Italy Capelvenere Middle Paleolithic 11 Italy Castelcivita Proto-Aurignacian 2 Italy Castelcivita Uluzzian 4 Italy Marcello Zei Middle Paleolithic 3 Italy Mario Bernardini Uluzzian 3 Italy Mario Bernardini Middle Paleolithic 21 Italy Riparo Bombrini Proto-Aurignacian 2 Italy Riparo Bombrini Middle Paleolithic 8 Italy Serra Cicora Proto-Aurignacian 2 Italy Serra Cicora Uluzzian 2 Italy Serra Cicora Middle Paleolithic 3 Italy Torre dell’Alto Middle Paleolithic 6 Italy Uluzzo Epigravettian 3 Italy Uluzzo Uluzzian 2 Italy Uluzzo C Uluzzian 3 Italy Uluzzo C Middle Paleolithic 17 Jordan Ain al-Buhira Late Ahmarian 1 Jordan Ain al-Buhira Ahmarian 1 Jordan Tabaqa Late Epipaleolithic 1 Jordan Tor al-Tareeq Middle Epipaleolithic 1 Jordan Tor al-Tareeq Early Epipaleolithic 2 Jordan Tor Sadaf Early Ahmarian 1 Jordan Tor Sadaf Transition 2 Jordan Tor Sageer Early Epipaleolithic 1 Jordan WHS 621 Middle Paleolithic 1 Jordan WHS 634 Middle Paleolithic 1 Jordan Yutil al-Hasa Late Epipaleolithic 1 Jordan Yutil al-Hasa Early Epipaleolithic 1 continued on next page

32 B arTON & Riel-Salvatore Table 2.1—continued Number of Country Site Industry Assemblages Jordan Yutil-al-Hasa Late Ahmarian 2 Baia de Fier-Pestera Muierii Middle Paleolithic 1 Romania Borosteni-Pestera Cioarei Middle Paleolithic 4 Romania Cosava-Cuca Aurignacian 2 Romania Nandru-Pestera Curata Middle Paleolithic 5 Romania Nandru-Pestera Spurcata Middle Paleolithic 1 Romania Ohaba Ponor Aurignacian 1 Romania Ohaba Ponor Middle Paleolithic 4 Romania Pestera Moieciu-Pestera Mare Upper Paleolithic 2 Romania Pestera Moieciu-Pestera Mare Middle Paleolithic 1 Romania Pestera Valea Coacazei Middle Paleolithic 1 Romania Rasnov-Gura Cheii Middle Paleolithic 1 Romania Romanesti-Dumbravita Aurignacian 1 Spain Beneito Solutreo-Gravettian 1 Spain Beneito Gravettian 1 Spain Beneito Aurignacian 1 Spain Beneito Middle Paleolithic 1 Spain Cova del Salt Middle Paleolithic 6 Spain Cova Negra Middle Paleolithic 15 Total 31 sites 167

each OIS provides a measure of the extremes of environmental variation that hominins faced during the Upper Pleistocene in western Eurasia. In fact, these values are highly correlated with mean retouch frequency values for the 167 assemblages (figure 2.5). Moreover, there is a general tendency for the resource forays of logistically mobile hunter-gatherers to extend over a much greater geographical area than the territories occupied by residentially mobile foragers (Binford 2001; Grove 2009; Kelly 1983). That is, even though residentially mobile foragers may move more often during the course of a year, they tend to do so within an area of lim- ited spatial extent. On the other hand, logistically organized hunter-gather­ ers may travel less frequently but acquire resources over a region of much greater geographic extent on an annual and lifetime basis. The geographic range over

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 33 Figure 2.2. Retouch frequencies for all sites, by Oxygen Isotope Stage (N = 167 assemblages). Box plots show median and mid-spread. Dashed regression line shows overall trend in retouch frequencies over the Upper Pleistocene; grey shading shows 95% confidence intervals around the trend line.

which hunter-gatherers forage also affects the size of their potential mating network (sensu Wobst 1974; see Amick 1996 for an extreme case) with poten- tial consequences for biocultural evolution that are discussed below.

Discussion These results have interesting and potentially important implications for biobehavioral change during the Upper Pleistocene of western Eurasia. One way to interpret this evidence would be to say that Neanderthals were residen- tially mobile foragers, and were replaced by logistically mobile anatomically modern humans (AMH), whose flexible land-use strategies were favored by deteriorating climate approaching the LGM. However, while this story makes for an appealing narrative, it is overly simplistic and is not supported by the evidence presented here.

34 B arTON & Riel-Salvatore Figure 2.3. Least-squares regression of temporal trends in retouch frequency and variance in retouch frequency seen in figure 2.1. Figure 2.4. (Top) GISP2 δ18O and retouch frequency for all dated sites (N = 55 assemblages). (Bottom) Retouch frequency for dated assemblages and temporal variation in environments, as shown by coefficient of variation for 100-year rolling mean of GISP2 δ18O.

Indeed, during OIS 5–4 all the lithic assemblages in the assemblages stud- ied are classified as Middle Paleolithic and the only human remains in west- ern Europe, at least, are classified as Neanderthal. Similarly, during OIS 2, all the lithic assemblages studied are classified as Upper Paleolithic and the only human remains found are classified as AMH. But these observations tell us nothing about the nature of the changes in human behavior and biology that transpired during the Upper Pleistocene nor their causes. During OIS 3, when human remains classified as both Neanderthal and AMH are found, as well as lithic assemblages classified as Middle and Upper Paleolithic, there is still much debate over which “kind” of hominin made which “kind” of assemblage at this time (Hublin 2013). But, it is likely that most Middle Paleolithic and Transitional assemblages were made by hominins classified as Neanderthals and that most assemblages labeled Upper Paleolithic were made by hominins considered AMH (Riel-Salvatore and Clark 2007; Zilhão and d’Errico 2003). Looking at the points and graphs in figure 2.2, it is obvious that logis- tical strategies have been practiced since OIS 5–3 by hominins who made

36 B arTON & Riel-Salvatore Figure 2.5. Least-squares multiple regression of mean retouch frequency (x-axis) vs. combination of range plus variance in δ18O in the GISP2 for OIS 5–2 (y-axis). assemblages classified as Middle Paleolithic (cf. Lieberman and Shea 1994). What changed over the course of the Upper Pleistocene was the frequency of highly residentially mobile strategies. This decline is seen from OIS 3–2 among assemblages classed as Upper Paleolithic as well as in those classed as Middle Paleolithic during OIS 5–3. That is, the vectored changes in lithic industries that we have documented across western Eurasia in the Upper Pleistocene cross-cut the typological classifications of Middle and Upper Paleolithic. This is especially clear if we look at the distribution of retouch frequency sepa- rately in assemblages classed as Middle Paleolithic, Transitional, and Upper Paleolithic within OIS 3, when all are found (figure 2.6). By definition, we classify as Upper Paleolithic assemblages those that contain higher frequencies of artifacts that were probably hafted in compound tools, such as backed blades. Such compound tools, being lightweight and maintain- able (sensu Torrence 1989) should be more commonly associated with logis- tical land-use strategies. Conversely, we classify as Middle Paleolithic those

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 37 Figure 2.6. Retouch frequencies for OIS 3 assemblages from western Eurasia classified as Middle Paleolithic, Transitional, and Upper Paleolithic. Individual assemblages shown as black dots. Box-plots in gray show median, mid-spread, and range. Dashed line shows linear regression for industries; grey shaded zone shows 95% confidence interval.

assemblages dominated by flake artifacts that were less likely to be hafted and more likely to be ad hoc tools typical of residentially mobile foragers (Nelson 1991). Yet while the mean values for retouch frequency differ for the Upper and Middle Paleolithic assemblages shown in figure 2.6, there is complete overlap in the distributions of both groups. That is, during OIS 3, both Middle and Upper Paleolithic assemblages range from ones indicating high levels of logis- tical mobility to ones indicating moderate amounts of residential mobility. Returning to human biology, even if only Neanderthals were the makers of Middle Paleolithic and Transitional assemblages, and only AMH made Upper Paleolithic assemblages, the trends we discuss would cross-cut these biological differences as they do lithic classifications. If, as is increasingly sug- gested by paleogenetic data (e.g., Fu et al. 2015; Hawks and Throckmorton 2013; Hershkovitz et al. 2015), the relationships between human skeletal

38 B arTON & Riel-Salvatore morphology, human behavior, and the forms of discarded stone artifacts and production debris are more complex than this, the data we have presented offer a new window into the dynamics of human biogeography and behavior during this time of rapid environmental change (cf. Green et al. 2010). The increased prevalence of logistical mobility strategies across the span of the Upper Pleistocene, regardless of how assemblages are classified or who may have made and used them, means that all hominins were undertaking an increasing number of longer distance forays to collect resources to return to a home base. This would have brought members of all western Eurasian hom- inin groups in contact with a larger and more diverse number of other humans. This resulting increase in opportunities for social and biological interactions across much wider geographic regions would have had profound impacts on human biological variation. The very small sample of human fossils known even in western Eurasia, means that paleoanthropologists still have virtually no information about the nature of biological variation within real hominin populations (i.e., con- temporaneous individuals and social groups within delimited regions) in the Upper Pleistocene. But those individuals that have been found in western Eurasia especially (i.e., those classified as Neanderthals) seem to have devel- oped and maintained a suite of derived characteristics of skeletal morphology from sometime in the Middle Pleistocene into OIS 3 (Harvati 2007; Hublin 2009; Klein 2003; Wolpoff et al. 2004). The presence of a biologically dis- tinctive regional population in this region is not surprising, given its geogra- phy as a long, narrow peninsula, especially during periods of continental and Alpine ice sheets. This geography would have left western European hominins semi-isolated, allowing them to emerge as a regionally distinctive population (Serangeli and Bolus 2008). It would also be easier for western European hom- inins to maintain these distinctive characteristics when residential mobility within spatially delimited “territories” was the common land-use strategy, lim- iting the social and biological interaction of any group to its immediate neigh- bors. This scenario is congruent with Voisin’s characterization of Neanderthals and the AMH as ring species, with gene flow between adjacent populations but considerable genetic differences between geographically distant popula- tions (Voisin 2006; see also Holliday 2006). Greater long-distance mobility, resulting from increasing reliance on logistical land-use strategies, would have considerably reduced the limitations to gene flow and cultural transmission imposed by geography and residential mobility patterns in this region. In other animals (and plants), a biogeographical change that increases inter- actions among members of different variant populations or even sister species

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 39 (e.g., removal of a geographic barrier or transportation by humans of one taxon into the range of another) commonly results in increased hybridization rates, leading to a rapid disappearance of one of them as a recognizably distinct vari- ant or species (Epifanio and Nielsen 2001; Wolf 1990). Recent DNA sequenc- ing of a 38,000-year-old Neanderthal specimen, estimates that the last com- mon ancestor of modern humans and Neanderthals lived between 670,000 and 120,000 years ago (95% CI), with a mean at 370,000 (Noonan et al. 2006), and other estimates of the time of divergence fall within this range (Garrigan and Kingan 2007; Green et al. 2006; Krause et al. 2007; Harvati 2007). Mammalian rates of evolution of hybrid inviability are considerably faster than other ver- tebrates, but still average 2–4 million years (Fitzpatrick 2004), considerably longer than the time since the Neanderthal/modern human divergence. This means that Neanderthals and the ancestors of modern humans probably could have produced viable hybrids during OIS 3 (50,000–27,000 years ago). Recent sequencing of Neanderthal DNA and comparisons with modern (i.e., twenty- first century) human genomes indicates that such genetic exchange did take place between Neanderthals and AMH in western Eurasia (Green et al. 2010). However, even viable hybrids often exhibit decreased or increased fertility with respect to one or both of the ancestral populations that can have sig- nificant impacts on parent populations (especially if the parent populations have had the opportunity to diverge to some degree evolutionarily), regard- less of whether they are different species or different populations of the same species (Demuth and Wade 2007). Increased hybridization, coupled with decreased hybrid fertility will reduce the overall fertility of a parent popula- tion by siphoning off reproductive capacity into the production of low fertility (or even sterile) hybrids. On the other hand, increased hybridization accom- panied by the higher fertility of “hybrid vigor” can ultimately replace all of one parent population with hybrids. Such extinction through hybridization is sufficiently common in nature to be an important concern in conservation biology with significant impacts on rare and endangered species (Epifanio and Nielsen 2001; Wolf 1990). The fact that hybrids are often difficult to dis- tinguish from one of the ancestral species without genetic analysis (Brisbin 1995) exacerbates these problems. The increased biological and cultural inter- action between Neanderthal populations and other hominin variants, due to the changes in land-use patterns documented here, is sufficient to account for the disappearance of Neanderthal skeletal morphology (i.e., extinction) across western Eurasia through significantly increased hybridization rates without invoking interpopulation or interspecific competition or other mechanisms. But as Brisbin (1995) points out, obvious skeletal evidence for the existence

40 B arTON & Riel-Salvatore of recognizable hybrids may be difficult to come by in spite of new genetic evidence for such interactions (Fu et al. 2015). A model of extinction of Neanderthals through hybridization also helps explain otherwise puzzling features of their spatiotemporal distribution in OIS 3 (Smith et al. 2005). With direct competition for resources between human groups, the most effective competitors end up with the most productive lands, and the less effective competitors are relegated to marginally productive areas. This has regularly been the case when agriculturalists competed with hunter- gatherers, and even when immigrating hunter-gatherers encounter long- established hunter-gatherers in a region (Fiedel 2004). However, recognizable Neanderthal remains from OIS 3 are found across the southern and south- western fringes of Europe. Especially as global climate cooled, and became increasingly unpredictable and continental ice sheets and tundra spread across northern Europe, morphological Neanderthals occupied the most mesic areas of Europe with the most diverse and easily procured resources. The earliest specimens of AMH, in contrast, are found in the much more challenging (i.e., comparatively marginal) areas of central/eastern Europe and the steppe south of the ice (Finlayson and Carrión 2007; Finlayson et al. 2006; Nigst et al. 2014; Trinkaus 2005; van Andel et al. 2003; Zilhão 2007). At a high cost in organiza- tion and technology, these human groups managed to survive and (eventually) flourish in these rigorous environments in spite of their being especially chal- lenging ecologically for hunter-gatherers originating in Africa. But they did not apparently displace Neanderthals, who were supposedly much less effec- tive competitors, from glacial Europe’s prime real estate for millennia. However, if Neanderthals disappeared due to hybridization rather than being out-competed in some as yet unspecified way, the spatiotemporal distributions of Pleistocene hominins makes more sense. The newcomers to western Eurasia (i.e., AMH) were initially relegated to marginal environments of the more frigid north and east by the established and well-adapted local populations (i.e., Neanderthals). We have not compiled lithic data on Upper Paleolithic assemblages from eastern/central Europe (but cf. Riel-Salvatore et al. 2008). However it would not be surprising if an analysis like the one presented here shows that OIS 3 hominins in these area faced environmental pressures to shift their land-use strategies toward logistical mobility somewhat earlier than pop- ulations to the more mesic south. Furthermore, as immigrants into a relatively new (i.e., for Upper Pleistocene hominins) and expanding ecosystem that was also devoid of other human competitors, that of the mammoth steppe south of growing ice sheets, there also are ecological reasons to expect higher mobility and longer distance moves in these groups (Barton et al. 2004).

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 41 As increasingly unpredictable and extreme environmental oscillations during OIS 3 favored logistical land-use strategies more broadly, previously isolated Neanderthals in the European heartland came to interact biologi- cally and culturally to a much greater degree with other human groups to the east and north. These even more mobile fringe populations would have been transmitting genes and ideas over even longer distances and served as constantly replenishing reservoirs of a geographically more extensive “mod- ern” human gene pool. Once the process passed a critical threshold of genetic interchange, recognizably distinct Neanderthal populations could have disap- peared comparatively rapidly, falling victim to genetic “globalization” (Smith et al. 2005). Given the geography of Europe, we would expect this process to move through the population from the east to the most isolated western- most tip of the European peninsula. In fact, traditionally, the latest dates for morphologically recognizable Neanderthals have been argued to come from the Iberian Peninsula ( Jordá Pardo 2007; Zilhão 2006), although recent data have admittedly challenged this and proposed a relatively sudden shift across Europe (Higham et al. 2014). Because humans are cultural animals, with much behavioral information transmitted by nongenetic means, we would expect hybridization to likewise occur in behavioral residues, including lithic refuse, at an increasing rate from OIS 3 onwards. Given the expected difficulties in recognizing hybrid homi- nins, we should not be surprised to find Middle Paleolithic artifacts made by individuals who appear to be morphologically modern or Upper Paleolithic artifacts made by individuals who resemble Neanderthals. We might or might not find mosaics of traits in some specimens—if we were certain which traits were indeed ancestral Neanderthal and which were ancestral non-Neanderthal. In conclusion, we want to point out two other implications of the model of human ecological dynamics that we have presented here. The first is that, if Neanderthals disappeared due to hybridization, it means that at least some Neanderthal genetic material may survive in modern populations, but it is no longer clustered into a recognizable constellation of morphological traits. Recent comparisons between Neanderthal and modern genomes suggest as much as 4 percent or even more of western European genes may have Neanderthal origins (Green et al. 2010; Hawks and Throckmorton 2013; Wall et al. 2009). In this respect, it is interesting that modern Europeans and their descendants are quite distinct from the rest of humanity in their lack of pig- mentation, ever-growing dense facial hair in males and general hirsuteness, and a tendency toward mid-facial prognathism—with similar characteristics also distinguishing Neanderthals from other contemporaneous hominins

42 B arTON & Riel-Salvatore (Lalueza-Fox et al. 2007; Wolpoff et al. 2004). The second is that the model we have developed to explain robust, large-scale spatiotemporal patterning in Upper Pleistocene lithic assemblages across western Eurasia is based on evidence that suggests all hominins in this region, including Neanderthals, responded in a similar fashion to significant environmental change. If Neanderthals became extinct due to hybridization, it was because they effec- tively adapted to the environmental uncertainty of the Upper Pleistocene. That is, their disappearance was not due to being ineffective competitors, but rather a result of their success in adapting to the challenges of global climate change.

References Aldeias, Vera, Paul Goldberg, Dennis Sandgathe, Francesco Berna, Harold L. Dibble, Shannon P. McPherron, Alain Turq, and Zeljko Rezek. 2012. “Evidence for Neandertal Use of Fire at Roc de Marsal ().” Journal of Archaeological Science 39(7):2414–2423. http://dx.doi.org/10.1016/j.jas.2012.01.039. Amick, Daniel. 1996. “Regional Patterns of Folsom Mobility and Land-Use in the American Southwest.” World Archaeology 27(3):411–426. http://dx.doi.org/10.1080/0 0438243.1996.9980317. Amick, Daniel, and Raymond P. Mauldin, eds. 1989. Experiments in Lithic Technology. BAR International Series 528. Oxford: British Archaeological Reports. Andrefsky, William, Jr. 2009. “The Analysis of Stone Tool Procurement, Production, and Maintenance.” Journal of Archaeological Research 17(1):65–103. http://dx.doi.org /10.1007/s10814-008-9026-2. Barton, C. Michael. 1990. “Beyond Style and Function: A View from the Middle Paleolithic.” American Anthropologist 92(1):57–72. http://dx.doi.org/10.1525/aa.1990 .92.1.02a00040. Barton, C. Michael. 1991. “Retouched Tools: Fact or Fiction? Paradigms for Interpreting Chipped Stone.” In Perspectives on the Past: Theoretical Biases in Mediterranean Hunter-Gatherer Research, ed. Geoffrey A. Clark, 143–163. Philadelphia: University of Pennsylvania Press. Barton, C. Michael. 1998. “Looking Back from the World’s End: Paleolithic Settlement and Mobility at Gibraltar.” In Las Culturas del Pleistoceno Superior en Andalucía, ed. J. L. Sanchidrián Torti and M. D. Simón Vallejo, 13–23. Nerja, Spain: Patronato de la Cueva de Nerja. Barton, C. Michael, Joan Bernabeu, Aura, Oreto Garcia, Lluis Molina, and Steven Schmich. 2004. “Historical Contingency, Nonlinearity, and the Neolithization of the Western Mediterranean.” In Current Issues in Mediterranean Landscape Archaeology,

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 43 ed. LuAnn Wandsnider and Effie Athanassopoulos, 99–124. Philadelphia: University of Pennsylvania Press. http://dx.doi.org/10.9783/9781934536285.99. Barton, C. Michael, and Geoffrey A. Clark. 1993. “Cultural and Natural Formation Processes in Late Quaternary Cave and Rockshelter Sites of Western Europe and the Near East.” In Formation Processes in Archaeological Context, ed. Paul Goldberg, David T. Nash, and Michael D. Petraglia, 33–52. Madison, WI: Prehistory Press. Barton, C. Michael, and Julien Riel-Salvatore. 2012. “Agents of Change: Modeling Biocultural Evolution in Upper Pleistocene Western Eurasia.” Advances in Complex Systems 15(1–2):1150003-1–1150003-24. Barton, C. Michael, and Julien Riel-Salvatore. 2014. “The Formation of Lithic Assemblages.” Journal of Archaeological Science 46:334–352. http://dx.doi.org/10.1016 /j.jas.2014.03.031. Barton, C. Michael, Julien Riel-Salvatore, John M. Anderies, and Gabriel Popescu. 2011. “Modeling Human Ecodynamics and Biocultural Interactions in the Late Pleistocene of Western Eurasia.” Human Ecology 39 (6): 705–725. http://dx.doi.org /10.1007/s10745-011-9433-8. Binford, Lewis R. 1980. “Willow Smoke and Dogs’ Tails: Hunter-Gatherer Settlement Systems and Archaeological Site Formation.” American Antiquity 45(1):4–20. http://dx.doi.org/10.2307/279653. Binford, Lewis R. 2001. Constructing Frames of Reference: An Analytical Method for Archaeological Theory Building Using Hunter-Gatherer and Environmental Data Sets. Berkeley: University of California Press. Brisbin, I. L. 1995. “Conservation of the Wild Ancestors of Domestic Animals.” Conservation Biology 9(5):1327–1328. http://dx.doi.org/10.1046/j.1523-1739.1995 .09051327.x. Clark, Amy E. 2008. “Changes in Occupation Intensity During the Lower and Middle Paleolithic at Tabun Cave, Israel.” Master’s Thesis, University of Arizona, Tucson. Clark, Amy E. 2015. “Changes in Land Use and Occupation Intensity at the Onset of the Middle Paleolithic? A View from Tabun Cave, Israel.” In Settlement Dynamics of the Middle Paleolithic and Middle Stone Age, Volume IV, ed. Nicholas J. Conard and Anne Delagnes. Tübingen Publications in Prehistory. Tübingen: Kerns Verlag. Conard, Nicholas J., and Manuel Will. 2015. “Examining the Causes and Consequences of Short-Term Behavioral Change during the Middle Stone Age at Sibudu, South Africa.” PLoS One 10(6):e0130001. http://dx.doi.org/10.1371/journal .pone.0130001. Demuth, J. P., and Michael J. Wade. 2007. “Population Differentiation in the Beetle Tribollium castaneum. I. Genetic Architecture.” Evolution: International Journal of Organic Evolution 61(3):494–509. http://dx.doi.org/10.1111/j.1558-5646.2007.00048.x.

44 B arTON & Riel-Salvatore Dibble, Harold L. 1987. “The Interpretation of Middle Paleolithic Scraper Morphology.” American Antiquity 52(1):109–117. http://dx.doi.org/10.2307/281062. Epifanio, John, and Jennifer Nielsen. 2001. “The Role of Hybridization in the Distribution, Conservation and Management of Aquatic Species.” Reviews in Fish Biology and Fisheries 10(3):245–251. http://dx.doi.org/10.1023/A:1016729132297. Fiedel, Stuart J. 2004. “Rapid Migrations by Arctic Hunting Peoples: Clovis and Thule.” In The Settlement of the American Continents: A Multidisciplinary Approach to Human Biogeography, ed. C. Michael Barton, Geoffrey A. Clark, David R. Yesner, and Georges A. Pearson, 79–84. Tucson: University of Arizona Press. Finlayson, Clive, and José S. Carrión. 2007. “Rapid Ecological Turnover and Its Impact on Neanderthal and Other Human Populations.” Trends in Ecology & Evolution 22(4):213–222. http://dx.doi.org/10.1016/j.tree.2007.02.001. Finlayson, Clive, Francisco Giles Pacheco, Joaquin Rodriguez-Vidal, Darren A. Fa, José María Gutierrez López, Antonio Santiago Pérez, Geraldine Finlayson, Ethel Allue, Javier Baena Preysler, Isabel Cáceres, et al. 2006. “Late Survival of Neanderthals at the Southernmost Extreme of Europe.” Nature 443(7113):850–853. http://dx.doi.org/10.1038/nature05195. Fitzpatrick, Benjamin M. 2004. “Rates of Evolution of Hybrid Inviability in Birds and Mammals.” Evolution; International Journal of Organic Evolution 58(8):1865– 1870. http://dx.doi.org/10.1111/j.0014-3820.2004.tb00471.x. Frison, George C. 1968. “A Functional Analysis of Certain Chipped Stone Tools.” American Antiquity 33 (2): 149–155. http://dx.doi.org/10.2307/278516. Fu, Qiaomei, Mateja Hajdinjak, Oana Teodora Moldovan, Silviu Constantin, Swapan Mallick, Pontus Skoglund, Nick Patterson, Nadin Rohland, Iosif Lazaridis, Birgit Nickel, et al. 2015. “An Early Modern Human from Romania with a Recent Neanderthal Ancestor.” Nature ( June):22; advance online publication. Garrigan, Daniel, and Sarah B. Kingan. 2007. “Archaic Human Admixture: A View from the Genome.” Current Anthropology 48(6):895–902. http://dx.doi.org/10 .1086/523014. Gould, Richard A. K., D. A. Koster, and A. H. L. Sontz. 1971. “The Lithic Assemblage of the Western Desert Aborigines of Australia.” American Antiquity 36(2):149–168. http://dx.doi.org/10.2307/278668. Green, Richard E., Johannes Krause, Adrian W. Briggs, T. Maricic, U. Stenzel, M. Kircher, N. Patterson, H. Li, W. Zhai, M. H. Y. Fritz, et al. 2010. “A Draft Sequence of the Neandertal Genome.” Science 328(5979):710–722. http://dx.doi.org /10.1126/science.1188021. Green, Richard E., Johannes Krause, Susan E. Ptak, Adrian W. Briggs, Michael T. Ronan, Jan F. Simons, Lei Du, Michael Egholm, Jonathan M. Rothberg, Maja

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 45 Paunovic, et al. 2006. “Analysis of One Million Base Pairs of Neanderthal DNA.” Nature 444(7117):330–336. http://dx.doi.org/10.1038/nature05336. Grove, Matt. 2009. “Hunter-Gatherer Movement Patterns: Causes and Constraints.” Journal of Anthropological Archaeology 28(2):222–233. http://dx.doi.org/10.1016/j.jaa .2009.01.003. Grove, Matt. 2010. “Logistical Mobility Reduces Subsistence Risk in Hunting Economies.” Journal of Archaeological Science 37(8):1913–1921. http://dx.doi.org/10 .1016/j.jas.2010.02.017. Harmand, Sonia, Jason E. Lewis, Craig S. Feibel, Christopher J. Lepre, Sandrine Prat, Arnaud Lenoble, Xavier Boës, Rhonda L. Quinn, Michel Brenet, Adrian Arroyo, et al. 2015. “3.3-Million-Year-Old Stone Tools from Lomekwi 3, West Turkana, Kenya.” Nature 521 (7552): 310–15. http://dx.doi.org/10.1038/nature14464. Harvati, Katerina. 2007. “Neanderthals and Their Contemporaries.” In Handbook of Paleoanthropology, ed. Winifred Henke and Ian Tattersall, 1717–1748. http://dx.doi .org/10.1007/978-3-540-33761-4_56. Hawks, John, and Zach Throckmorton. 2013. “The Relevance of Archaic Genomes to Modern Human Origins.” In The Origins of Modern Humans: Biology Reconsidered, ed. Fred H. Smith and James C. Ahern, 339–354. Hoboken, New : Wiley- Blackwell. http://dx.doi.org/10.1002/9781118659991.ch10. Hershkovitz, Israel, Ofer Marder, Avner Ayalon, Miryam Bar-Matthews, Gal Yasur, Elisabetta Boaretto, Valentina Caracuta, Bridget Alex, Amos Frumkin, Mae Goder-Goldberger, et al. 2015. “Levantine Cranium from (Israel) Foreshadows the First European Modern Humans.” Nature 520(7546):216–219. http://dx.doi.org/10.1038/nature14134. Higham, Tom, Katerina Douka, Rachel Wood, Christopher Bronk Ramsey, Fiona Brock, Laura Basell, Marta Camps, Alvaro Arrizabalaga, Javier Baena, Cecillio Barroso-Ruíz, et al. 2014. “The Timing and Spatiotemporal Patterning of Neanderthal Disappearance.” Nature 512(7514):306–309. http://dx.doi.org/10.1038 /nature13621. Holliday, T. 2006. “Neanderthals and Modern Humans: An Example of a Mammalian Syngameon?” In Neanderthals Revisited: New Approaches and Perspectives, ed. Katerina Harvati and Terry Harrison, 281–297. Dordrecht, Netherlands: Springer. http://dx.doi.org/10.1007/978-1-4020-5121-0_16. Hublin, Jean-Jacques. 2009. “The Origin of Neandertals.” Proceedings of the National Academy of Sciences of the United States of America 106(38):16022–16027. http://dx.doi .org/10.1073/pnas.0904119106. Hublin, Jean-Jacques. 2013. “The Makers of the Early Upper Paleolithic in Western Eurasia.” In Origin of Modern Humans: Biology Reconsidered, ed. Fred H. Smith

46 B arTON & Riel-Salvatore and James C. Ahern, 223–252. Hoboken, New Jersey: Wiley-Blackwell. http://dx .doi.org/10.1002/9781118659991.ch6. Jordá Pardo, Jesús F. 2007. “The Wild River and the Last Neanderthals: A Palaeoflood in the Geoarchaeological Record of the Jarama Canyon (Central Range, Guadalajara Province, Spain).” Geodinamica Acta 20(4):209–217. http://dx .doi.org/10.3166/ga.20.209-217. Kelly, Robert L. 1983. “Hunter-Gatherer Mobility Strategies.” Journal of Anthropological Research 39:277–306. Kelly, Robert L. 1995. The Foraging Spectrum: Diversity in Hunter-Gatherer Lifeways. Washington, DC: Smithsonian Institution Press. Kimbel, W. H., R. C. Walter, D. C. Johanson, K. E. Reed, J. L. Aronson, Z. Assefa, C. W. Marean, G. G. Eck, R. Bobe, E. Hovers, et al. 1996. “Late Pliocene Homo and Oldowan Tools from the Hadar Formation (Kada Hadar Member), Ethiopia.” Journal of Human Evolution 31(6):549–561. http://dx.doi.org/10.1006/jhev.1996.0079. Klein, Richard G. 2003. “Paleoanthropology: Whither the Neanderthals?” Science 299(5612):1525–1527. http://dx.doi.org/10.1126/science.1082025. Krause, Johannes, Carles Lalueza-Fox, Ludovic Orlando Wolfgang Enard, Richard E. Green, Hernán A. Burbano, Jean-Jacques Hublin, Catherine Hänni, Javier Fortea, Marco de la Rasilla, et al. 2007. “The Derived FOXP2 Variant of Modern Humans Was Shared with Neandertals.” Current Biology 17(21):1908–1912. http://dx.doi.org/10.1016/j.cub.2007.10.008. Kuhn, Steven L. 1992. “On Planning and Curated Technologies in the Middle Paleolithic.” Journal of Anthropological Research 48:185–214. Kuhn, Steven L. 1994. “A Formal Approach to the Design and Assembly of Mobile Toolkits.” American Antiquity 59(3):426–442. http://dx.doi.org/10.2307/282456. Kuhn, Steven L. 2004. “Middle Paleolithic Assemblage Formation at Riparo Mochi.” In Processual Archaeology: Exploring Analytical Strategies, Frames of Reference and Culture Process, ed. Amber Lynn Johnson, 31–60. Westport, CT: Greenwood Publishing Group. Kuhn, Steven L., and Amy E. Clark. 2015. “Artifact Densities and Assemblage Formation: Evidence from Tabun Cave.” Journal of Anthropological Archaeology 38:8–16. http://dx.doi.org/10.1016/j.jaa.2014.09.002. Lalueza-Fox, Carles, Holger Rompler, David Caramelli, C. Staubert, G. Catalano, D. Hughes, N. Rohland, E. Pilli, L. Longo, S. Condemi, et al. 2007. “A Melano­ cortin 1 Receptor Allele Suggests Varying Pigmentation Among Neanderthals.” Science 318(5855):1453–1455. http://dx.doi.org/10.1126/science.1147417. Laville, Henri, Jean-Philippe Rigaud, and James Sackett. 1980. Rock Shelters of the Perigord. New York: Academic Press.

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 47 Lieberman, Daniel E., and John J. Shea. 1994. “Behavioral Differences between Archaic and Modern Humans in the Levantine Mousterian.” American Anthropologist 96 (2): 300–332. http://dx.doi.org/10.1525/aa.1994.96.2.02a00030. Magne, Martin P., and David Pokotylo. 1981. “A Pilot Study in Bifacial Lithic Reduction Sequences.” Lithic Technology 10:34–47. McCall, Grant S. 2012. “Ethnoarchaeology and the Organization of Lithic Technology.” Journal of Archaeological Research 20(2):157–203. http://dx.doi.org/10 .1007/s10814-011-9056-z. McPherron, Shannon P., Zeresenay Alemseged, Curtis W. Marean, Jonathan G. Wynn, Denné Reed, Denis Geraads, René Bobe, and Hamdallah A. Béarat. 2010. “Evidence for Stone-Tool-Assisted Consumption of Animal Tissues before 3.39 Million Years Ago at Dikika, Ethiopia.” Nature 466(7308):857–860. http://dx.doi .org/10.1038/nature09248. Meese, D. A., A. J. Gow, R. B. Alley, G. A. Zielinski, P. M. Grootes, M. Ram, K. C. Taylor, P. A. Mayewski, and J. F. Bolzan. 1997. “The Greenland Ice Sheet Project 2 Depth-Age Scale: Methods and Results.” Journal of Geophysical Research 102(C12):26411–26423. http://dx.doi.org/10.1029/97JC00269. Nelson, Margaret C. 1991. “The Study of Technological Organization.” In Archaeological Method and Theory, vol. 3. ed. Michael B. Schiffer, 57–100. Tucson: University of Arizona Press. Nigst, Philip R., Paul Haesaerts, Freddy Damblon, Christa Frank-Fellner, Carolina Mallol, Bence Viola, Michael Götzinger, Laura Niven, Gerhard Trnka, and Jean- Jacques Hublin. 2014. “Early Modern Human Settlement of Europe North of the Alps Occurred 43,500 Years Ago in a Cold Steppe-Type Environment.” Proceedings of the National Academy of Sciences of the United States of America 111(40):14394– 14399. http://dx.doi.org/10.1073/pnas.1412201111. Noonan, James P., Graham Coop, Sridhar Kudaravalli, D. Smith, J. Krause, J. Alessi, F. Chen, D. Platt, S. Paabo, J. K. Pritchard, et al. 2006. “Sequencing and Analysis of Neanderthal Genomic DNA.” Science 314(5802):1113–1118. http://dx.doi.org/10 .1126/science.1131412. Nowell, April, and Iain Davidson. 2010. Stone Tools and the Evolution of Human Cognition. Boulder: University Press of Colorado. Otten, Jennifer J., Jennifer Pitzi Hellwig, and Linda D. Meyers, eds. 2006. Dietary Reference Intakes: The Essential Guide to Nutrient Requirements. Washington, DC: National Academies Press. Parry, William J., and Robert L. Kelly. 1987. “Expedient Core Technology and Sedentism.” In The Organization of Core Technology, ed. Jay K. Johnson and Carol A. Marrow, 284–304. Boulder: Westview Press.

48 B arTON & Riel-Salvatore Riel-Salvatore, Julien. 2007. “The Uluzzian and the Middle-Upper Paleolithic Transition in Southern Italy.” PhD dissertation, Arizona State University, Tempe. Riel-Salvatore, Julien. 2009. “What Is a “Transitional” Industry? The Uluzzian of Southern Italy as a Case Study.” In Sourcebook of Paleolithic Transitions: Methods, Theories, and Interpretations, ed. Marta Camps and Parth Chauhan, 377–396. New York: Springer. http://dx.doi.org/10.1007/978-0-387-76487-0_25. Riel-Salvatore, Julien. 2010. “A Niche Construction Perspective on the Middle– Upper Paleolithic Transition in Italy.” Journal of Archaeological Method and Theory 17(4):323–355. http://dx.doi.org/10.1007/s10816-010-9093-9. Riel-Salvatore, Julien, and C. Michael Barton. 2004. “Late Pleistocene Technology, Economic Behavior, and Land-Use Dynamics in Southern Italy.” American Antiquity 69(2):257–274. http://dx.doi.org/10.2307/4128419. Riel-Salvatore, Julien, and C. Michael Barton. 2007. “New Quantitative Perspectives on the Middle-Upper Paleolithic Transition: The View from the Northern Mediterranean.” In Early Upper Paleolithic “Transitional” Industries: New Questions, New Methods, ed. Julien Riel-Salvatore and Geoffrey A. Clark. BAR International Series 1620. Oxford: British Archaeological Reports. Riel-Salvatore, Julien, and Geoffrey A. Clark, eds. 2007. New Approaches to the Study of Early Upper Paleolithic “Transitional” Industries in Western Eurasia: Transitions Great and Small. BAR International Series S1620. Oxford: British Archaeological Reports. Riel-Salvatore, Julien, Gabriel Popescu, and C. Michael Barton. 2008. “Standing at the Gates of Europe: Human Behavior and Biogeography in the Southern Carpathians during the Late Pleistocene.” Journal of Anthropological Archaeology 27(4):399–417. http://dx.doi.org/10.1016/j.jaa.2008.02.002. Sandgathe, Dennis M. 2006. Examining the Levallois Reduction Strategy from a Design Theory Point of View. BAR International Series 1417. Oxford: British Archaeological Reports. Semaw, S., P. Renne, J. W. K. Harris, C. S. Feibel, R. L. Bernor, N. Fesseha, and K. Mowbray. 1997. “2.5-Million-Year-Old Stone Tools from Gona, Ethiopia.” Nature 385(6614):333–336. http://dx.doi.org/10.1038/385333a0. Serangeli, Jordi, and Michael Bolus. 2008. “Out of Europe - The Dispersal of a Successful European Hominin Form.” Quartär 55:83–98. Shott, Michael J. 1994. “Size and Form in the Analysis of Flake Debris: Review and Recent Approaches.” Journal of Archaeological Method and Theory 1(1):69–110. http://dx.doi.org/10.1007/BF02229424. Shott, Michael J. 1998. “Status and Role of Formation Theory in Contemporary Archaeological Practice.” Journal of Archaeological Research 6(4):299–329. http://dx.doi.org/10.1007/BF02446082.

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 49 Shott, Michael J. 2008. “Lower Paleolithic Industries, Time, and the Meaning of Assemblage Variation.” In Time in Archaeology: Time Perspectivism Revisited, ed. Simon Holdaway and LuAnn Wandsnider, 61–75. Salt Lake City: University of Utah Press. Smith, Fred H., Ivor Jankovic, and Ivor Karavanic. 2005. “The Assimilation Model, Modern Human Origins in Europe, and the Extinction of Neandertals.” Quaternary International 137(1):7–19. http://dx.doi.org/10.1016/j.quaint.2004.11.016. Surovell, Todd A. 2009. Toward a Behavioral Ecology of Lithic Technology: Cases from Paleoindian Archaeology. Tucson: University of Arizona Press. Torrence, Robin. 1989. “Tools as Optimal Solutions.” In Time, Energy and Stone Tools, ed. Robin Torrence, 1–6. Cambridge: Cambridge University Press. Trinkaus, Erik. 2005. “Early Modern Humans.” Annual Review of Anthropology 34(1):207–230. http://dx.doi.org/10.1146/annurev.anthro.34.030905.154913. van Andel, T. H., S. W. Davies, and B. Wening. 2003. “The Human Presence in Europe during the Last Glacial Period I: Human Migrations and the Changing Climate.” In Neanderthals and Modern Humans in the European Landscape dur- ing the Last Glaciation, ed. T. H. van Andel and S. W. Davies, 31–56. Cambridge: McDonald Institute for Archaeological Research. Voisin, J.-L. 2006. “Speciation by Distance and Temporal Overlap: A New Way of Understanding Neanderthal Evolution.” In Neanderthals Revisited: New Approaches and Perspectives, ed. K. Harvati and T. Harisson, 299–314. New York: Springer. http://dx.doi.org/10.1007/978-1-4020-5121-0_17. Wall, Jeffrey D., Kirk E. Lohmueller, and Vincent Plagnol. 2009. “Detecting Ancient Admixture and Estimating Demographic Parameters in Multiple Human Populations.” Molecular Biology and Evolution 26(8):1823–1827. http://dx.doi.org /10.1093/molbev/msp096. White, J. Peter, and David H. Thomas. 1972. “What Mean These Stones? Ethno- taxonomic Models and Archaeological Interpretations in the New Guinea Highlands.” In Models in Archaeology, ed. David L. Clarke, 275–308. London: Methuen. Wobst, H. Martin. 1974. “Boundary Conditions for Paleolithic Social Systems: A Simulation Approach.” American Antiquity 39(2):147–178. http://dx.doi.org /10.2307/279579. Wolf, Eric R. 1990. “Distinguished Lecture: Facing Power—Old Insights, New Questions.” American Anthropologist 92(3):586–596. http://dx.doi.org/10.1525 /aa.1990.92.3.02a00020. Wolpoff, Milford H., B. Mannheim, A. Mann, John Hawks, Rachel Caspari, Karen Rosenberg, David Frayer, George Gill, and Geoffrey Clark. 2004. “Why Not the

50 B arTON & Riel-Salvatore Neandertals?” World Archaeology 36(4):527–546. http://dx.doi.org/10.1080/00438240 42000303700. Zilhão, João. 2006. “Chronostratigraphy of the Middle-to-Upper Paleolithic Transition in the Iberian Peninsula.” Pyrenae 37:7–84. Zilhão, João. 2007. “TheE mergence of Ornaments and Art: An Archaeological Perspective on the Origins of Behavioral ‘Modernity.’” Journal of Archaeological Research 15(1):1–54. http://dx.doi.org/10.1007/s10814-006-9008-1. Zilhão, João, and Francesco d’Errico, eds. 2003. The Chronology of the Aurignacian and of the Transitional Technocomplexes: Dating, Stratigraphies, Cultural Implications. Lisbon: Instituto Portugues de Arqueologia.

Athic Li Perspective on Ecological Dynamics in the Upper Pleistocene 51

3

Research in the Mimbres Mogollon region of south- The Significance western New Mexico has documented a long period of “Persistent Places” of pithouse use prior to the construction of pueblos. in Shaping Regional Many of the characteristics that we see associated with Settlement History Mimbres pueblos in the Classic period (ad 1100–1150) appear to have developed during the previous Pithouse The Case of the period (ad 200–1000). Continuity in material culture Mimbres Mogollon can be seen in a number of traits, primarily in artifacts and burial practices. In this chapter, I argue that the continuity observed Barbara J. Roth in material culture extends to the social realm, and that this continuity and the relationships it represents are visible on the landscape. The long-term use of specific locations in the Mimbres region was one factor that shaped the economic and social organization of both pithouse and pueblo groups, and ultimately affected the trajectory of culture change in this area. Thus, pit- house and pueblo occupations were intricately linked— spatially, materially, and socially. To this end, I use the concept of “persistent place” and apply it to Pithouse-period and Classic pueblo–period occupations in the Mimbres River valley. Archaeologists­­ working in the Southwest have identified “persistent places” as locales that groups either occupied continu- ously or returned to repeatedly over many generations (Clark and Gilman 2012; Schlanger 1992; Schriever 2012; Varien 1999). Schlanger (1992:97) was the first to formally discuss them and she noted that these locales were usually “marked by cultural features that focus DOI: 10.5876/9781607324942.c003

53 reoccupation.” Not surprisingly, these spots are also located in prime resource areas, generally with access to reliable water, expanses of arable land, and/ or wild plants and game. Schlanger (1992) and Varien (1999, 2002) describe how these persistent places were used to designate and distinguish ancestral land. The ways in which the cultural landscape was perceived and used by prehistoric groups were thus associated with specific social strategies aimed at marking “place” (Zedeño and Bowser 2009; see Schroeder and Goldstein, chapter 7, this volume). Given the evidence for repeated and/or continuous long-term use of specific locales along the Mimbres River by Pithouse and Classic Mimbres pueblo groups, the application of this concept to these loca- tions is useful for addressing the role of the landscape in structuring social interaction and land tenure.

Pes r istent Places and Cultural Landscapes Schlanger (1992:97) defined persistent places in the Mesa Verde region in the northern Southwest as “places that were repeatedly used during long- term occupations of regions.” Although simple at its most basic level, the concept incorporates a number of more complex anthropological phenomena, including the making and marking of place, and how community identity was structured within a particular landscape. Varien (1999, 2002) stressed that community persistence in the Mesa Verde region existed despite substantial household mobility. The Mesa Verde cultural landscape was made up of a series of long-lived communities that provided the social context for interac- tion and movement in the region. Significant to the discussion here, Varien (1999) linked these communities to forms of land tenure and resource access, noting that the placement of residential sites was the means by which house- holds claimed land. In this sense, these places set up a system of land tenure that assured access to productive resources (Varien 1999:208). Adler (1996) and Schriever (2012) have also shown the link between persis- tent places and the social means by which groups asserted land tenure. In a case study of Mesa Verde, Adler (1996) found that stable communities had claims to the best lands, a pattern also observed by Shafer (2003) at NAN Ranch in the Mimbres River valley. Schriever (2012) examined the interconnections between mobility, land tenure, and social identity in the San Simon region of southeastern Arizona, arguing that persistent sites with evidence of multigen- erational occupation of single residential spaces are associated with heritable land tenure. It appears that persistent places in many portions of the Southwest, including the Mimbres area, are physical manifestations of this phenomenon.

54 Roth The concept of persistent places also borrows from recent research on cultural landscapes by both archaeologists and cultural anthropologists (Anschuetz et al. 2001; Ashmore and Knapp 1999; David and Thomas 2008; David and Wilson 2002; Feld and Basso 1996; Gupta and Ferguson 1997; Hirsch and O’Hanlon 1995; Rodman 1992; Snead 2002; Ucko and Layton 1999; Van Dyke 2011; Whittlesey et al. 1997; Zedeño 1997). In these studies, the ecological aspects of land use, although recognized as an important factor in the occupa- tion of an area, remain one component of the use of a landscape, while social and ideological factors are also considered (Whittlesey 2009). The concept of placemaking is essential to understanding the role of these persistent places in past societies (Feld and Basso 1996; Zedeño and Bowser 2009). As Preucel and Meskell (2004:215) note, place is the outcome of the social process of valuing space. Placemaking is thus central to the develop- ment of community identity, and serves to structure social behavior and inter- action (Gupta and Ferguson 1997). Communities transform “physical spaces into meaningful places” (Anschuetz et al. 2001:158). Zedeño and Bowser (2009:5) argue, however, that the analytical challenge is finding meaning in archaeological places using the archaeological record. Anschuetz et al. (2001:175) stress that landscapes (and hence persistent places) are “historically contingent” and derive meaning from both past and current activities. Persistent places are thus inextricably linked with ancestral ties that influence the way that community identity was established. This is supported by a wealth of ethnographic data (see essays in David and Wilson 2002; Hirsch and O’Hanlon 1995) and is perhaps best illustrated in Australia, where Smith (1999) talks of “landscapes imbued with ancestral associations” that were instrumental to social identity (see also Layton 1999; Morphy 1995; Van Dyke 2011; Holdway et al., chapter 5 this volume). The notion of persistent places thus encompasses a range of social phe- nomena, including community identity, social interaction, and land tenure. Landscapes in this broader sense are culturally constructed, so persistent places were rendered meaningful by specific cultural practices (David and Wilson 2002; Feld and Basso 1996; Gupta and Ferguson 1997). For example, ritual or ceremonial items and features may be used to mark or maintain a place and express its significance to others (Anschuetz et al. 2001). Adler (2002:204) argues that the construction of great kivas in the northern Southwest served to symbolize community identity and, in this sense, marked important places. All of these factors appear to be interrelated in many southwestern puebloan communities, including Mimbres communities. The notion of “making place” appears to have been central to the development of community identity and

“Pe rSISTent Places” in Shaping Regional Settlement History 55 both facilitated and structured social interaction. The evidence for persistent places in the Mimbres region is explored below, as is the significance of these places to Pithouse-period and pueblo groups.

Bgack round to the Study The basis for this study comes from long-term work in the Mimbres River valley that focused on delineating settlement patterns and land-use strategies. Interest in the Mimbres region developed early. Bandelier traveled through the area in the 1880s, and Clement Webster, Jesse Walter Fewkes, and Nels Nelson surveyed there in the early 1900s. The focus quickly shifted to excava- tions of the large Classic Mimbres pueblo sites, however, and the 1920s and 1930s witnessed the excavation of many of the major pueblo sites, including Cameron Creek (Bradfield 1931), Mattocks (Nesbitt 1931), Swarts (Cosgrove and Cosgrove 1932), and Galaz (Anyon and LeBlanc 1984). During this same time period Emil W. Haury excavated at the Harris site, a large pithouse site in the central portion of the Mimbres valley. He used data from the site in defining the Mogollon as a distinct cultural group and set up the cultural sequence for the Pithouse period (Haury 1936). Although the Cosgroves did some survey work while they were excavating at Swarts, Mimbres archaeology prior to the 1960s primarily focused on excavation. The University of Michigan Mimbres Survey, led by Arthur J. Jelinek, was conducted during the winter and spring of 1967 and represents the first major archaeological reconnaissance in the region. The survey focused on the Mimbres River valley and surrounding areas, including the uplands adjacent to the river, the Deming plain to the south, drainages to the east of the Mimbres River and west toward the Cliff Valley. Jelinek’s survey, James Fitting’s investi- gations in the Burro Mountains and Cliff Valley (Fitting 1971, 1972, 1973), and Don Graybill’s survey of the Gila National Forest (Graybill 1975) represent the first systematic attempts to place Mimbres sites in a temporal and regional context. By documenting significant variability in site size, distribution, and temporal span of occupation at sites throughout the region, they showed that what it meant to be “Mimbres Mogollon” was not just living in large, cobble adobe pueblos along the river. Subsequent to these early surveys, archaeologists from the Mimbres Foundation conducted a major survey of the Mimbres valley (Blake et al. 1986), identifying sites dating from the Early Pithouse period through the Cliff phase (ca ad 200 to 1450). The survey data indicated substantial population increase over time during the Classic Mimbres period, followed by rapid population

56 Roth decline. This decline has been linked by a number of researchers to resource depletion and drought (Creel 2006a; Minnis 1985), although Pool (2013) has recently argued that these stressors were not the sole cause of the reorganiza- tion of Mimbres society after ad 1000, and social factors likely played some role. One of the clearest patterns to come out of the Mimbres Foundation survey and subsequent excavations in the valley was the demonstration that most of the largest pueblos in the valley such as Galaz, Old Town, and NAN Ranch had large pithouse components beneath them, although the extent of these components could not always be discerned (Anyon and LeBlanc 1984; Creel 2006b; Shafer 2003, 2006). The distribution of these substantial pueblo sites with large pithouse com- ponents has important implications for understanding population dynamics and social interaction during the Pithouse and pueblo periods, encompass- ing both the ecological reasons for site locations (Swanson et al. 2012) and the social strategies involved in their long-term use. In the remainder of this chapter, I discuss how settlement data can be used to define persistent places in order to delineate these processes in the Mimbres region.

Pes r istent Places in the Mimbres River Valley In applying the concept of persistent place to the Mimbres River valley, I argue that certain well-watered areas along the river adjacent to expanses of arable land exhibit evidence of long-term use that is indicative of ancestral marking and land tenure, as has been documented in other portions of the Southwest. This pattern is clearly visible during the pueblo period, but it appar- ently began during the Pithouse period. Survey and excavation data are used to illustrate that a broader view of landscapes can strengthen our understanding of the social dynamics that were occurring during the Pithouse and pueblo periods. The pattern of pithouse occupations adjacent to and beneath the larger pueblos is consistent across the valley and extends into the adjacent tribu- tary valleys at sites such as Gatton’s Park and Ponderosa Ranch (Stokes 1995). If these locales are viewed from the perspective of persistent places located within a cultural landscape, then it is apparent that they represent a complex of linked sites. Three distinct portions of the Mimbres River valley are discussed here as exhibiting evidence of being persistent places, culminating in the construc- tion of large, well-known pueblo sites. These include NAN Ranch and sur- rounding pithouse communities in the southern valley (Shafer 2003), the Beauregard, Montezuma, and Mitchell complex in the northern portion of

“Pe rSISTent Places” in Shaping Regional Settlement History 57 Figure 3.1. “Persistent Places” in the Mimbres valley.

the valley, and the McAnally, Harris, and Mattocks sites in the north-central valley (figure 3.1). Other areas may also represent persistent places; Creel’s (2006b) long-term work at Old Town indicates that it, too, represents this same pattern. Sites such as Swarts (Cosgrove and Cosgrove 1932) and Three Circle/Elk Ridge may also be persistent places, but the data to evaluate this are less robust than for the three localities discussed here. The three places discussed in this chapter are included because they are situ- ated in different portions of the Mimbres valley where long-term investment

58 Roth in particular locations on the landscape is clearly illustrated. The characteris- tics of these persistent places include: 1. All are located in portions of the valley with access to expanses of arable land and dense stands of wild resources. 2. All exhibit evidence of long-term occupation, from at least ad 500 (and sometimes substantially earlier) to 1130. When present, the Early Pithouse sites (ad 200–550) associated with these persistent places are often spatially distinct from the Late Pithouse (ad 550–1000) and large pueblo components, as they are located on high landforms directly above the later components. TheseE arly Pithouse components likely represent the initial settlement of the valley; visible domestic and public architecture was used as a material signature on these high points to mark place and establish initial land tenure. The Late Pithouse and pueblo components are on river terraces topographically below these early components. In all but the Harris/Mattocks case, the pueblo components have large pithouse components beneath them. 3. Ritual architecture, in the form of large, deep communal structures dur- ing the Pithouse period and open plazas surrounded by room blocks in the pueblo period, are highly visible and would have served to mark the cultural landscape. Adler (2002) has emphasized how community identity was expressed through ritual architecture in the northern Southwest, and it appears that it served a similar function in the Mimbres valley for asserting ceremonial rights and establishing a sense of community identity. In the remainder of this chapter, I describe how each of these characteris- tics is manifested in the three locations argued to be persistent places in the Mimbres valley.

NAN Ranch Shafer’s (2003) long-term work at NAN Ranch has provided excellent data on the Pithouse-period sites around the NAN Ranch pueblo and on the sequence of development from the Georgetown (ad 550–650) through the Classic Mimbres period at the NAN Ranch Ruin itself. This work has docu- mented Early and Late Pithouse and pueblo occupations that were centered on a very productive portion of the Mimbres floodplain, with a reliable water supply, abundant wild resources, and expanses of arable land. The Y Bar site is anE arly Pithouse–period site located approximately 0.5 km south of NAN pueblo. It consists of five visible pithouse depressions and

“Pe rSISTent Places” in Shaping Regional Settlement History 59 the remains of a large communal structure. The site is located on a high ridge, consistent with other Early Pithouse–period sites in the area. Shafer (2003) argues that it was occupied seasonally and represents a relatively short-term occupation. The placement of a site with a large communal structure on a prominent hilltop above a very productive portion of the river may have been the first step in identifying this location as important, beginning the process of marking place. Household mobility is compatible with this, as the persis- tence of the place is the critical variable; both Schlanger (1992) and Varien (2002) have shown that a persistent place does not necessarily have to be persistently occupied. The idea that the area around NAN Ranch was an important place is supported by the presence of the full sequence of Late Pithouse–period occupation beneath the NAN pueblo (starting at ad 600). Shafer (2003, 2006) has detailed the complex series of changes that occurred as groups at NAN established their land tenure and began irrigating, including changes in architecture, storage, and ceremonial architecture. He notes that there is continuity in the occupation at NAN, as shown by repair and rebuilding of rooms, and proposes a lineal sequence of generations living there (Shafer 2003). He argues that “core households” established the initial room blocks at NAN pueblo and retained rights to prime agricultural land. The mark- ing of place by these core households was thus linked to land tenure and resource access.

Beauregard, Montezuma, and Mitchell A similar pattern is observed farther north in the valley, where the Beauregard, Montezuma, and Mitchell sites show long-term use of a lush portion of the river from the Pithouse through pueblo periods (figure 3.2). The Pithouse-period Beauregard site is located atop a large ridge 70 m above the floodplain (Creel and Anyon 2003). Twenty-three pithouse depressions were identified there in two separate concentrations, separated by a natural saddle. The houses date to theG eorgetown (ad 550–650) and San Francisco (ad 650–750) phases of the Late Pithouse period and possibly extend into the early portion of the Three Circle phase (ad 750–1000). A communal structure dating to the San Francisco/early Three Circle phase is also present on this ridgetop (Anyon n.d.). During the Three Circle phase, groups moved downslope to the Monte­ zuma site, located on a finger ridge directly beneath the Beauregard site and above a broad expanse of floodplain, perhaps to get closer to their

60 Roth Figure 3.2. Montezuma site (photo taken from the Beauregard site). agricultural fields. Mimbres Foundation archaeologists view these occupa- tions as complementary, with groups moving to Montezuma after aban- doning Beauregard (Anyon n.d.). Significant to the argument here, they retained the use of the same landform. A Classic Mimbres pueblo was then built on top of the Three Circle pithouses and this was occupied into the late Classic period. A second pueblo, the Mitchell site, was built on a lower ter- race across the river from Montezuma, yet still adjacent to the same expanse of arable land and other resources. Unfortunately, the Mitchell site was so severely looted that very little information is available on it. It is possible that the construction of this pueblo was either subsequent to the construc- tion of the room blocks at Montezuma or that it was built at the same time as the Montezuma pueblo rooms by a separate group of households. Its proximity to Montezuma indicates that it is part of the same community, however. These three sites document continuity of use of this particular por- tion of the river floodplain over many centuries. As at NAN, this shows long-term commitment to a particular place starting in the early portion of the Pithouse period.

“Pe rSISTent Places” in Shaping Regional Settlement History 61 McAnally, Harris, and Mattocks On initial inspection, an exception to this pattern appears to be the Mattocks site, a large Classic-period pueblo in the north-central portion of the valley that lacks a substantial Late Pithouse–period component. If the cultural landscape rather than the individual site is used to examine the concept of persistent place, however, then this area exhibits a very similar pattern to that observed at the other two persistent places discussed here. The Early Pithouse–period McAnally site is located on an isolated hilltop directly above the Mattocks ruin (figure 3.1; Diehl and LeBlanc 2001). Fifteen pithouse depressions were identified at the site and two houses were excavated. No communal structures were found, but it is possible that one of the unexcavated depressions repre- sents an early communal structure. The Harris site, a large Late Pithouse–period site with Georgetown, San Francisco, and Three Circle phase components and communal structures dat- ing to all three phases, is located on a river terrace approximately one mile north of Mattocks (figure 3.1; Haury 1936; Roth 2015). Unlike other Late Pithouse–period sites in the area, the Harris site lacks a pueblo component. It appears that instead of building a pueblo atop their pithouses, groups occupy- ing the Harris site dispersed in the late ad 900s, with some households mov- ing downstream to the Mattocks site. TheE arly Pithouse–period McAnally site had already, in perceptible ways, marked this location on the landscape as an important spot. Although it is not clear at this time why the people from Harris moved to Mattocks, in keeping within the same “place” on the cultural landscape, it apparently represents a persistent place akin to those cases where major pueblos were built atop earlier pithouse sites.

Upland Sites Although the best data for discerning persistent places come from the Mimbres River valley as a result of more extensive survey and excavation work in that area, work at the Lake Roberts Vista site, an upland site in a tributary valley to the Gila River (figure 3.3), suggests that it also is a persistent place, although it shows a somewhat different pattern than that seen in the Mimbres River valley. The site is located on a ridgetop above Sapillo Creek, and contains Georgetown, San Francisco, Three Circle, and Classic Mimbres–period com- ponents, all located on the same landform (Roth 2007; Stokes and Roth 1999). A large communal structure was built on the edge of the site during the Three Circle phase. This communal structure is larger than would be expected for the size of the pithouse component, and Stokes and Roth (1999) have argued

62 Roth Figure 3.3. Location of Lake Roberts Vista site, Gila National Forest.

that it served the surrounding community, which included other pithouse sites. The site thus represents a prominent visible place on the landscape that was used over long periods of time. A similar situation may be present elsewhere in the Sapillo Valley at the Ponderosa Ranch and Gatton’s Park sites (Stokes 1995), but no professional excavations have been done at these sites so this inference is based solely on surface data.

“Pe rSISTent Places” in Shaping Regional Settlement History 63 Discussion Persistent places in the Mimbres River valley represent long-term ties to specific locations that can reasonably be assumed to have been occupied by related households over many centuries. These places are marked on the land- scape by both domestic and ceremonial architectural features. The earliest indication of the marking of place extends back to the Early Pithouse period (at Y Bar and McAnally) or Georgetown phase (at Beauregard). These early sites are all positioned on prominent hill- or ridgetops. These high landforms would have increased their visibility to other groups, thus facilitating interac- tion and exchange (Diehl 2001), and would have served to mark significant resource zones along the river, establishing initial ties to the land. Shafer (2003, 2006) has discussed the presence of core households at NAN Ranch and other pueblo sites in the Mimbres valley, arguing that these house- holds used ancestral ties dating back to the Late Pithouse period to claim the best agricultural lands. Adler (2002) has also discussed the importance of ancestral ties in delimiting land tenure. Persistent places provided visible means by which groups established access to land and resources, thereby communicating this to others in an unambiguous way. It appears from this marking of place that the cultural landscape was an important aspect of social organization and interaction throughout the Mimbres cultural sequence. Persistent places remained an essential component of the landscape despite relatively dramatic changes in subsistence, household organization, ritual (Creel and Anyon 2003), and ideology (Gilman et al. 2014), suggesting that these cultural ties to the landscape were perhaps paramount in Mimbres soci- ety through time. Another role of these persistent places was to reinforce community identity. Anschuetz et al. (2001:158) have discussed how communities transform “physi- cal spaces into meaningful places.” The continuity of occupation observed at these sites points to the signaling of long-term commitment to that particular location. This is illustrated in the presence of large visible communal struc- tures that would have communicated this sense of shared identity to others. The cultural landscape was thus crucial to the continued sense of the commu- nity and the communication of that to others. All of this changed at the end of the Classic period (ca. ad 1130). At this time, severe climatic conditions and depletion of resources led to the aban- donment of many large pueblos in the central and northern portion of the Mimbres River valley (Creel 2006a; Minnis 1985; Pool 2013). The succeed- ing Reorganization period (ad 1150–early 1200s) witnessed movement into smaller hamlets outside the valley (Hegmon et al. 1998; Nelson 1999; Nelson

64 Roth et al. 2006) and into adobe pueblos adjacent to some of the larger ones within the valley, especially in the southern portion (Creel 1997). The long-term strat- egy of being “place-focused” may have created conditions of population size and resource depletion that could not be sustained with a continued focus on a single place in the valley. The use of place in constructing Mimbres com- munity identity changed, and perhaps with it the notion of what Mimbres was, also changed.

Cuoncl sions Persistent places, locales that were repeatedly used over many generations, are present in the Mimbres River valley in areas that provided key resources (water, land, and flora). These places were marked with visible features on the landscape as early as the Early Pithouse period. This marking of place and subsequent long-term use was apparently a means by which households obtained and retained access to these resources. They also structured commu- nity identity and likely facilitated interaction and movement within the valley. They remained important components of the occupation of the valley through the pueblo period, serving to define and specify community identity and land tenure in visible, socially meaningful ways. Continuity of occupation has long been observed in the Mimbres River valley and has consistently been interpreted as important to understanding cultural change in the region. By taking a cultural landscape approach and interpreting these occupations using the concept of persistent places, it is pos- sible to begin to delineate the role that social factors played in the way these groups adapted and interacted in the valley.

References Adler, Michael A. 1996. “Land Tenure, Archaeology, and the Ancestral Pueblo Social Landscape.” Journal of Anthropological Archaeology 15(4):337–371. http://dx.doi.org /10.1006/jaar.1996.0013. Adler, Michael A. 2002. “Negotiating the Village: Community Landscapes in the Late Pre-Historic American Southwest.” In Inscribed Landscapes: Marking and Making Place, ed. Bruno David and Meredith Wilson, 200–216. Honolulu: University of Press. Anschuetz, Kurt F., Richard H. Wilshusen, and Cherie L. Scheick. 2001. “An Archaeology of Landscapes: Perspectives and Directions.” Journal of Archaeological Research 9(2):157–211. http://dx.doi.org/10.1023/A:1016621326415.

“Pe rSISTent Places” in Shaping Regional Settlement History 65 Anyon, Roger, and Steven A. LeBlanc. 1984. The Galaz Ruin. Albuquerque: Maxwell Museum of Anthropology. Ashmore, Wendy, and A. Bernard Knapp. 1999. Archaeologies of Landscape: Contemporary Perspectives. Malden, MA: Blackwell Publishing. Blake, Michael, Steven A. LeBlanc, and Paul E. Minnis. 1986. “Changing Settlement and Population in the Mimbres Valley, SW New Mexico.” Journal of Field Archaeology 13:439–464. Bradfield, Wesley. 1931. Cameron Creek Village: A Site in the Mimbres Area in Grant County, New Mexico. Monograph No. 1. Santa Fe: School of American Research. Clark, Jeffrey J., and Patricia A.G ilman. 2012. “Persistent and Permanent Pithouse Places in the Basin and Range Province of Southeastern Arizona.” In Southwestern Pithouse Communities, ad 200–900, ed. Lisa C. Young and Sarah A. Herr, 61–77. Tucson: University of Arizona Press. Cosgrove, H. S., and C. B. Cosgrove. 1932. The Swarts Ruin: A Typical Mimbres Site in Southwestern New Mexico. Papers of the Peabody Museum of American Archaeology and Ethnology 15. Cambridge: Harvard University. Creel, Darrell. 1997. “Interpreting the End of the Mimbres Classic.” In Prehistory of the Borderlands, ed. John Carpenter and Guadalupe Sanchez, 25–32. Arizona State Museum Archaeological Series No. 186. Tucson: University of Arizona. Creel, Darrell. 2006a. “Environmental Variation and Prehistoric Culture in the Mimbres Area, Southwestern New Mexico.” In Environmental Change and Human Adaptation in the Ancient American Southwest, ed. David E. Doyel and Jeffrey S. Dean, 204–213. Salt Lake City: University of Utah Press. Creel, Darrell. 2006b. Excavations at the Old Town Ruin, Luna County, New Mexico, 1989–2003. Sante Fe: US Bureau of Land Management, New Mexico State Office. Creel, Darrell, and Roger Anyon. 2003. “New Interpretations of Mimbres Public Architecture and Space: Implications for Cultural Change.” American Antiquity 68(1):67–92. http://dx.doi.org/10.2307/3557033. David, Bruno, and Julian Thomas, eds. 2008. Handbook of Landscape Archaeology. Walnut Creek, CA: Left Coast Press. David, Bruno, and Meredith Wilson. 2002. Inscribed Landscapes: Marking and Making Place. Honolulu: University of Hawaii Press. Diehl, Michael W. 2001. “Competing Models of Upland Mogollon Pithouse Period Lifestyles.” In Early Pithouse Villages of the Mimbres Valley and Beyond, ed. Michael W. Diehl and Steven A. LeBlanc, 25–35. Papers of the Peabody Museum of Archaeology and Ethnology, vol. 83. Cambridge, MA: Harvard University.

66 Roth Diehl, Michael W., and Steven A. LeBlanc, eds. 2001. Early Pithouse Villages of the Mimbres Valley and Beyond. Papers of the Peabody Museum of Archaeology and Ethnology, vol. 83. Cambridge, MA: Harvard University. Feld, Steven, and Keith H. Basso. 1996. Senses of Place. Santa Fe: School of American Research Press. Fitting, James E. 1971. The Burris Ranch Site, Dona Ana County, New Mexico. Southwest New Mexico Research Reports 1. Cleveland: Case Western Reserve University. Fitting, James E. 1972. “Preliminary Notes on Cliff Valley Settlement Patterns.” Artifact 10:15–30. Fitting, James E. 1973. Four Archaeological Sites in the Big Burro Mountains of New Mexico. Center of Anthropological Study Monograph 1. Las Cruces, NM: Center of Anthropological Study. Gilman, Patricia A., Marc Thompson, andK ristina Wyckoff. 2014. “Ritual Change and the Distant: Mesoamerican Iconography, Scarlet Macaws, and Great Kivas in the Mimbres Region of Southwestern New Mexico.” American Antiquity 79(1):90– 107. http://dx.doi.org/10.7183/0002-7316.79.1.90. Graybill, Donald A. 1975. Mimbres-Mogollon Adaptations in the Gila National Forest, Mimbres District, New Mexico. US Forest Service Southwestern Region Archaeological Report 9. Albuquerque: US Forest Service. Gupta, Akhi, and James Ferguson. 1997. “Culture, Power, and Place: Ethnography at the End of an Era.” In Culture, Power, Place: Explorations in Critical Anthropology, ed. Akhi Gupta and James Ferguson, 1–29. Durham: Duke University Press. http://dx.doi.org/10.1215/9780822382089. Haury, Emil W. 1936. The Mogollon Culture of Southwestern New Mexico. Medallion Papers No. 20. Globe, Arizona: Gila Pueblo. Hegmon, Michelle, Margaret C. Nelson, and Susan M. Ruth. 1998. “Abandonment and Reorganization in the Mimbres Region of the American Southwest.” American Anthropologist 100(1):148–162. http://dx.doi.org/10.1525/aa.1998.100.1.148. Hirsch, Eric, and Michael O’Hanlon. 1995. The Anthropology of Landscape: Perspectives on Place and Space. Oxford: Clarendon Press. Layton, Robert. 1999. “The Alawa Totemic Landscape: Ecology, Religion, and Politics.” In The Archaeology and Anthropology of Landscape, ed. Peter J. Ucko and Robert Layton, 219–239. London: Routledge. http://dx.doi.org/10.4324/978020 3202449_chapter_16. Minnis, Paul E. 1985. Social Adaptation to Food Stress: A Prehistoric Southwestern Example. Chicago: University of Chicago Press.

“Pe rSISTent Places” in Shaping Regional Settlement History 67 Morphy, Howard. 1995. “Landscape and the Reproduction of the Ancestral Past.” In The Anthropology of Landscape: Perspectives on Place and Space, ed. Eric Hirsch and Michael O’Hanlon, 184–209. Oxford: Clarendon Press. Nelson, Margaret C. 1999. Mimbres during the Twelfth Century: Abandonment, Continuity, and Reorganization. Tucson: University of Arizona Press. Nelson, Margaret C., Michelle Hegmon, Stephanie Kulow, and Karen Gust Schollmeyer. 2006. “Archaeological and Ecological Perspectives on Reorganization: A Case Study from the Mimbres Region of the U.S. Southwest.” American Antiquity 71(3):403–432. http://dx.doi.org/10.2307/40035359. Nesbitt, Paul. 1931. The Ancient Mimbrenos, Based on Investigations at the Mattocks Ruin, Mimbres Valley, New Mexico. Logan Museum Bulletin 4. Beloit, WI: Logan Museum. Pool, Michael D. 2013. “Mimbres Mogollon Farming: Estimating Prehistoric Agricultural Production during the Classic Mimbres Period.” In Soils, Climate, and Society, ed. John D. Wingard and Sue E. Hayes, 85–107. Boulder: University Press of Colorado. Preucel, Robert W., and Lynn Meskell. 2004. “Places.” In A Companion to Social Archaeology, ed. Lynn Meskell and Robert W. Preucel, 215–229. Malden, MA: Blackwell. Rodman, Margaret C. 1992. “Empowering Place: Multilocality and Multivocality.” American Anthropologist 94(3):640–656. http://dx.doi.org/10.1525/aa.1992.94.3 .02a00060. Roth, Barbara J. 2007. “The Late Pithouse Period Occupation of the Lake Roberts Vista Site.” In Exploring Variability in Mogollon Pithouses, ed. Barbara J. Roth and Robert Stokes, 5–11. Arizona State University Anthropological Research Papers No. 58. Tempe: Arizona State University. Roth, Barbara J., ed. 2015. “Archaeological Investigations at the Harris Site, LA 1867, Grant County, Southwestern New Mexico.” Report on file, Department of Anthropology, University of Nevada, Las Vegas. Schlanger, Sarah. 1992. “Recognizing Persistent Places in Anasazi Settlement Systems.” In Space, Time, and Archaeological Landscapes, ed. Jacqueline Rossignol and LuAnn Wandsnider, 91–112. New York: Plenum Press. http://dx.doi.org /10.1007/978-1-4899-2450-6_5. Schriever, Bernard A. 2012. “Mobility, Land Tenure, and Social Identity in the San Simon Basin of Southeastern Arizona.” Kiva 77(4):413–438. http://dx.doi.org /10.1179/kiv.2012.77.4.003. Shafer, Harry J. 2003. Mimbres Archaeology at the NAN Ranch Ruin. Albuquerque: University of New Mexico Press.

68 Roth Shafer, Harry J. 2006. “Extended Families to Corporate Groups: Pithouse to Pueblo Transformation of Mimbres Society.” In Mimbres Society, ed. Valli Powell-Marti and Patricia A. Gilman, 15–31. Tucson: University of Arizona Press. Smith, Claire. 1999. “Ancestors, Place, and People: Social Landscapes in Aboriginal Australia.” In The Archaeology and Anthropology of Landscape, ed. Peter J. Ucko and Robert Layton, 189–205. London: Routledge. http://dx.doi.org/10.4324/97802 03202449_chapter_14. Snead, James E. 2002. “Ancestral Pueblo Trails and the Cultural Landscape of the Pajarito Plateau, New Mexico.” Antiquity 76(293):756–765. http://dx.doi.org /10.1017/S0003598X00091201. Stokes, Robert. 1995. “Prehistoric Settlement Patterns in the Sapillo Creek Valley, Gila National Forest, New Mexico.” MA thesis, Eastern New Mexico University, Portales, NM. Stokes, Robert, and Barbara J. Roth. 1999. “Mobility, Sedentism, and Settlement Patterns in Transition: The Late Pithouse Period in the Sapillo Valley, New Mexico.” Journal of Field Archaeology 26:423–434. Swanson, Steve, Roger Anyon, and Margaret C. Nelson. 2012. “Southern Mogollon Pithouse Period Settlement Dynamics, Land Use, and Community Development, ad 200–1000.” In Southwestern Pithouse Communities, ad 200–900, ed. Lisa C. Young and Sarah A. Herr, 95–109. Tucson: University of Arizona Press. Ucko, Peter J., and Robert Layton, eds. 1999. The Archaeology and Anthropology of Landscape. New York: Routledge. http://dx.doi.org/10.4324/9780203202449. Van Dyke, Ruth. 2011. “Anchoring Identities: Iconic Landforms across San Juan Time and Place.” In Movement, Connectivity, and Landscape Change in the Ancient Southwest, ed. Margaret C. Nelson and Colleen Strawhacker, 403–422. Boulder: University Press of Colorado. Varien, Mark D. 1999. Sedentism and Mobility in a Social Landscape: Mesa Verde and Beyond. Tucson: University of Arizona Press. Varien, Mark D. 2002. “Persistent Communities and Mobile Households.” In Seeking the Center Place, ed. Mark D. Varien and Richard H. Wilshusen, 163–184. Salt Lake City: University of Utah Press. Whittlesey, Stephanie M. 2009. “Mountains, Mounds, and Meaning: Metaphor in the Hohokam Cultural Landscape.” In The Archaeology of Meaningful Places, ed. Brenda J. Bowser and Maria Nieves Zedeño, 73–89. Salt Lake City: University of Utah Press. Whittlesey, Stephanie M., Richard Ciolek-Torrello, and Jeffrey H. Altschul, eds. 1997. Vanishing River: Landscape and Lives of the Lower Verde Valley. Tucson: Statistical Research Institute Press.

“Pe rSISTent Places” in Shaping Regional Settlement History 69 Zedeño, Maria Nieves. 1997. “Landscapes, Land Use, and the History of Territory Formation: An Example from the Puebloan Southwest.” Journal of Archaeological Method and Theory 4(1):67–103. http://dx.doi.org/10.1007/BF02428059. Zedeño, Maria Nieves, and Brenda J. Bowser. 2009. “The Archaeology of Meaningful Places.” In The Archaeology of Meaningful Places, ed. Brenda J. Bowser and Maria Nieves Zedeño, 1–14. Salt Lake City: University of Utah Press.

70 Roth 4

As durable material items representing hominin actions Reductive Technology and activities over much of the past 3.3 million years, and the Epipaleolithic chipped-stone artifacts are one of the main lines of of the Middle East evidence about prehistoric behaviors. It is no sur- and North Africa prise, then, that the classification and interpretation of chipped-stone assemblages often has taken on a life of its own, with considerable researcher investment into Deborah I. Olszewski one paradigm or another, and lively debates between proponents of contrasting viewpoints. The well-known Old World typologies of Bordes (1979), Clark and Kleindienst (1974), L.S.B. and M. D. Leakey (1951; 1966), de Sonneville-Bordes and Perrot (1954, 1955, 1956), Tixier (1963), and updates or geographical expansions of them (e.g., Debénath and Dibble 1994; Demars and Laurent 1992; Hours 1974; Goring-Morris 1987) cre- ated standardized lists of tool types that allowed com- parability between assemblages. Most of the typologies characterize retouched artifacts (although some excep- tions exist, for example, Bordes’s inclusion of tech- nological types such as Levallois flakes and Levallois points in his tool typology). Most typologies, however, are ultimately heuristic devices used to order a diverse array of variability in chipped-stone artifact morphol- ogy, and few archaeological typologies actually contain discrete types (Adams and Adams 1991). Several decades ago, it was noted that the interpreta- tion of lithic types has a long history of alliance with a phylogenetic or history-based paradigm ( Jelinek 1976:26). One of the most famous alternative paradigms DOI: 10.5876/9781607324942.c004

71 (lithic types are associated with activities) was the focus of the “Bordes-Binford debate” (Binford and Binford 1966; Bordes 1961). More recently, researchers such as Dibble (1987, 1995), Toth (1985), and McPherron (1999), using contexts in the Old World, and Frison (1968) and Flenniken and Wilke (1989; but see Bettinger et al. 1991) using contexts in the New World, have suggested that lithic reduction processes form a powerful paradigm for the interpretation of lithic types and assemblages. Although current research has added more care- ful examinations of the lithic technology(ies) associated with assemblages, par- ticularly chaîne opératoire, refitting, and reduction approaches (e.g., Bar-Yosef and Van Peer 2009; Bleed 2001; Bourguignon et al. 2004; Lindly et al. 2000; Sellet 1993), these themselves often are linked to the historical reconstruction paradigm (but see Carr and Bradbury 2011; Tostevin 2011; among others). None of the alternative paradigms, therefore, have completely overturned an adher- ence to the notion that lithic types (and therefore the assemblages that con- tain them) are linked to specific groups of people in prehistory, particularly for research undertaken in the Old World (Clark 1993; Shea 2008:207; Shea 2013). Added to this is the fact that original interpretations of some specific lithic types have been either broadened or completely reversed in some cases. For example, burins often are associated with an interpretation of engraving or incising (e.g., Leguay 1877:286–287), although some researchers attributed addi- tional functions to burins (e.g., Becker and Wendorf 1993; Knecht 1988; Vaughan 1985). It has been only recently that scholars have come to accept that burins may have had many functions and, in some cases, may not have been tools, but cores (Barton et al. 1996; Olszewski 2007a; see also articles in McPherron 2007). On the other hand, an instance of complete reversal of interpretation is seen in the contrast between Mary Leakey’s (1966:463) focus on Oldowan chopper and chopping tools as the important “desired” elements (i.e., the “tools”) and Toth’s (1985, 1987) analyses suggesting that the flakes struck from the choppers and chopping tools are the “desired” tools, while the choppers and chopping tools are actually cores. Here, of course, such flakes usually are not retouched, and thus would not be identified as formal tools in most typologies. Much of the debate engendered by different paradigms for lithic assem- blages in the Old World has used assemblages of the Middle Paleolithic/ Middle Stone Age or earlier. It might be asked if these interpretative issues also are applicable to later periods, such as the Epipaleolithic, which contains a multitude of named archaeological cultures, often within small geographic regions (e.g., Olszewski 2008:51–53). Many of the distinctions made between these Epipaleolithic archaeological cultures are those of lithic types, especially microlith forms, as well as the presence or absence of microburin technique.

72 Olszewski Although microlith types are conceptualized as distinct forms, anyone who has classified these into type categories knows that each type contains a wide range of variability (as predicted by Adams and Adams 1991). It is only rarely, however, that potential overlap between microlith types and its implica- tions has been considered (e.g., Neeley and Barton 1994; Olszewski 1993a). Following Jelinek (1976), it is possible that such potential overlap is one facet that may be best explained using a lithic reduction sequence paradigm (see also Rollefson, chapter 10, this volume, for a biface example). This chapter focuses on five Epipaleolithic lithic assemblages from various regions in the Middle East and North Africa and examines whether lithic reduction pro- cesses link some microlith types into a sequence of reduction for a single arti- fact, while also considering other reduction processes and situational contexts that may explicate cases of dominance in an assemblage of specific tool types.

T he Sites and Lithic Assemblages The formation of lithic assemblages reflects cumulative human behav- iors, which result from the decisions people made as they engaged in vari- ous activities during multiple visits to site locales; in some cases, natural taphonomic processes also can impact site assemblage formation (e.g., the so-called Tayacian, which has been shown to result from natural taphonomy [Dibble et al. 2006]). The sample of Epipaleolithic assemblages used in this essay are from the eastern Middle East (), the western Middle East (Tor Sageer and Tor at-Tareeq), and western North Africa (Smugglers’ Cave [Grotte des Contrebandiers]; figure 4.1). These assemblages are comparable in the sense that all contain nongeometric forms of microliths and microburin technique, which allows for a broad comparison of lithic attributes that can be used to examine contextual features of lithic reduction processes at specific sites and thus serve as a proxy of human behaviors at those locales; most sites also contain geometric microliths to varying degrees. By using typological names (e.g., nongeometric microliths) here, it is not the intention to reify these but to show that this terminology can be used and thought of in ways beyond standard typological thinking (e.g., discrete categories or immutable types). All of the assemblages used here were studied by the author.

Warwasi Rockshelter, The rockshelter at Warwasi, located in the Tang-i-knisht valley in the foot- hills of the Zagros Mountains some 12 km to the northeast of Kermanshah,

Re dUCTIVe Technology and the Epipaleolithic 73 Figure 4.1. Archaeological sites mentioned in the text.

was test excavated by Bruce Howe, working with Robert Braidwood in 1960 (Braidwood et al. 1961); the trench was approximately 8 m × 1 m. The rock- shelter contains a long sequence of Middle Paleolithic, Upper Paleolithic, and Epipaleolithic deposits; analyses of these have been published elsewhere (Dibble and Holdaway 1993; Olszewski 1993a, 1996, 1999, 2001b, 2007a, 2007b; Olszewski and Dibble 1994, 2006). TheE pipaleolithic lithic assemblages at Warwasi, which represent the upper 1.5 m of the sequence, are attributed to the Zarzian industry. There are 15 arbitrary levels of 10 cm each that can be grouped into four Zarzian horizons—Unit 1 is the oldest and Unit 4 is the youngest (Olszewski 1993b). Two of these (Units 2 and 4) are used in this study (table 4.1). The Zarzian industry was originally identified by Garrod (1930) during her 1928 excavations at the Iraqi site of Zarzi Cave. The industry is characterized by bladelet technology and has a sequence in which nongeometric microliths typify the early horizons and geometric forms (mainly scalene triangles and trapezes, with rarer lunates) are added during later phases. These character- istics also are present at Warwasi, where Unit 1 contains only nongeomet- rics (especially lamelles Dufour and pointed bladelets), and Units 2–4 con- tain nongeometrics (pointed, truncated, and curved backed bladelets) and

74 Olszewski Table 4.1. Overview of site tool assemblages. Smugglers’ Warwasi 2 Warwasi 4 Tor Sageer Tor at-Tareeq Cave Tools* % N % N % N % N % N ES 13.6 119 2.7 33 6.1 55 2.9 31 9.0 144 BU 3.1 27 0.2 3 2.5 23 3.0 32 3.2 52 BP 4.0 35 0.9 11 1.2 11 0.7 7 5.2 84 BO 1.3 11 5.5 68 0.7 6 0.6 6 0.5 8 TR 3.6 32 2.1 26 2.1 19 6.0 64 1.4 22 NG 28.2 248 36.7 450 64.8 586 55.0 587 23.9 384 GE 11.0 97 7.4 91 2.0 18 0.4 4 0.6 10 ND 17.1 150 28.7 352 5.3 48 9.9 106 4.1 66 RP 13.2 116 12.6 155 12.7 115 19.6 209 24.3 390 SP 3.8 33 1.3 16 1.1 10 1.0 11 25.5 409 MT 0.8 7 1.5 18 0.3 3 0.8 9 0.5 8 VA 0.3 3 0.3 4 1.1 10 0.1 1 1.7 27 Total (878) (1,227) (904) (1,067) (1,604) *ES = endscrapers; BU = burins; BP = backed pieces; BO = borers; TR = truncations; NG = nongeo- metric microliths; GE = geometric microliths; ND = notch/denticulates; RP = retouched pieces; SP = special tools; MT = multiple tools; VA = varia.

geometrics (mainly scalene and elongated scalene triangles). The rare lunates are present in Units 3 and 4, where they coincide with the presence of curved backed bladelets. Microburin technique is present. No radiocarbon dates are available for Warwasi, although Units 2 and 4 are comparable to the Zarzian from the rockshelter at Palegawra in (Braidwood and Howe 1960:57–59), which was radiocarbon dated using bone to 14,400 ± 760 (UCLA–1703A) and 13,350 ± 460 (UCLA–1714D) uncal BP (Turnbull and Reed 1974:84).

Tor Sageer and Tor at-Tareeq, Jordan These two sites represent an open-air context (Tor at-Tareeq) and a rock- shelter (Tor Sageer) in the Wadi al-Hasa region in the western highlands of Jordan, about 120 km southeast of Amman. Tor at-Tareeq was first tested in 1984 with a 44 m × 1 m test trench, which was divided into eight 5 m × 1 m sections and one 4 m × 1 m section (Clark et al. 1988:258). The site is present in the upper 15 m of the trench (Steps A, B, and C). Two additional 2 m × 2 m

Re dUCTIVe Technology and the Epipaleolithic 75 tests (Units B and C) were excavated in 1992 (Neeley et al. 1998), five 1 m× 1 m tests (Squares B1, C1, C2, C3, and C4) were dug in 2000 (Olszewski et al. 2001), and a further 11 units (I98, I99, J98, J99, L97, L98, L99, M97, M98, M99, and S97) were excavated in Area A in 2012 (Olszewski and al-Nahar 2014). The sequence containsE arly Epipaleolithic deposits in the A and B areas of the site, as well as in the lower deposits in the C area. The upper deposits in the C area have been attributed to the Middle Epipaleolithic (e.g., Neeley et al. 1998; Olszewski 2000). A portion of the excavations at Tor at-Tareeq reached bedrock (Step A, all of Step B except the 1 m adjacent to Step C, the northwestern corner of Unit B, Square B1, and all the units in Area A in the 2012 field season); in those tests that did not, each was excavated a meter or more in total depth. Test excavations at Tor Sageer were undertaken in 1997 and 1998 (Coinman et al. 1999; Olszewski et al. 1998). A total of six contiguous 1 m × 1 m tests were excavated (Units B3, B4, C4, D3, D4, and E4). Both Early Epipaleolithic and Late Upper Paleolithic are present at this site (Olszewski 2016). Bedrock was reached about 75–85 cm below ground surface. Early Epipaleolithic lithic assemblages at these two sites are attributed to the Nebekian industry, which was first described by Rust at Jabrud Rockshelter 3 in Syria (Rust 1950:107–111) and later suggested for many sites in Jordan as the historically most appropriate nomenclature (Byrd 1988:263; Byrd and Garrard 2013; Goring-Morris and Belfer-Cohen 1997:76; Olszewski 2006). The Nebekian is a bladelet-based industry characterized as containing narrow forms of nongeometric microliths (curved, double curved, pointed, backed and truncated bladelets including La Mouillah points, and rare Qalkhan points) along with the earliest use of the microburin technique in the Levant (Garrard et al. 1988:325–326; Olszewski 2003). There are rare examples of geo- metric forms, principally trapezes with one or both ends showing microburin scars. Units used for analysis here include Stratum I at Tor Sageer and the 2000 season tests (Squares B1 and C1–C4) at Tor at-Tareeq (see table 4.1). Both sites have been radiocarbon dated using charcoal samples. These indi- cate that Stratum I at Tor Sageer postdates 24,630–24,150 cal BP, while Tor at-Tareeq was occupied from about 21,780–18,270 cal BP (dates have been calibrated using IntCal13, v2.2.4 from dates presented in Olszewski 2003:232; Neeley et al. 1998:303).

Smugglers’ Cave, The site of Smugglers’ Cave (Grotte des Contrebandiers) is about 17 km from Rabat, along the coastal road to Casablanca. It is situated some 270 m

76 Olszewski from the Atlantic shoreline, at an elevation of approximately 9 masl (Roche 1963:190). The cave was excavated by J. Roche (1963, 1973) in the mid-to-late 1950s and in the 1960s and 1970s (Roche and Texier 1976), as well as in 1994 by Bouzouggar (1997). Epipaleolithic materials were recovered from excava- tions between 1955 and 1957; this occupation was reported to be confined to an area near the mouth of the cave covering roughly 8 m × 2.75 m and attain- ing a maximum depth of about 0.75 m (Roche 1963:190–191). Four seasons of new excavations (2007–2010) have recently been undertaken at the cave; in 2008–2010, these included investigations into the Epipaleolithic occupation (in which the author was involved). The Epipaleolithic in this region is classified as the Iberomaurusian indus- try (Camps 1974; Garcea and Giraudi 2006; Pallary 1909; Roche 1963; Tixier 1963:10), although some researchers refer to this industry as Upper Paleolithic (e.g., Barton et al. 2005). The author restudied the Roche collections from the 1950s, which form the basis for the materials reported here (see table 4.1; also Olszewski et al. 2011). The Iberomaurusian emphasizes bladelet manufacture for the production of microliths, but it also contains a significant flake produc- tion component. Microliths are mainly nongeometric forms (pointed, curved, and blunt-ended bladelets, as well as Ain Keda points), although there are a small number of geometrics (mainly lunates). Microburin technique is present. The industry also is characterized by a significant number of pièces esquillées (“wedges”), many of which are exceptionally small (such pieces are also known from other Later Stone Age lithic assemblages in other parts of Africa, where they are often called scaled pieces or outils écaillés (e.g., Close 1989; Kusimba 2001; Willoughby 2001). These might have been tools, but it is equally likely that they are cores for the production of minuscule flakes (< 2 cm). There are no reliable radiocarbon dates for Smugglers’ Cave; however, deposits with analogous Iberomaurusian materials from Kehf al Hammar, , and Ghar Cahar in Morocco have provided a series of radiocarbon dates between 21,000 and 11,000 cal BP (Barton et al. 2005:92; Bouzouggar et al. 2008:8).

Ovv er iew of the Lithic Assemblages Although the assemblages from the five contexts at the four sites vary widely in geographic and temporal distribution, all are Epipaleolithic lithic industries emphasizing nongeometric microliths (see table 4.1), which, as noted above, allows this set of forms to be examined for behavioral implications within the sequence of “types” present in each assemblage rather than just noting its occurrence using typological classification. The four sites also share ubiquitous

Re dUCTIVe Technology and the Epipaleolithic 77 types such as notch/denticulates and retouched pieces; these two categories often can be proxies for incidental edge damage (e.g., trampling), particu- larly when the retouch is characterized by “nibbling” and/or inconsistent steep, alternate removals and the notches are undistinguished (i.e., not regularly produced as in denticulated edges characteristic of some sickle types or saws). For the most part, then, notch/denticulates and retouched pieces at all the site contexts except Smugglers’ Cave are of little diagnostic value. Smugglers’ Cave, however, has few notch-denticulates and the retouched pieces mainly are characterized by consistent, regular retouch, indicating that lithic materi- als suffered little in the way of taphonomic damage. Because nongeometric microliths typify these assemblages, and each also contains evidence for the use of microburin technique to snap bladelets that can then be backed or retouched into microliths, the next step is to exam- ine what forms of nongeometrics are present in each of these Epipaleolithic assemblages. As shown in table 4.2, there are several “named” types (Dufour, Qalkhan, Ain Keda, and Ouchtata) that tend to be regionally specific, although not always (e.g., Ouchtata). Setting aside nondiagnostic types such as blunt, inverse, partially backed, double backed and other, table 4.3 indicates that nongeometrics mainly are distributed between three categories—pointed types, curved types, and types with truncations/microburin scars. Pointed nongeometrics range from 15.2 percent to 35.9 percent; curved types from 7.1 percent to 30.7 percent; and, the truncated group from 9.3 percent to 40.8 percent (shown regrouped in table 4.4). Both Warwasi Unit 2 and Smugglers’ Cave have low frequencies of curved types, which is interesting given that Ain Keda points (present at Smugglers’ Cave) have a curved edge. The explanation, however, may be related to the fact that the curved edge on Ain Keda points is not the backed edge, but the nonretouched edge. This differs from the classic curved backed bladelet type in which the curved edge is the backed edge. All of the sites except Smugglers’ Cave show moderate to high frequencies of the truncated group. Calculations of the restricted microburin index (Imbtr), following Goring- Morris (1987:50), use the total number of microburins divided by the total number of microburins plus nongeometrics plus geometrics × 100 (exclud- ing fragments, but including truncated bladelets). These calculations result in the following values: Imbtrs—Warwasi Unit 2: 6.9; Warwasi Unit 4: 7.3; Tor Sageer: 23.9; Tor at-Tareeq: 30.2; and Smugglers’ Cave: 12.6. At this level of analysis, it is clear that habitual use of microburin technique is less common at Warwasi and at Smugglers’ Cave than at the Jordanian sites of Tor Sageer and Tor at-Tareeq.

78 Olszewski Table 4.2. Nongeometric microliths. Smugglers’ Warwasi 2 Warwasi 4 Tor Sageer Tor at-Tareeq Cave Type %N%N%N%N%N Dufour 8.9 22 9.8 44 2.4 14 0.2 1 — — Qalkhan — — — — 1.2 7 0.3 2 — — Ain Keda — — — — — — — — 5.7 22 Ouchtata — — — — 6.3 37 0.5 3 6.8 26 pointed 28.6 71 15.8 71 8.2 48 11.6 69 11.2 43 double — — — — 4.3 25 0.8 5 — — pointed curved 5.6 14 13.3 60 15.2 89 13.8 82 9.6 37 La — — — — 6.0 35 2.9 17 0.5 2 Mouillah backed 4.4 11 8.4 38 17.2 101 7.4 44 1.8 7 and truncated truncated 10.1 25 8.7 39 4.9 29 10.6 63 4.2 16 blunt 5.2 13 2.0 9 0.5 3 2.0 12 8.6 33 inverse — — 0.4 2 3.4 20 1.7 10 0.5 2 double 5.6 14 7.1 32 2.4 14 1.3 8 3.4 13 backed partially 9.3 23 8.4 38 8.9 52 5.7 34 11.5 44 backed other 2.0 5 6.9 31 1.5 9 1.5 9 6.0 23 fragment 20.2 50 19.1 86 17.6 103 39.6 235 30.2 116 Total (248) (450) (586) (594) (384)

Microliths and Microburins Because microburin technique is widely interpreted as a method used in the manufacture of microliths, it is possible to ask if certain microlith forms might represent “stages” in microlith manufacture. This question does not assume or predict that “finished” types (always a semantically loaded term), such as trapezes or lunates or backed and truncated bladelets, are modified from one “finished” form into another (e.g., as assumed by some to be an outcome predicted by Neeley and Barton 1994). Rather, the question here is if there are microlith forms that are linked in a series of manufacturing

Re dUCTIVe Technology and the Epipaleolithic 79 Table 4.3. Nongeometric microliths, excluding fragments. Smugglers’ Warwasi 2 Warwasi 4 Tor Sageer Tor at-Tareeq Cave Type %N%N%N%N%N Dufour 11.1 22 12.1 44 2.9 14 0.3 1 — — Qalkhan — — — — 1.4 7 0.6 2 — — Ain Keda — — — — — — — — 8.2 22 Ouchtata — — — — 7.7 37 0.8 3 9.7 26 pointed 35.9 71 19.5 71 9.9 48 19.2 69 16.0 43 double — — — — 5.2 25 1.4 5 — — pointed curved 7.1 14 16.5 60 18.4 89 22.8 82 13.8 37 La — — — — 7.2 35 4.7 17 0.7 2 Mouillah backed 5.6 11 10.4 38 20.9 101 12.3 44 2.6 7 and truncated truncated 12.6 25 10.7 39 6.0 29 17.5 63 6.0 16 blunt 6.6 13 2.5 9 0.6 3 3.3 12 12.3 33 inverse — — 0.5 2 4.1 20 2.8 10 0.7 2 double 7.1 14 8.8 32 2.9 14 2.2 8 4.8 13 backed partially 11.6 23 10.4 38 10.8 52 9.4 34 16.4 44 backed other 9.6 5 17.3 31 4.8 9 4.7 9 13.4 23 Total (198) (364) (483) (359) (268)

stages (or reduction sequence). The clearest example of this appears to be two of the forms in the truncated group—La Mouillah points and backed and truncated bladelets. As defined by Tixier (1963:106), a La Mouillah point is a bladelet with abrupt backing along one edge that terminates in a distal or proximal piquant trièdre (which is a type of microburin scar). Because of this definition, a bladelet that has an unbacked lateral edge with a truncated end cannot be part of a sequence of microlith types involving La Mouillah points because the truncation (while modifying an end of a bladelet) is not a microburin scar. Step 1 in this sequence is to modify a bladelet in order to snap it using microburin technique. There are three modification choices. The unmodified

80 Olszewski Table 4.4. Grouped nongeometric microliths. Smugglers’ Warwasi 2 Warwasi 4 Tor Sageer Tor at-Tareeq Cave Type %N%N%N%N%N Dufour 11.1 22 12.1 44 2.9 14 0.3 1 — — pointed 35.9 71 19.5 71 24.2 117 22.0 79 33.9 91 types curved 7.1 14 16.5 60 18.4 89 22.8 82 13.8 37 types trun- 18.2 36 21.1 77 34.1 165 34.5 124 9.3 25 cated types blunt 6.6 13 2.5 9 0.6 3 3.3 12 12.3 33 inverse — — 0.5 2 4.1 20 2.8 10 0.7 2 double 7.1 14 8.8 32 2.9 14 2.2 8 4.8 13 backed partially 11.6 23 10.4 38 10.8 52 9.4 34 16.4 44 backed other 9.6 5 17.3 31 4.8 9 4.7 9 13.4 23 Total (198) (364) (483) (359) (268) bladelet can be abruptly backed along most of one lateral edge, ending in a pseudo-notch, and then snapped at the pseudo-notch; the bladelet can be notched and snapped at the notch; or, the bladelet can be laterally backed and notched, and then snapped at the notch (Olszewski 2001a:312). It is the pseudo-notch technique that Tixier (1963:108) describes for generating a La Mouillah point (figure 4.2). The resulting microburins, regardless of which option of preparing the bladelet was chosen, would look identical to each other as all would have a portion of what appeared to be a notch. Step 2 in the sequence is to discard the microburin while retaining the bladelet portion. If Tixier is correct in assuming that the pseudo-notch tech- nique was used, then this bladelet, which has lateral backing and a piquant trièdre microburin scar, is the La Mouillah point. Alternatively, if the bladelet was snapped using only the notch technique, it would be necessary to laterally back one edge of the bladelet in order to produce a La Mouillah point. Finally, Step 3 in the sequence is to modify the laterally backed bladelet (the La Mouillah point) by truncating the end that has the microburin scar. The resulting nongeometric microlith is a backed and truncated bladelet.

Re dUCTIVe Technology and the Epipaleolithic 81 Figure 4.2. From microburin to microlith.

Data from the five site occupations in the MiddleE ast and North Africa can be examined to ascertain whether or not this microlith type’s reduction sequence is a reasonable interpretation. At first glance, the data appear equivo- cal (table 4.5). That is, backed and truncated bladelets appear in assemblages without La Mouillah points (Warwasi) and have variable frequencies in the other assemblages (Tor Sageer, Tor at-Tareeq, and Smugglers’ Cave). If, how- ever, consideration of the intensity of microburin technique is added (table 4.6), it is clear that high levels of Imbtr are much more likely to result in the presence of La Mouillah points. The presence of La Mouillah points in the site assemblages thus appears to reflect two factors. First are the effects of sample size on diversity (larger samples tend to yield a greater diversity of types); and, second, it is possible that the microburin technique was used at Warwasi not for making backed and truncated bladelets, but for the manufac- ture of geometric microliths (which are considerably more frequent in these Zarzian occupations than at the other sites used here; see table 4.1). Minute observations on the assemblage from Tor Sageer, which focus on microburin notch placement on microburins and La Mouillah points, as well as whether it is the proximal or distal end of a backed and truncated bladelet that is truncated, offer additional support for these two microlith types as part of a reduction sequence. As shown in figure 4.3a, the majority (69.6%)

82 Olszewski Table 4.5. La Mouillah points and backed and truncated bladelets (within microliths). Smugglers’ Warwasi 2 Warwasi 4 Tor Sageer Tor at-Tareeq Cave Type %N%N%N%N%N La — — — — 7.2 35 4.7 17 0.7 2 Mouillah backed & 5.6 11 10.4 38 20.9 101 12.3 44 2.6 7 truncated

Table 4.6. La Mouillah points and backed and truncated bladelets (within microliths) with Imbtr. Smugglers’ Warwasi 2 Warwasi 4 Tor Sageer Tor at-Tareeq Cave Type %N%N%N%N%N La — — — — 7.2 35 4.7 17 0.7 2 Mouillah backed & 5.6 11 10.4 38 20.9 101 12.3 44 2.6 7 truncated Imbtr 6.9 7.3 22.0 30.2 12.6

of the microburins are proximal pieces. This means that the resulting snapped bladelet most often will have a microburin scar on the proximal end of the piece. When La Mouillah points are examined for microburin scar placement (figure 4.3b), the majority (68.8%) have a proximal piquant trièdre microburin scar. If La Mouillah points and backed and truncated bladelets represent two elements of a sequence, the backed and truncated elements should have proxi- mal truncations, and, indeed, this is the case in the Tor Sageer assemblage (figure 4.3c; 75%). Based on the data presented above, it is reasonable to conclude that, in Epipaleolithic assemblages reflecting primarily nongeometric microlith pro- duction where microburin technique is used, La Mouillah points represent an intermediate “stage” in the sequence from snapping a bladelet to manufactur- ing a backed and truncated bladelet. It is somewhat ironic that a manufac- turing stage has been given a special type name, which semantically implies that the piece itself was a desired end product (and thus perhaps supporting viewpoints that see lithic typologies as problematic; e.g., Bisson 2000; Chase, chapter 13, this volume; Clark 2002; Monnier 2006).

Re dUCTIVe Technology and the Epipaleolithic 83 Figure 4.3. Tor Sageer frequencies of (A) proximal or distal microbu- rins, (B) proximal or distal microburin scars on La Mouillah points, and (C) proximal or distal truncations on backed and truncated bladelets. Figure 4.4. Piece ésquillée artifact from Smugglers’ Cave.

Pièces Esquillées (“Splintered Pieces”) The site of Smugglers’ Cave in Morocco provides another example of a tool type that quite likely is not a tool, but that does reflect intensive use of stone raw materials, and that has been singled out as a marker for the Iberomaurusian Epipaleolithic tradition (Camps 1974:64). These are pièces esquillées—small artifacts with opposing edges that appear “splintered” on one or both surfaces (figure 4.4). Pièces esquillées have been variously interpreted as bipolar wedges used to split a grooved organic material such as wood, bone, or antler (e.g., Bardon et al. 1906:14–15; Chauchat et al. 1985; Tixier 1963:146) or as bipolar cores (e.g., Escalon de Fonton 1969). In the tool (wedge) interpretation, the bipolar splintering is explained as a byproduct of shock waves when hitting the wedge to force the groove in the organic material to split. Alternatively, in the core interpretation, pièces esquillées (which are usually flakes or blades)

Re dUCTIVe Technology and the Epipaleolithic 85 are assumed to be placed against an anvil and then struck in order to remove small flakes from opposing ends of the piece. Such small flakes are presumably a desired end product. Using data from Smugglers’ Cave, it is possible to examine the pièce esquil- lée as core interpretation. This “tool” type (which is included in the Special Tools [SP] class in table 4.1) constitutes 18.6 percent of all Iberomaurusian tools at the site. Their metric dimensions can be compared against other core types, as seen in table 4.7 (see also Olszewski et al. 2011). In a progression of potential core use and reduction strategies, the single platform cores are larg- est, followed by opposed platform cores (which presumably represent further reduction intensity), and then pièces esquillées (which on a continuum of use are the most reduced cores). In one sense, then, the main difference between small cores and pièces esquillées is the nature of the support—cores are on nod- ules while pièces esquillées are on flakes or blades. While it could be argued that pièces esquillées are smaller than single or opposed platform cores because the flakes or blades that are used for pièces esquillées would have been struck from such cores, it is also true that their length as a flake or blade is one that would have necessitated larger core sizes than those now represented in the single and opposed core sample. In the larger context, it is apparent that there are many small flakes and bladelets at the site, perhaps suggesting an emphasis on their manufacture. For example, there are 60 cores-on-flakes (27.2 mm length, 18.7 mm width, 8.3 mm thickness, and 6.5 g weight), from which only small flake removals can be made. They are similar in size to the single and opposed platform cores. In addition, the single and opposed platform cores in their current form only produced flakes or bladelets with a maximum length of 27 mm–32 mm, a size that falls within the range of microlith dimensions. Many of the final remov- als from these cores, of course, are quite a bit smaller. With this in mind, the small flakes removed from pièces esquillées fit the pattern of production of small flakes in the Iberomaurusian. While the function of such small flakes is not known, they are a widely acknowledged deliberate end product in a variety of geographical areas and temporal periods (e.g., Barrera and Kirch 1973; Dibble and McPherron 2006; Shackley 2005; Steenhuyse 2007).

Discussion This chapter presents two detailed examinations of lithic reduction processes and named tool types from Epipaleolithic contexts in the Middle East and North Africa. Such named tool types often are the basis for arguing that lithics

86 Olszewski Table 4.7. Pièces esquillées compared to cores at Smugglers’ Cave, Morocco. Single Platform Cores Opposed Platform Cores Pièces esquillées Length 32.7 mm 27.2 mm 18.7 mm Width 25.7 mm 19.0 mm 12.7 mm Thickness 15.9 mm 13.2 mm 4.6 mm Weight 22.4 g 10.1 g 2.3 g N 83 88 299

can help identify prehistoric cultural groups, primarily because such tool types have been interpreted as desired end products that result from cultural selec- tion (or traditions of making tools) that are specific to group identities. This is the long-held perspective of phylogeny or history-based approaches (e.g., Bar-Yosef 1991:375; see also Pirie 2004:699–700), but as Donaldson (1991:352) points out, there is an enormous amount of variability in the Epipaleolithic— even within small regions—and this variability can no longer be treated as if it is the result of simply time or ethnicity. In fact, maintaining a strong link- age between ethnicity and lithic types is to some degree counterproductive, because it imposes a straitjacket of “immutable” lithic types on interpretations of prehistoric behaviors (e.g., Bisson 2000). Most researchers (Epipaleolithic or other archaeologists), even those favor- ing strong ties between lithic types and ethnic groups, certainly recognize the host of other variables that influence lithic assemblages, such as access to stone raw material, abundance of stone raw material, relative level of mobility, and features of subsistence activities (e.g., commentary in Goring-Morris et al. 1996; Kaufman 1995). What is less often appreciated, particularly in Old World con- texts, is that lithic types are “works in progress,” that is, just as the reduction of a core can be studied as part of a chaîne opératoire or of a reduction sequence, so too can the production of tools, as in the example of microburins, La Mouillah points, and backed and truncated bladelets discussed above. Naturally, such sequences of reduction cannot, and should not, be applied indiscriminately to lithic assemblages that happen to contain a particular type or types, as can be seen from the several Middle Eastern and North African examples used above. That is to say, we must avoid the trap of applying a generic static explanation for lithic reduction sequences that ends up being analogous to the static model of immutable lithic types (see also Chase, chapter 13, this volume). Just as settlement decisions and subsistence choices are dynamic processes, so too are those choices that affect the production and reduction of lithic

Re dUCTIVe Technology and the Epipaleolithic 87 artifacts (e.g., see Shea 2013; Shott et al. 2011). All of these choices are, most importantly, situational. Thus, for example, coastal populations of the entity identified as Iberomaurusian appear to have maximized the use of high-­ quality flint while occupying their sites. As discussed above, this is reflected not only in the small cores (which produce small flakes and blades/bladelets), but also in the use of flakes as cores (the well-recognized category of cores- on-flakes), and in the use of flakes and blades as bipolar cores (traditionally categorized as the “tool” type of pièces esquillées). If we were to find these same groups in noncoastal areas and/or in situations with other types of stone raw material access and abundance, would we find them continuing to maximize lithic reduction in these ways? Or would we not even recognize them as Iberomaurusian because we would lack one of their defining lithic features? If many lithic types reflect contextual responses, then we are not tracing groups in prehistory, but rather the application of strategies as circumstances neces- sitate (e.g., Lombard 2012:147–148). Lithic reduction is a powerful explanatory mechanism ( Jelinek 1976). And, in many cases, it is the most parsimonious of the available alternatives in behavioral interpretations. Beyond this, however, rather than limiting the human capacity for innovation and creativity, as static explanations of lithic types often do (because once invented, an immutable type must spread to other groups from the region of invention), models based on lithic reduction processes allow for the re-creation, reinvention, and rediscovery of particular lithic types and lithic strategies as the contingencies of a situation warranted (i.e., examples of technological convergence should be expected). In other words, examining lithic reduction processes and choices speaks to behavioral adaptations of prehistoric groups in a far more direct way than it ever did to a historical trajectory.

Acknowledgments Field research in Jordan was made possible by grants from the National Science Foundation (SBR–9618766), the Wenner Gren Foundation (GR 6278), and the National Geographic Society (6695–00); analysis of the Iberomaurusian collections in Morocco by funding from the University Research Foundation, University of Pennsylvania (to Harold L. Dibble); and analysis of the Epipaleolithic materials from Warwasi by the American Philosophical Society.

88 Olszewski References Adams, William Y., and Ernest W. Adams. 1991. Archaeological Typology and Practical Reality. New York: Cambridge University Press. http://dx.doi.org/10.1017/CBO 9780511558207. Bardon, L., J. Bouyssonie, and A. Bouyssonie. 1906. “Outils Écaillées par Percussion.” Revue de l’école d’Anthropologie XVI:170–175. Barrera, William M., Jr., and Patrick Vinton Kirch. 1973. “Basaltic-Glass Artefacts from Hawaii: Their Dating and Prehistoric Uses.” Journal of the Polynesian Society (N. Z.) 82:177–187. Barton, R.N.E., A. Bouzouggar, S. N. Collcutt, R. Gale, T.F.G. Higham, L. T. Humphrey, S. Parfitt, E. Rhodes, C. B. Stringer, and F. Malek. 2005. “The Late Upper Palaeolithic Occupation of the Moroccan Northwest Maghreb during the Last Glacial Maximum.” African Archaeological Review 22(2):77–100. http://dx.doi .org/10.1007/s10437-005-4190-y. Barton, C. Michael, Deborah I. Olszewski, and Nancy R. Coinman. 1996. “Beyond the Graver: Reconsidering Burin Function.” Journal of Field Archaeology 23(1):111– 125. http://dx.doi.org/10.2307/530611. Bar-Yosef, Ofer. 1991. “Stone Tools and Social Context in Levantine Prehistory.” In Perspectives on the Past: Theoretical Biases in Mediterranean Hunter-Gatherer Research, ed. Geoffrey A. Clark, 371–395. Philadelphia: University of Pennsylvania Press. Bar-Yosef, Ofer, and Philip Van Peer. 2009. “The Chaîne Opératoire Approach in Middle Paleolithic Archaeology.” Current Anthropology 50(1):103–131. http://dx.doi .org/10.1086/592234. Becker, Mark, and Fred Wendorf. 1993. “A Microwear Study of a Late Pleistocene Quadan Assemblage from Southern .” Journal of Field Archaeology 20:389–398. Bettinger, Robert L., James F. O’Connell, and David Hurst Thomas. 1991. “Projectile Points as Time Markers in the Great Basin.” American Anthropologist 93(1):166–172. http://dx.doi.org/10.1525/aa.1991.93.1.02a00140. Binford, Lewis R., and Sally R. Binford. 1966. “A Preliminary Analysis of Functional Variability in the Mousterian of Levallois Facies.” American Anthropologist 68(2):238–295. http://dx.doi.org/10.1525/aa.1966.68.2.02a001030. Bisson, Michael. 2000. “Nineteenth Century Tools for Twenty-First Century Archaeology? Why the Middle Paleolithic Typology of François Bordes Must Be Replaced.” Journal of Archaeological Method and Theory 7(1):1–48. http://dx.doi .org/10.1023/A:1009578011590. Bleed, Peter. 2001. “Trees or Chains, Links or Branches: Conceptual Alternatives for Consideration of Stone Tool Production and Other Sequential Activities.” Journal

Re dUCTIVe Technology and the Epipaleolithic 89 of Archaeological Method and Theory 8(1):101–127. http://dx.doi.org/10.1023 /A:1009526016167. Bordes, François. 1961. “Mousterian Cultures in France.” Science 134(3482):803–810. http://dx.doi.org/10.1126/science.134.3482.803. Bordes, François. 1979. Typologie du Paléolithique ancien et moyen. Cahiers du Quaternaire 1. Paris: Editions du Centre National de la Recherche Scientifique. Bourguignon, Laurence, J.-P. Faivre, and Alain Turq. 2004. “Ramification des Chaînes Opératoires: Une Spécificité du Moustérien?” Paléo 16:37–48. Bouzouggar, Abdeljalil. 1997. “Économie des Matières Premières et du Débitage dans la Séquence Atérienne de la Grotte d’El Mnasra I (Ancienne grotte des Contrebandiers—Maroc).” Préhistoire Anthropologie Méditerranéennes 6:35–52. Bouzouggar, Abdeljalil, R.N.E. Barton, S. Blockley, C. Bronk-Ramsey, S. N. Collcutt, R. Gale, T.F.G. Higham, L. T. Humphrey, S. Parfitt, E. Turner, et al. 2008. “Reevaluating the Age of the Iberomaurusian in Morocco.” African Archaeological Review 25(1-2):3–19. http://dx.doi.org/10.1007/s10437-008-9023-3. Braidwood, Robert J., and Bruce Howe. 1960. Prehistoric Investigations in Iraqi Kurdistan. Studies in Ancient Oriental Civilization No. 31. Chicago: University of Chicago Press. Braidwood, Robert J., Bruce Howe, and Charles A. Reed. 1961. “The Iranian Prehistoric Project.” Science 133(3469):2008–2010. http://dx.doi.org/10.1126/science .133.3469.2008. Byrd, Brian F. 1988. “Late Pleistocene Settlement Diversity in the Azraq Basin.” Paléorient 14(2):257–264. http://dx.doi.org/10.3406/paleo.1988.4472. Byrd, Brian F., and Andrew N. Garrard. 2013. “Regional Patterns in Late Palaeolithic Chipped Stone Production and Technology in the Levant.” In Beyond the Fertile Crescent: Late Palaeolithic and Neolithic Communities of the Jordanian Steppe, The Azraq Basin Project, ed. Andrew N. Garrard and Brian F. Byrd, vol. 1., 350–393. Oakville, CT: Oxbow Books. Camps, Gabriel. 1974. Le Civilisations Préhistoriques de l’Afrique du Nord et du Sahara. Paris: Centre National de la Recherche Scientifique. Carr, Philip J., and Andrew P. Bradbury. 2011. ““Learning from Lithics: A Perspective on the Foundation and Future of the Organization of Technology.” Special Issue: Reduction Sequence, Chaîne Opératoire, and Other Methods: The Epistemologies of Different Approaches to Lithic Analysis.” PaleoAnthropology 2011:305–319. Chauchat, Cl., C. Normand, J.-P. Raynal, and R. Santa-Maria. 1985. “Le Retour de la Pièce Esquillée.” Bulletin de la Societé Préhistorique Française 82:35–41. Clark, Geoffrey A. 1993. “Paradigms in Science and Archaeology.” Journal of Archaeological Research 1(3):203–234. http://dx.doi.org/10.1007/BF01326535.

90 Olszewski Clark, Geoffrey A. 2002. “Observations on Paradigmatic Bias in French and American Paleolithic Archaeology.” In The Role of American Archeologists in the Study of the European Upper Paleolithic, ed. Lawrence Straus, 19–26. BAR International Series 1048. Oxford: British Archaeological Reports. Clark, Geoffrey A., John Lindly, Marcia Donaldson, Andrew Garrard, Nancy Coinman, Joseph Schuldenrein, Suzanne Fish, and Deborah Olszewski. 1988. “Excavations at Middle, Upper and Epipaleolithic Sites in the Wadi 'Hasa, West-Central Jordan.” In The Prehistory of Jordan in 1986, ed. Andrew Garrard and Hans-Georg Gebel, 209–285. BAR International Series 396. Oxford: British Archaeological Reports. Clark, J. Desmond, and Maxine R. Kleindienst. 1974. “The Stone Age Cultural Sequence: Terminology, Typology and Raw Material.” In Kalambo Falls Prehistoric Site, vol. 2. ed. J. Desmond Clark, 71–106. London: Cambridge University Press. Close, Angela E. 1989. “Report on Site E-81-1: An Early Kubbaniyan Site on the Edge of the Kubbaniyan Dune Field.” In The Prehistory of Wadi Kubbaniya, Volume 3. Late Paleolithic Archaeology, ed. Fred Wendorf, Romuald Schild, and Angela E. Close, 490–523. Dallas: Southern Methodist University Press. Coinman, Nancy R., Deborah I. Olszewski, Kholood Abdo, Todd G. Clausen, Jason B. Cooper, Jason Fox, Maysoon al-Nahar, Edward Richey, and Leslie Saele. 1999. “The Eastern Hasa Late Pleistocene Project: Preliminary Report on the 1998 Season.” Annual of the Department of Antiquities of Jordan 43:9–25. Debénath, André, and Harold L. Dibble. 1994. Handbook of Paleolithic Typology. Volume One: Lower and Middle Paleolithic of Europe. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Demars, P.-Yves, and Pierre Laurent. 1992. Types d’Outils Lithiques du Paléolithique Supérieure en Europe. Paris: Presses du Centre National de la Recherche Scientifique. de Sonneville-Bordes, Denise, and Jean Perrot. 1954. “Lexique Typologique du Paléolithique Supérieur. I. Grattoirs—II. Outils Solutréens.” Bulletin de la Société Préhistorique Française 51(7):327–335. http://dx.doi.org/10.3406/bspf.1954.3106. de Sonneville-Bordes, Denise, and Jean Perrot. 1955. “Lexique Typologique du Paléolithique Supérieur. III. Outils composites—Perçoirs.” Bulletin de la Société Préhistorique Française 52(1/2):76–79. http://dx.doi.org/10.3406/bspf.1955.3156. de Sonneville-Bordes, Denise, and Jean Perrot. 1956. “Lexique Typologique du Paléolithique Supérieur. IV. Burins—V. Outillage à Bord Abattu—VI. Pièces Tronquées—VII. Lames Retouchées—VIII. Pièces Variées—IX. Outillage Lamellaire. Pointe Azilienne.” Bulletin de la Société Préhistorique Française 53(7):408–412. http://dx.doi.org/10.3406/bspf.1956.3357.

Re dUCTIVe Technology and the Epipaleolithic 91 Dibble, Harold L. 1987. “The Interpretation of Middle Paleolithic Scraper Morphology.” American Antiquity 52(1):109–117. http://dx.doi.org/10.2307/281062. Dibble, Harold L. 1995. “Middle Paleolithic Scraper Reduction: Background, Clarification, and Review of the Evidence to Date.” Journal of Archaeological Method and Theory 2(4):299–368. http://dx.doi.org/10.1007/BF02229003. Dibble, Harold L., and Simon J. Holdaway. 1993. “The Middle Paleolithic Industries of Warwasi.” In The Paleolithic Prehistory of the Zagros-Taurus, ed. Deborah I. Olszewski and Harold L. Dibble, 75–99. University Museum Symposium Series. Philadelphia: University of Pennsylvania Museum. Dibble, Harold L., and Shannon P. McPherron. 2006. “The Missing Mousterian.” Current Anthropology 47(5):777–803. http://dx.doi.org/10.1086/506282. Dibble, Harold L., Shannon P. McPherron, Philip Chase, William R. Farrand, and André Debénath. 2006. “Taphonomy and the Concept of Paleolithic Cultures: The Case of the Tayacian from Fontéchevade.” PaleoAnthropology 2006:1–21. Donaldson, Marcia L. 1991. “Historic Biases in Modern Perceptions of the Levantine Epipaleolithic.” In Perspectives on the Past: Theoretical Biases in Mediterranean Hunter-Gatherer Research, ed. Geoffrey A. Clark, 341–352. Philadelphia: University of Pennsylvania Press. Escalon de Fonton, M. 1969. “La Pièce Esquillée—Essai d’Interpretation.” Bulletin de la Societé Préhistorique Française 66(3):76. http://dx.doi.org/10.3406 /bspf.1969.10365. Flenniken, J. Jeffrey, and Philip J. Wilke. 1989. “Typology, Technology, and Chronology of Great Basin Dart Points.” American Anthropologist 91(1):149–158. http://dx.doi.org/10.1525/aa.1989.91.1.02a00100. Frison, George. 1968. “A Functional Analysis of Certain Chipped Stone Tools.” American Antiquity 3(2):149–155. http://dx.doi.org/10.2307/278516. Garcea, Elena A. A., and Carlo Giraudi. 2006. “Late Quaternary Human Settlement Patterning in the Jebel Gharbi.” Journal of Human Evolution 51(4):411–421. http://dx.doi.org/10.1016/j.jhevol.2006.05.002. Garrard, Andrew N., Alison V. G. Betts, Brian F. Byrd, Susan Colledge, and Christopher Hunt. 1988. “Summary of Palaeoenvironmental and Prehistoric Investigations in the Azraq Basin.” In The Prehistory of Jordan: The State of Research in 1986, ed. Andrew N. Garrard and Hans-Georg Gebel, 311–337. BAR International Series 396. Oxford: British Archaeological Reports. Garrod, Dorothy A. E. 1930. “The Palaeolithic of Southern Kurdistan: Excavations in the Caves of Zarzi and Hazar Merd.” Bulletin of the American School of Prehistoric Research 6:8–43.

92 Olszewski Goring-Morris, A. Nigel. 1987. At the Edge: Terminal Pleistocene Hunter-Gatherers in the Negev and Sinai. BAR International Series 361. 2 vols. Oxford: British Archaeological Reports. Goring-Morris, A. Nigel, and Anna Belfer-Cohen. 1997. “The Articulation of Cultural Processes and Late Quaternary Environmental Changes in Cisjordan.” Paléorient 23(2):71–93. http://dx.doi.org/10.3406/paleo.1997.4653. Goring-Morris, A. Nigel, Donald O. Henry, James L. Phillips, Geoffrey A. Clark, C. Michael Barton, and Michael P. Neely. 1996. “Pattern in the Epipalaeolithic of the Levant: Debate after Neeley and Barton.” Antiquity 70:130–137. Hours, Francis. 1974. “Remarques sur l’Utilisation des Listes-Types pour l’Étude du Paléolithique Supérieur et de l’Épipaléolithique du Levant.” Paléorient 2(1):3–18. http://dx.doi.org/10.3406/paleo.1974.4172. Jelinek, Arthur J. 1976. “Form, Function, and Style in Lithic Analysis.” In Cultural Change and Continuity: Essays in Honor of James Bennett Griffin, ed. C. E. Cleland, 19–33. New York: Academic Press. Kaufman, Daniel. 1995. “Microburins and Microliths of the Levantine Epipaleolithic: A Comment on the Paper by Neeley and Barton.” Antiquity 69(263):375–381. http://dx.doi.org/10.1017/S0003598X00064784. Knecht, Heidi. 1988. Upper Paleolithic Burins: Type, Form, and Function. BAR International Series 434. Oxford: British Archaeological Reports. Kusimba, Sibel Barut. 2001. “The Early Later Stone Age in East Africa: Excavations and Lithic Assemblages from Lukenya Hill.” African Archaeological Review 18(2):77–123. http://dx.doi.org/10.1023/A:1011032025300. Leakey, Louis S. B. 1951. Olduvai Gorge. Cambridge: Cambridge University Press. Leakey, Mary D. 1966. “A Review of the Oldowan Culture from Olduvai Gorge, .” Nature 210(5035):462–466. http://dx.doi.org/10.1038/210462a0. Leguay, Louis. 1877. “Les Procèdes Employés pour la Gravure et la Sculpture des Os avec le Silex.” Bulletin de la Societé d’Anthropologie de Paris 12(1):280–296. http://dx .doi.org/10.3406/bmsap.1877.3234. Lindly, John M., Roman F. Beck, and Geoffrey A. Clark. 2000. “Core Reconstructions and Lithic Reduction Sequences at WHS 623X: An Upper Paleolithic Site in the Wadi al-Hasa, West-Central Jordan.” In The Archaeology of the Wadi al-Hasa, West-Central Jordan, Volume 2, Excavations at Middle, Upper, and Epipaleolithic Sites, ed. Nancy R. Coinman, 211–226. Anthropological Research Papers No. 52. Tempe: Arizona State University. Lombard, Marlize. 2012. “Thinking Through the Middle Stone Age of Sub-Saharan Africa.” Quaternary International 270:140–155. http://dx.doi.org/10.1016/j.quaint .2012.02.033.

Re dUCTIVe Technology and the Epipaleolithic 93 McPherron, Shannon P. 1999. “Ovate and Pointed Handaxe Assemblages: Two Points Make a Line.” Préhistoire Européenne 14:9–32. McPherron, Shannon P., ed. 2007. Tools Versus Cores. Alternative Approaches to Stone Tool Analysis. Newcastle: Cambridge Scholars Press. Monnier, Gilliane. 2006. “Testing Retouched Flake Tool Standardization During the Middle Paleolithic: Patterns and Implications.” In Transitions before the Transition: Evolution and Stability in the Middle Paleolithic and Middle Stone Age, ed. E. Hovers and S. Kuhn, 57–83. New York: Springer. http://dx.doi.org/10.1007/0-387-24661 -4_4. Neeley, Michael P., and C. Michael Barton. 1994. “A New Approach to Interpreting Late Pleistocene Microlith Industries in Southwest Asia.” Antiquity 68:275–288. Neeley, Michael P., Jane D. Peterson, Geoffrey A. Clark, Suzanne K. Fish, and Margaret Glass. 1998. “Investigations at Tor al-Tareeq: An Epipaleolithic Site in the Wadi el-Hasa, Jordan.” Journal of Field Archaeology 25:295–317. Olszewski, Deborah I. 1993a. “Zarzian Microliths from Warwasi Rockshelter, Iran: Scalene Triangles as Arrow Components.” In Hunting and Animal Exploitation in the Later Paleolithic and Mesolithic of Eurasia, ed. Gail Larsen-Peterkin, Harvey M. Bricker and Paul A. Mellars, 199–205. Archaeological Papers No. 4. Washington, DC: American Anthropological Association. http://dx.doi.org/10.1525/ap3a.1993.4 .1.199. Olszewski, Deborah I. 1993b. “The Zarzian Occupation of Warwasi Rockshelter, Iran.” In The Paleolithic Prehistory of the Zagros-Taurus, ed. Deborah I. Olszewski and Harold L. Dibble, 206–236. University Museum Symposium Series. Philadelphia: University of Pennsylvania Museum. Olszewski, Deborah I. 1996. “The Lithic Transition to the Early Neolithic in the Zagros Region: Zarzian and M’lefaatian Industries.” In Neolithic Chipped Stone Industries of the Fertile Crescent, ed. Stefan Kozlowski and Hans-Georg Gebel, 183–192. Studies in Early Near Eastern Production, Subsistence, and Environment. Berlin: Ex Oriente. Olszewski, Deborah I. 1999. “The Early Upper Paleolithic in the Zagros Mountains.” In Dorothy Garrod and the Progress of the Palaeolithic. Studies in the Prehistoric Archaeology of the Near East and Europe, ed. William Davies and Ruth Charles, 167–180. Oxford: Oxbow Books. Olszewski, Deborah I. 2000. “The Epipaleolithic in the Wadi al-Hasa: An Overview.” In The Archaeology of the Wadi al-Hasa, West-Central Jordan, Volume 2. Excavations at Middle, Upper, and Epipaleolithic Sites, ed. Nancy R. Coinman, 227–243. Anthropological Research Papers No. 52. Tempe: Arizona State University.

94 Olszewski Olszewski, Deborah I. 2001a. “My ‘Backed and Truncated Bladelet,’ Your ‘Point’: Terminology and Interpretation in Levantine Epipaleolithic Assemblages.” In Beyond Tools: Redefining the PPN Lithic Assemblages of the Levant, ed. Isabella Caneva, Cristina Lemorini, Daniela Zampetti, and Paolo Biagi, 303–318. Studies in Early Near Eastern Production, Subsistence, and Environment 9. Berlin: Ex Oriente. Olszewski, Deborah I. 2001b. “Ruminations on the Early Upper Paleolithic and a Consideration of the Zagros Aurignacian.” In Questioning the Answers: Re-solving Fundamental Problems of the Early Upper Paleolithic, ed. Maureen A. Hays and Paul T. Thacker, 79–89. BAR International Series 1005. Oxford: British Archaeological Reports. Olszewski, Deborah I. 2003. “The Conundrum of the Levantine Late Upper Palaeolithic and Early Epipalaeolithic: Perspectives from the Wadi al-Hasa, Jordan.” In More Than Meets the Eye: Studies in Upper Palaeolithic Diversity in the Near East, ed. Nigel Goring-Morris and Anna Belfer-Cohen, 230–241. Oxford: Oxbow Books. Olszewski, Deborah I. 2006. “Issues in the Levantine Epipaleolithic: The Madamaghan, Nebekian, and Qalkhan (Levant Epipaleolithic).” Paléorient 32(1):19–26. http://dx.doi.org/10.3406/paleo.2006.5168. Olszewski, Deborah I. 2007a. “Carinated Tools, Cores, and Mobility: The Zagros Aurignacian Example.” In Tools Versus Cores: Alternative Approaches to Stone Tool Analysis, ed. Shannon P. McPherron, 91–106. Newcastle: Cambridge Scholars Press. Olszewski, Deborah I. 2007b. “Issues in the Development of the Early Upper Paleolithic and a ‘Transitional’ Industry from the Zagros Region.” In New Approaches to the Study of Early Upper Paleolithic “Transitional” Industries in Western Eurasia: Transitions Great and Small, ed. Julien Riel-Salvatore and Geoffrey A. Clark, 131–142. British Archaeological Reports International Series 1620. Oxford: Archaeopress. Olszewski, Deborah I. 2008. “The Palaeolithic Period, Including the Epipaleolithic.” In Jordan: An Archaeological Reader, ed. Russell Adams, 35–69. Oakville, CT: Equinox Publishing. Olszewski, Deborah I. 2016. “Late Upper Paleolithic and Initial Epipaleolithic in the Marshlands: A View from Tor Sageer, Wadi al-Hasa, Jordan.” In Fresh Fields and Pastures New: Papers Presented in Honor of Andrew M.T. Moore, ed. K. Lillois and M. Chazan, 41–53. Leiden: Sidestone Press. Olszewski, Deborah I., and Maysoon al-Nahar. 2014. “The 2012 Excavations in the Area A Early Epipaleolithic at Tor at-Tareeq, Wadi al-Hasa.” Neo-Lithics 1/14:25–32.

Re dUCTIVe Technology and the Epipaleolithic 95 Olszewski, Deborah I., Nancy R. Coinman, Todd G. Clausen, Jason B. Cooper, Henrik Jansson, Maysoon al-Nahar, Utsav Schurmans, Leslie S. Saele, Amy J. Sampson, and Jason R. Thompson. 2001. “The Eastern Hasa Late Pleistocene Project: Preliminary Report on the 2000 Field Season.” Annual of the Department of Antiquities of Jordan 45:39–60. Olszewski, Deborah I., Nancy R. Coinman, Joseph Schuldenrein, Todd G. Clausen, Jason B. Cooper, Jason Fox, J. Brett Hill, Maysoon al-Nahar, and John Williams. 1998. “The Eastern Hasa Late Pleistocene Project: Preliminary Report on the 1997 Season.” Annual of the Department of Antiquities of Jordan 42:53–74. Olszewski, Deborah I., and Harold L. Dibble. 1994. “The Zagros Aurignacian.” Current Anthropology 35(1):68–75. http://dx.doi.org/10.1086/204242. Olszewski, Deborah I., and Harold L. Dibble. 2006. “To Be or Not to Be Aurignacian: The Zagros Upper Paleolithic.” In Towards a Definition of the Aurignacian, ed. Ofer Bar-Yosef and João Zilhão, 355–373. Trabalhos de Arqueologia 45. Lisbon: American School of Prehistoric Research/Instituto Português de Arqueologia. Olszewski, Deborah I., Utsav A. Schurmans, and Beverly A. Schmidt. 2011. “The Epipaleolithic (Iberomaurusian) from Grotte des Contrebandiers, Morocco.” African Archaeological Review 28(2):97–123. http://dx.doi.org/10.1007/s10437 -010-9086-9. Pallary, P. 1909. Instructions pour la Recherché Préhistorique dans le Nord-Ouest de l’Afrique. Mémoires de la Société Historique Algérienne No. 3. Algiers. Pirie, Anne. 2004. “Constructing Prehistory: Lithic Analysis in the Levantine Epipalaeolithic.” Journal of the Royal Anthropological Institute 10(3):675–703. http://dx.doi.org/10.1111/j.1467-9655.2004.00207.x. Roche, Jean. 1963. L’Épipaléolithique Marocain. Paris: Librairie Marcel Didier. Roche, Jean. 1973. “Notes sur les Activités de Chantiers (1972).” Bulletin d’Archéologie Marocaine 9:147–153. Roche, Jean, and Jean-Pierre Texier. 1976. “Découverte de Restes Humains dans un Niveau Atérien Supérieur de la Grotte des Contrebandiers, à Temara (Maroc).” Comptes Rendus de l’Académie des Sciences 282:45–47. Rust, Alfred. 1950. Die Höhlenfunde von Jabrud (Syrien). Neumünster: Karl Wachholtz Verlag. Sellet, Frédéric. 1993. “Chaîne Opératoire: The Concept and Its Applications.” Lithic Technology 8:106–122. Shackley, M. Steven. 2005. Obsidian: Geology and Archaeology in the North American Southwest. Tucson: University of Arizona Press.

96 Olszewski Shea, John J. 2008. “The Lower and Middle Paleolithic in the Middle East and Neighboring Regions.” Evolutionary Anthropology 17(5):205–207. http://dx.doi.org /10.1002/evan.20183. Shea, John J. 2013. “Lithic Modes A—I: A New Framework for Describing Global- Scale Variation in Stone Tool Technology Illustrated with Evidence from the East Mediterranean Levant.” Journal of Archaeological Method and Theory 20(1):151–186. http://dx.doi.org/10.1007/s10816-012-9128-5. Shott, Michael J., John M. Lindly, and Geoffrey A. Clark. 2011. “Continuous Modeling of Core Reduction: Lessons from Refitting Cores from WHS 623x, an Upper Paleolithic Site in Jordan.” Special Issue: Reduction Sequence, Chaîne Opératoire, and Other Methods: The Epistemologies of Different Approaches to Lithic Analysis. PaleoAnthropology 2011:320–333. Steenhuyse, Alexandre. 2007. “Continuity and Change in Lithic Technological and Economic Strategies: The Middle and Upper Paleolithic Industries of the Brive Basin (France).” PhD dissertation, University of Pennsylvania, Philadelphia. Tixier, Jacques. 1963. Typologie de l’Épipaléolithique du Maghreb. Paris: Arts et Métiers Graphiques. Tostevin, Gilbert B. 2011. “Levels of Theory and Social Practice in the Reduction Sequence and Chaîne Opératoire Methods of Lithic Analysis.” Special Issue: Reduction Sequence, Chaîne Opératoire, and Other Methods: The Epistemologies of Different Approaches to Lithic Analysis. PaleoAnthropology 2011:351–374. Toth, Nicholas. 1985. “The Oldowan Reassessed: A Close Look at Early Stone Artifacts.” Journal of Archaeological Science 12(2):101–120. http://dx.doi.org /10.1016/0305-4403(85)90056-1. Toth, Nicholas. 1987. “The First Technology.” Scientific American 256(4):112–121. http://dx.doi.org/10.1038/scientificamerican0487-112. Turnbull, Priscilla F., and Charles A. Reed. 1974. The Fauna from the Terminal Pleistocene of Palegawra Cave. A Zarzian Occupation Site in Northeastern Iraq. Fieldiana: Anthropology, Vol. 63(3). Chicago: Field Museum of Natural History. Vaughan, Patrick. 1985. “The Burin-Blow Technique: Creator or Eliminator?” Journal of Field Archaeology 12:488–496. Willoughby, Pamela R. 2001. “Middle and Later Stone Age Technology from the Lake Rukwa Rift, Southwestern Tanzania.” South African Archaeological Bulletin 56(173/174):34–45. http://dx.doi.org/10.2307/3889026.

Re dUCTIVe Technology and the Epipaleolithic 97

5

Beginning in the final decades of the twentieth century Context and Complexity archaeologists realized that, rather than treat artifacts on the Arid Margins as distinct entities with forms reflective of either style of Australia or function, a more materialist stance was needed in which form was seen as the outcome of the life-his- Assessing Human Responses to tory of the object. Form was not simply reducible to an Unpredictable Environment style or function but varied according to the contexts in which artifact assemblages were located. In addi- tion, both the nature and accessibility of raw material Simon J. Holdaway, might affect artifact form, and assemblage composition Justin I. Shiner, might be attributed to the combined effects of occu- Patricia C. Fanning, and pation duration, production technology, and tool reuse. Matthew J. Douglass Hence, function needed to be assessed comparatively and contextually at what today would be described as a landscape scale. Here we discuss a set of processes that help to explain stone artifact assemblage variability in surface assemblages from western New South Wales (NSW), Australia (figure 5.1). We make use of assemblages described and analyzed during a dozen years of research in western NSW as part of the Western New South Wales Archaeology Program (WNSWAP) (Holdaway and Fanning 2014; Holdaway et al. 2000). In addition to more than 160,000 individual artifact records, the results of geochronological research are considered with geomorphological studies, which together pro- vide the framework necessary for investigating patterns in stone artifact assemblage composition. DOI: 10.5876/9781607324942.c005

99 Figure 5.1. Western NSW, Australia, study locations: Stud Creek, Nundooka, Fowlers Gap, Peery Lake (Rutherfords Creek), Burkes Cave, and Pine Point- Langwell (Conservation, Karz). Environmental Background Australia is the land “of droughts and flooding rains” (Mackellar 1908), with fluctuations in rainfall closely linked to ENSO (El Niño-Southern Oscillation) variation. There are indications that the current system of cycles began during the mid-Holocene (Sandweiss et al. 1999), but identifying the pattern of year-to-year environmental changes that make up the current sys- tem can be a challenge archaeologically. In the arid and semiarid regions of Australia, sedimentary deposits with suitable paleoenvironmental information are relatively rare. This rarity is partly the result of a lack of suitable proxy indicators for past environments, but it is also one of matching the scale at which environmental change is studied with the scale at which the archaeological record can be resolved (Allen et al. 2008; Holdaway and Fanning 2010). Our approach is based on the analysis of multiple radiocarbon determi- nations obtained from heat-retainer hearths, together with optically stimu- lated luminescence (OSL) determinations obtained from sediments on which the hearths rest (Fanning and Holdaway 2001; Fanning et al. 2008, 2009a; Holdaway and Fanning 2014; Holdaway et al. 2002, 2005; Rhodes et al. 2009). The hearth age determinations indicate the temporal pattern of Aboriginal occupation, while the sediment chronology informs on past erosional and depositional episodes. At the Stud Creek study area, for instance, the sediment stratigraphy of valley fill deposits dating to the Holocene shows a number of unconformities. We interpreted these as evidence of past floods that eroded the valley floor existing at the time and removed any preexisting deposits (Fanning and Holdaway 2001). In the Nundooka (ND) study area, our OSL age estimates help to confirm evidence of paleofloods identified in a separate study by Jansen and Brierley (2004). Radiocarbon age estimates from hearths at this location are all more recent than the OSL estimates for the paleoflood deposits described by Jansen and Brierley (Fanning et al. 2007). This evidence suggests floods as a mechanism for the removal of artifacts belonging to prior occupations, a process that explains the relative lack of early to mid-Holocene age deposits in our study area (Fanning et al. 2009b). Furthermore, gaps in radiocarbon age determinations from heat-retainer hearths from our study areas indicate periods when hearths were not con- structed or from whence hearths have been removed from the surface. The evidence suggests that our western NSW artifact assemblages are the result of occupations that were susceptible to climatic extremes, either drought that limited the ability of Aboriginal groups to inhabit some areas, or floods that removed the surfaces adjacent to drainage systems, a change that may also

Cxonte t and Complexity on the Arid Margins of Australia 101 have affected the suitability of particular locations for occupation (Holdaway and Fanning 2014; Holdaway et al. 2010). The precision with which we can see correlations and gaps in the record in western NSW is a consequence of the large number of determinations we have acquired. However, there is evidence from other locations that such changes are not limited to the last 2,000 years. For instance, a number of archaeolo- gists have argued for marked environmental changes beginning in the mid-­ Holocene and leading to a period of greater unpredictability in resource avail- ability (e.g., Hiscock 2003; Veth 1993). Both our studies and those of others suggest that, for humans, the environment of arid and semiarid Australian environment was both relatively depauperate in resources and unpredictable in the short and long term (Holdaway, Fanning, and Douglass 2013; Holdaway et al. 2015). We argue that this environmental context has had a marked effect on the nature of occupation in the past. Evidence from the stone artifact assem- blages supports this argument and comprises the remainder of this chapter.

Stone Artifacts Western NSW stone artifact assemblages are dominated by flakes and cores, with a smaller ground-stone component (the remains of seed grind- ers, and rare ground axe heads) (figure 5.2). Retouched flakes are uncommon, averaging less than 5 percent of all artifacts within assemblages (depend- ing on the method of calculation). Hammerstones are an occasional find. Morphological blades are rare, as are cores that suggest formal reduction technologies. Cores are predominantly worked unifacially (from a single direction) or bifacially (from two directions across a single edge) and occa- sionally the bipolar technique was used to work quartz. The majority of tool forms consist of scrapers (one or two margins of continuous retouch) and notched flakes (either single or multiple notches). Specialized tool forms are rare and restricted to the tula adze (a hafted flake usually of silcrete that is resharpened on the distal end until it has almost reached the platform, giving the appearance of a slug), geometric microliths, and Pirri points (a unifacially flaked point) (figure 5.3). Tools are generally lightly retouched, with intensive retouch only consistently evident on tula adzes (Holdaway et al. 2004, 2008a; Shiner et al. 2005, 2007). Many assemblages contain seed-grinding imple- ments, but these account for a small proportion of the overall assemblage count and are often broken. It is likely that complete seed-grinding imple- ments are underrepresented in many assemblages due to their removal by artifact collectors (Griffiths 1996).

102 Holdaway, Shiner, Fanning & Douglass Figure 5.2. Part of a grinding dish used to process seeds. Complete grinding dishes were a target for collectors and are rarely found today.

Two types of raw material dominate assemblages: silcrete and quartz. The proportion of these for selected assemblages (table 5.1) provides an indica- tion of the variability in raw material representation across western NSW. Sources are located throughout the western NSW rangelands as either out- crop or stone pavement, but are less common on the sand plains (Douglass and Holdaway 2011). The quality of quartz, largely determined by the presence of internal frac- ture planes within the individual nodules, varies greatly. As noted above, there is evidence for some bipolar reduction, but the majority of quartz cores indicate handheld reduction (e.g., unifacial and bifacial core platforms). Both quartz and silcrete artifacts often exhibit a rounded cortex characteristic of procurement from stone pavements or streambeds (figure 5.4). Quartzite and sandstone sometimes occur within assemblages, but in low propor- tions. Ground-stone implements were manufactured from these materials and quartzite was also flaked. The same types of artifact forms are manu- factured from quartz and silcrete, although the frequency and intensity of

Cxonte t and Complexity on the Arid Margins of Australia 103 Figure 5.3. Australian small tool tradition artifacts (Gould 1969): (left) Pirri point, (center) backed microliths, and (right) tula adze slug.

manufacture can vary considerably (Holdaway et al. 2004, 2008a; Holdaway and Fanning 2014).

Mobility Determining how artifacts are made allows inferences about mobility. Working artifacts in particular ways reduces the quantity of material trans- ported, or they may be carefully designed to fulfill particular tasks (Doelman and Holdaway 2011). Archaeologists have spent much time considering the relationship between artifact manufacture, transport, and form but studies have often mainly focused on retouched tools (Hiscock 1994, 2006; Nelson 1991; Torrence 1989) rather than the much more numerous core and flake com- ponent (Sullivan, chapter 6, this volume). An alternative is to consider the distribution of material that enters and leaves a location. A number of authors have discussed the difference between provisioning individuals and provisioning places with either artifacts or mate- rials (e.g., Elston 1990; Kuhn 1994). While provisioning locations makes sense

104 Holdaway, Shiner, Fanning & Douglass Table 5.1. Flaked stone artifact frequency (and proportion) by sample area and raw material. Raw Burkes Fowlers Rutherfords Stud Material Cave Conservation Creek Karz Nundooka Creek Creek 1 Silcrete 4,307 2,127 267 5,258 1,699 21,533 24,756 (67.2) (15.5) (5.4) (36.1) (41.1) (89.1) (95.6) Other 167 124 47 187 133 2,601 88,4 (2.6) (0.9) (1.0) (1.3) (3.2) (10.8) (3.4) Quartz 1,935 11,441 4,605 9,139 2,306 26 (0.1) 23,7 (30.2) (83.6) (93.6) (62.5) (55.7) (0.9) Total 6,409 13,692 4,919 14,611 4,138 24,160 25,877

Figure 5.4. Raw material sources in western NSW: creek beds and (inset) stony desert pavements. in areas where raw material occurs in only limited places and transport costs are therefore high (e.g., Webb 1993), raw material in much of western NSW is widely available, particularly if it is obtained from the gibber cobbles that

Cxonte t and Complexity on the Arid Margins of Australia 105 make up the widespread stony desert pavements (figure 5.4). Moving large quantities of raw material long distances from sources in this environment therefore makes little sense. Provisioning of places may occur when returning on a regular basis, which is to be expected where resources occur predictably, but is much less likely where resource patches are unpredictable. In western NSW, ensuring that individuals always had stone available is consistent with the need to exploit resources whenever these were encoun- tered, as well as ethnographic accounts that describe Aboriginal people car- rying stone artifacts wrapped in their hair or in leather bags (summarized in Holdaway and Douglass 2012). As discussed by Kuhn (1995), access to raw material is not solely a problem of linear distance to a source. For foragers, a more pressing concern is the time it takes to find the stone source and manu- facture an artifact needed to exploit a particular food resource. Our studies employ a three-step method, based on the presence of cortex, as a means of assessing the transport of material, whether or not an artifact is retouched or made from local or imported materials (Dibble et al. 2005; Douglass 2010; Douglass and Holdaway 2011; Douglass et al. 2008; Holdaway et al. 2008c; Lin et al. 2010). First, the total mass of an assemblage is divided by an estimate of the average mass of the individual nodules from which cores were flaked, thus giving an estimate of the number of nodules reduced. Next, estimated nodule frequency (or number) is multiplied by the total surface area of the nodules. This provides an estimate of the expected cortical surface area that should be present in an assemblage. Finally, this value is compared to the actual quantity of cortex observed in the assemblage and expressed as a ratio. Applying this method, which is a slightly modified version of Dibble et al.’s (2005), to assemblages from our western NSW study locations returns values of the cortex ratio consistently below 1.0, indicating that cortex is underrep- resented (Douglass 2010; Douglass et al. 2008). Results vary by raw material (table 5.2), with values for quartz being higher than those for silcrete, indicat- ing that the products of quartz reduction found in these assemblages are more complete than the products of silcrete reduction. The addition of many non-cortical flakes and cores to assemblages would explain the observed values of the cortex ratio. However, the local abundance of raw materials at our sampling locations (some assemblages are located directly adjacent to sources of high-quality stone) makes the scenario of extensive raw material importation seem unlikely (Douglass and Holdaway 2011). A more probable explanation for the disparity in cortex proportions is the removal of material from assemblages for use elsewhere. Cortex ratios less than 1.0 reflect a tendency toward the removal of artifacts with a greater

106 Holdaway, Shiner, Fanning & Douglass Table 5.2. Cortex ratio for selected assemblages from western NSW (adapted from Douglass 2010). Observed Cortical Expected Cortical Assemblage Material Surface Area (cm2) Surface Area (cm2) Cortex Ratio Conservation Quartz 26,567 40,295 0.66 Conservation Silcrete 1,342 5,903 0.23 Fowlers Creek Quartz 18,477 29,524 0.63 Fowlers Creek Silcrete 351 999 0.35 Karz Quartz 5,535 10,924 0.51 Karz Silcrete 732 5,569 0.13 Nundooka Quartz 8,461 12,358 0.68 Nundooka Silcrete 3,105 8,397 0.37 Stud Creek 1 Silcrete 56,073 109,191 0.51

cortical surface area to volume ratio than the nodules from which they were produced. This would result from the selective removal of large blanks that would tend to have cortex on their dorsal surface as a consequence of their size (Douglass 2010; Douglass et al. 2008).

C oping with Uncertainty Stone is heavy, but for those without the technology of metals it is essential for a host of day-to-day activities. Given this, it should be neither surprising that it was transported nor surprising that people sought the most efficient means to move it. Kuhn (1994) considered the efficiency of the shape and size of transported lithic artifacts, arguing that, while transporting cores provided great flexibility in the forms produced, transporting small flakes provided the greatest efficiency in terms of flake mass relative to useable cutting edge. The evidence from analyses of cortex strongly points to the transport of flakes both from sampling locations in the alluvial valleys and from so-called quarry locations adjacent to silcrete outcrops (Douglass 2010; Douglass et al. 2008; Holdaway et al. 2008c). People were mobile, an observation sustained by analyses of assemblages that reveal low proportions of retouched tools (par- ticularly those suggesting extended use lives). Selection of flakes for transport indicates a concern to have stone at hand. However, transporting stone in a lithic-rich landscape seems somewhat incongruous. Australian stone artifacts are characterized as expedient (Flenniken and White 1985), and producing

Cxonte t and Complexity on the Arid Margins of Australia 107 artifacts on an ad hoc basis as required might seem a more energetically effi- cient solution. A return to the paleoenvironmental interpretation discussed above helps to resolve this issue. Evidence from a number of western NSW study locations indicates a correlation between the abundance of the archaeological record and periods of major environmental changes (Holdaway and Fanning 2014; Holdaway et al. 2002, 2008b, 2010). Comparison of the ages of the heat-retainer hearths and the ages of the surfaces on which they rest, obtained with OSL, indicates that the apparent increase in the number of sites (as well as the quan- tity of material in these sites) dating to the last 2,000 years is an artifact of preservation. There is a more recent archaeological record because there are more recent surfaces preserved where this record has accumulated (Holdaway and Fanning 2014; Holdaway et al. 2008b). Earlier surfaces have been removed, transporting and burying artifacts in huge volumes of sediment (Fanning et al. 2009b). Stone artifacts appear abundant because there is high visibility but this apparent abundance does not mean high populations or continuous occupation. If, as Gould (1991) and, to a degree, Allen (1974) suggest, some Australian environments are impoverished in a comparative sense, with seasonally available resources that are patchy in their distribution and prone to fluctuations, then the archaeological evidence that indicates the transport of flakes even in lithic rich environments may make more sense (Holdaway et al. 2013). Flakes were trans- ported because there was a need to exploit resources as they were encountered. It was always better to have access to stone immediately than to lose the poten- tial to exploit a resource by first having to spend time finding utilizable stone.

Risk Risk, when applied to stone artifacts, reflects the problem of being “caught short,” that is, not having the correct implements to exploit a resource when it is available. Elston (1990) differentiates between venture and contingency risk. Venture risk refers to the cost of lithic procurement whereas contingency risk is a measure of the cost of not having sufficient tools or stone for the task at hand. As discussed above, the wide availability of stone in the stony desert country of western NSW would argue against venture risk being significant, although some stone was moved from place to place. However, many of the assemblage characteristics suggest contingency risk was important. Results based on the cortex ratio analysis discussed above indicate a concern for using material effi- ciently. There was an emphasis on transporting material in the more efficient form of flakes rather than cores.

108 Holdaway, Shiner, Fanning & Douglass This efficiency in lithic transportation and use is not, however, matched by aspects of the food quest. Edwards and O’Connell (1995) discuss the evidence for seed grinding in Australia, commenting that the calorific return for seeds makes them a food of choice only when there are few alternatives. Using an optimal foraging model, use of seeds occurs only after exhaustion of resources that are more productive through overuse or higher population levels. In our western NSW study area, seed-grinding tools are present but mostly in the form of fragments of grinding dishes reused as heat retainers in hearths. Complete dishes are rare in the field, but these were a target for nineteenth- century collectors and, as Allen (1974) comments, many are now in museums or private collections. Seeds are an inefficient resource from which to obtain calories but were at times available in relatively dense patches, notably in the alluvial valleys adjacent to creeks after periods of rain. In an environment with rapidly diminishing resource abundance away from major water sources such as the Darling River (figure 5.1), utilization of seeds may have been more by way of necessity than optional. As Bamforth and Bleed (1997) comment, dis- cussions of risk must include situations in which loss of a resource becomes critical because there are no other resources to take its place. Under such situ- ations, humans will adapt by making use of foods that are available even if returns relative to labor cost are not great. Such behavior is not as inconsistent with the efficient use of lithic material as it may first appear. Appeals to risk may explain mid-Holocene changes in the nature of Australian stone artifact assemblages, but careful consideration in the way risk is determined is needed. Hiscock (1994), for instance, sees the advent of key retouched tool types, like the tula adze and backed microliths, as part of a suite of technologies that allowed occupation of the arid zone. He argues that abundance-orientated, mid-Holocene technological strategies based upon points and backed artifacts reflect increased foraging risks and procure- ment costs resulting from local scale factors linked to climatic variability and changing social dynamics. These replaced older assemblages, characterized by scrapers rather than microliths and adzes, where strategies that involved the extension of the use-life of an artifact through resharpening (i.e., the Frison effect) prevailed (Hiscock 2006). A different picture emerges, however, if occupation duration and mobility are considered rather than simply the presence of resharpening. Raw mater­ ial reduction intensity reflects occupation duration and the associated costs of material acquisition (see also Olszewski, chapter 4, this volume). Flakes are transported, not just tools, with risk managed by ensuring that individuals were supplied with the stone needed to have sharp edges available when required.

Cxonte t and Complexity on the Arid Margins of Australia 109 Resharpened forms played only a limited role in the technological systems of western NSW (contra other contexts such as western Eurasia; see Barton and Riel-Salvatore, chapter 2, this volume). Backed microliths and points are minor constituents of assemblages as are frequently resharpened scrapers. Veth (1993; Veth et al. 2011) also argues for the invention of key technologies: adzes for wood working, and deep wells and complex exchange systems that facilitated desert expansion from the refuges that protected human popula- tions during the Last Glacial Maximum. During the mid-to-late Holocene, greater levels of social complexity are argued for and these are linked to popu- lation increase (Smith 2004, 2005) or changes toward reduced mobility (Veth 2005a, 2005b, 2006). According to this scenario, environmental changes forced social, economic, and technological changes to sustain and ultimately grow population levels. In the parlance of evolutionary theory, population changes of this form fol- low a strategy of maximizing reproductive fitness (M. Allen 2004; Madsen et al. 1999). Resources augment population growth, the benefits of which are most apparent when the productivity of the environment is easy to predict. In these types of environments, the effective carrying capacity shows low vari- ability. As M. Allen (2004:200) comments, variation in carrying capacity may have unpleasant consequences for individuals and small groups but the effects are tolerable and populations survive. However, maximization of reproduc- tive fitness is not the only path to success. Long-term, a strategy of minimiz- ing reproductive fitness is more successful when environmental productivity is difficult to predict. To survive in a highly variable environment, adaptation must cope with a variety of risks and uncertainties. Bad years can lead to cata- strophic population losses, so the adaptations often form a type of buffer to alleviate the risk of such occurrences. Following this reasoning, the risk of not using grass seed in western NSW was that alternative, more profitable resources might not eventuate. Grass seeds either acted to sustain a population at a certain level, or alternatively were a resource that filled caloric gaps left by shortfalls in other resources. Whichever scenario best explains the evidence, grass seed exploitation pro- vided a mechanism that could offset particularly poor years. Curation of flakes provided access to a sharp edge whenever needed without the need to spend time locating suitable stone. By provisioning people rather than places (with the exception of large seed-grinding dishes), Aboriginal people were always in the position to exploit resources when they were encountered. Paleoenvironmental inferences derived from both heat-retainer hearths and OSL dates on sediments indicate an environment that was subject to periodic

110 Holdaway, Shiner, Fanning & Douglass floods (Holdaway et al. 2010). Comments byG ould (1991) indicate a resource base that was unpredictable in the short term (see also Holdaway et al. 2013). These sets of evidence point to Aboriginal people adapting to an unpredict- able environment with a limited range of resources, one where adaptation involved strategies designed to counter the uncertainty of resource availability. Certainly the transportation of stone artifacts, the concern with contingency risk rather than venture risk, and the need to exploit even the least efficient energy sources would fit with this concept of a buffering adaptation and a strategy of minimizing reproductive fitness.

Settlement and Ceremony Rhys Jones famously commented on the time Aboriginal people had to develop a complex ceremonial life outside that related to food gathering ( Jones 1977). Lourandos (1985, 1997) made much of the social changes that occurred with intensification, arguing that Aboriginal people were equiva- lent to examples of complex hunter-gatherers in other parts of the world. The Australian archaeological record appears to support intensification because the amount of archaeological material, as well as the diversity of types of this material, increases from the mid-Holocene. However, in the case of western NSW, these appearances are deceptive. While abundant, the archaeological record shows little evidence for redun- dancy in place use. A lack of redundancy in resource distribution means that artifact assemblages are variable, but this variability does not partition into discrete categories. Despite a desire to define functional site types, there is no agreed-upon set of criteria on which to base site type identification. Instead, site type definition involves a range of activities generalized to such an extent that, even when comparisons are made between continents using materials deposited by groups who cannot have shared any culture historical relation- ships, site types appear similar (Holdaway and Wandsnider 2006; Shiner 2008). Settlement pattern archaeology as traditionally conceived is ineffective when based on such variability (Allen et al. 2008; Holdaway and Fanning 2008). The complexity of Aboriginal social structure and religious beliefs is justifi- ably famous and the subject of extended, anthropological study. Archaeologists have sought the origins of this system, arguing that ideological systems, like other aspects of social life, are interpretable as part of an adaptive strategy that allowed access to neighboring areas during periods of resource depletion (e.g., Gould 1980; Witter 2007). Witter argues that the last 2,000 years in western NSW saw a proliferation of ceremonial sites identified on the basis of large

Cxonte t and Complexity on the Arid Margins of Australia 111 size (implying the contiguous presence of multiple groups of people); exten- sive charcoal-rich deposits, the presence of rare and specialized artifacts; and locations near rock art or dense resources capable of feeding large numbers of people. In addition, a spatial distribution of features includes hearths, work- shops, and art; evidence for seed grinding; and differences in the composition of artifact scatters (greater and more extensive than in “ordinary” large camps). As with studies that seek to define other functional site types, these criteria effectively conflate temporal and spatial variability. Large site size on its own does not imply large numbers of people, nor is spatial proximity an indica- tor of contemporaneity (Holdaway and Fanning 2014; Holdaway et al. 2005). The size of a site as exposed on the surface is a function of geomorphological processes rather than occupation size, and the same is true of charcoal-rich deposits (Fanning and Holdaway 2004). Deflation affects most sites, so pres- ervation of charcoal, together with any sediment, is the exception rather than the rule (Holdaway et al. 2008a). Its presence or absence cannot be used to estimate the number of people at a particular time and place. Because we are unable to show that features were in use at the same time, the spatial proxim- ity of hearths, workshops, art, grinding implements, and artifact scatters does not support the inference of large numbers of people (Holdaway et al. 2008b). Where absolute age determinations for hearths are available, adjacent hearths differ substantially in age, so spatial proximity cannot be used to imply con- temporaneity (Holdaway et al. 2002, 2005; Shiner 2008). The presence of rare artifact forms may relate to large numbers of people but may also be an indica- tor of extended occupation duration (Holdaway et al. 2008a). We have studied two of the four ceremonial centers Witter identifies as “key Dreaming Track sites” in western NSW: Sandy Creek at Fowlers Gap and Burkes Cave (labeled as Bourkes Cave by Witter) (Holdaway and Fanning 2014; Holdaway et al. 2008a; Shiner et al. 2005, 2007). Neither site has stone artifact assemblages consistent with the criteria listed above. In the case of Sandy Creek, the OSL and radiocarbon chronology shows repeated uses sepa- rated by centuries, rather than a single continuous occupation. Ethnohistoric accounts of ceremonies exist in western NSW and there is little reason to doubt their existence in the past. Certainly, rock art studies indicate a long history of activity and stone arrangements are common to some areas. Littleton’s (2007) study of burials also indicates ceremonial activ- ity in the past. But, unlike Witter (2007), Littleton makes no claim of a strong relationship between time and space as a key component of ceremonial activ- ity. Instead, she argues that the accumulation of burials at one location, form- ing what at first sight may appear to be cemeteries, are in fact a palimpsest

112 Holdaway, Shiner, Fanning & Douglass of individual burials that accumulated through time as a response to the long-term presence of relatively high ground suitable for interment. Applying Littleton’s reasoning to Witter’s interpretations, we might expect the evidence for ceremonial activity to accumulate in a wide variety of locations as the resource abundance necessary for relatively large numbers of people to con- gregate together varied in time and space (Holdaway and Allen 2012). These places would of course be difficult to identify archaeologically since they would not differ in content from any other. A more productive direction is to consider settlement and ceremony as a buffering adaptation employing the evolutionary theory cited above. As many scholars have commented, trade and exchange systems may act to ensure access to new regions during times of hardship (e.g., Gould 1980). However, there is another way to think about these systems as well as other ceremonial behavior. Dunnell (1989) argues that non-reproductive activities—what might be thought of as cultural elaboration or ceremonial activities—provide “a sink of ‘excess’ time and resources that can be devoted to subsistence reproduc- tion under stressful conditions” (Dunnell 1989:245). Madsen et al. (1999) relate the expenditure of energy on cultural elaboration to environmental predict- ability. Those groups who inhabit unpredictable environments and practice cultural elaboration will experience a lower variance in long-term fitness than those who refrain from such elaboration. In western NSW, cultural elabora- tion seen in rock art, the creation of Dreaming Tracks and stone arrangements, exchange systems, and burials might not, as Jones (1977) once suggested, be a reaction to available leisure time. Instead, it may be seen as the material manifestations of the practice of Aboriginal sociopolitics, a net outcome of which was a hedging strategy that provided a buffer against drastic changes in an unpredictable environment. It might be, as Dunnell proposed, that cultural elaboration served to “mop up” excess time and resources when times were good, allowing these to be released when conditions deteriorated. But it is also possible to interpret this behavior in a way closer to Gould’s notions: cul- tural elaboration developed in such a way that it facilitated mobility, thereby allowing local population reduction and, at times, local population extinction through complete abandonment (Holdaway and Allen 2012).

Population Understanding past population dynamics rests on different types of analy- ses: settlement demography, dating curves, and carrying capacity approaches (e.g., Kirch and Rallu 2007). With the exception of mounds, habitation

Cxonte t and Complexity on the Arid Margins of Australia 113 structures do not preserve in much of Australia, so archaeologists have used counts of stone artifacts as a proxy for past population numbers, combining these with reconstructions of past settlement systems. However, criticisms are leveled at counts of artifacts as a direct measure of past populations because of the influence of different technologies producing variable numbers of objects from each knapping episode (e.g., Hiscock 1986). Differential preservation and visibility also have an effect on the number of artifacts present (Fanning and Holdaway 2004). An alternative is to use the numbers of radiocarbon determinations as the “artifact” to be counted (the dating curves of Kirch and Rallu 2007; e.g., Smith et al. 2008; Williams et al. 2010). But as discussed above, the age of the record and therefore the number of radiocarbon deter- minations is very dependent on surface preservation as well as preservation of the material dated (usually charcoal). Therefore, like counting artifacts, the number of radiocarbon determinations may provide an inaccurate estimate of past population numbers. Finally, population numbers have been based on estimates of environmental carrying capacity, most notably by Birdsell (1953). Broadly speaking, these techniques support three models of population increase in Australia: Birdsell (1953, 1957) suggested a relatively rapid popula- tion increase after initial human colonization, up to a level determined by the carrying capacity, and then a steady state after that; Beaton (1990) argued for a steady increase in population from colonization until European arrival; and Lourandos (1983; Lourandos and Ross 1994) argues for a late efflorescence in population numbers. All three models deal with the length of Australian pre-European history rather than the specifics of population change in one region during the late Holocene. However, local application of the continental models is problematic since all three involve a Malthusian increase in popula- tion, albeit at different rates and different times. Such an increase cannot have occurred in all parts of the continent at the same time. Growth rates of less than 1 percent would, given the length of the Holocene, produce population levels well beyond even the revisionist population estimates of Butlin (1983). At one time it was thought that high rates of population growth were pos- sible for hunter-gatherer populations but were controlled by cultural prac- tices (e.g., polygyny, marriage age restrictions, infanticide, etc.). However, the ethnographic studies on which these were based have lately been called into question. Venereal disease is thought to have distorted the fertility measures made for twentieth-century hunter-gatherer populations (Pennington 2001). According to Boone (2002), the apparent contradiction of high rates of pop- ulation increase yet low overall population numbers can only be solved by recognizing the limits imposed by subsistence efficiency. Population crashes

114 Holdaway, Shiner, Fanning & Douglass every 50 years brought about by climatically and/or environmentally imposed decreases in the resource base with an impact on 25 percent of the population will, over time, lead to essentially stable population numbers, albeit numbers that show fluctuations within any one 50-year cycle. Reduce the impact to 15 percent and the modeled population shows an exponential increase. The problems with direct measures of past population numbers outlined above still remain, but some inferences may be made concerning popula- tion dynamics nevertheless. Occupation at any one location was intermittent (Holdaway and Fanning 2014; Holdaway et al. 2002, 2005). While people may well have journeyed to neighboring locations, comparison of the record from our different study areas, like Stud Creek and Fowlers Gap, suggests that population levels were never large enough to occupy many locations at once. Added to this, correlations with global climatic changes suggest that regions were at times occupied at very much reduced population levels, in concert with variation in environmental productivity (Holdaway et al. 2010). Intermittent occupation suggests the importance of subsistence efficiency for interpreting past population levels, as does the evidence for the efficient use of stone and the reliance on resources with high processing costs. If the environment was as unpredictable as suggested above, then Aboriginal people faced a challenge in adapting to this uncertainty. Mobility was no doubt a key adaptive strategy (Gould 1991). If a fluctuating environment led to marked differences in local productivity, then the shift from good to bad years encouraged variations in population levels at the local level as people moved into an area and then away. Long term, and over very large geographic regions, these fluctuations evened out, providing an essentially stable population, but locally, populations pos- sibly showed marked presence/absence oscillations. The complex ceremonial life Aboriginal people developed needs to be inter- preted in accordance with these types of population fluctuations. Knowledge transfer was necessary despite the fact that populations were small and, at times, dispersed and parts of the landscape might not be visited within a generation or more. Under these conditions the inadequacy of a Cartesian- based model, where all information is locked to particular places in the land- scape, becomes understandable. Rather than attempting to memorize lists of resources available at particular locations at different times of the year, which at any rate would be subject to change given environmental shifts, it would be more useful to pass on information concerning where resources might be found, given particular environmental conditions (Holdaway and Allen 2012).

Cxonte t and Complexity on the Arid Margins of Australia 115 Cuoncl sions Understanding stone artifact assemblage composition requires a consider- ation of context: the age of the artifacts, the raw material form, the tech- nology employed, and the artifact functions. Access to raw material has an impact on assemblage composition as has mobility and occupation duration. Understanding why people made particular economic choices requires the integration of large numbers of data sets, including not only the measure- ment of many stone artifacts, but also a good understanding of occupation and geochronology. In arid and semiarid Australia, many Australian analysts, faced with thou- sands of artifacts exposed in apparently undatable surface deposits, have retreated to the study of retouched tools rather than a consideration of whole assemblages. Despite recognition of the difficulty of estimating the size of past populations based on artifacts frequencies, assessments of mid-to-late Holocene Australia have frequently sought explanations for apparent popula- tion increase. The research summarized here indicates that it is unretouched flakes, rather than just retouched tools, that show evidence of curation. Aboriginal people were concerned with using raw material efficiently but did this by provision- ing people rather than places. The movement of raw material in the form of flakes provided the most efficient means of resource transport. Such econo- mizing behavior ties in with other indications of a life adapted for habita- tion in a region where resources were sparse and unpredictable. Rather than intensification involving a movement toward coping with an ever-increasing population, the mid-to-late Holocene record from Australia’s NSW desert margins indicates strategies where population levels varied locally but were stable regionally, ensuring long-term human survival in the face of a fluctuat- ing and unpredictable climate.

Acknowledgments We are indebted to the Indigenous traditional owners of country in west- ern NSW for their permission to undertake research, and for their help and support with fieldwork. The initial impetus for the research came from Dan Witter, formerly with the NSW National Parks and Wildlife Service, and initial funding came from the Australian Institute for Aboriginal and Torres Strait Islander Studies. Students from La Trobe, Auckland, and Macquarie Universities provided invaluable labor in the field. We thank the University of NSW and the NSW National Parks and Wildlife Service for permission

116 Holdaway, Shiner, Fanning & Douglass to conduct the research at Fowlers Gap Arid Zone Research Station, and in Sturt and Paroo-Darling National Parks. Work at Pine Point and Langwell was conducted with the support of the Harvy and Harrison families. Funding was provided by La Trobe University, the University of Auckland, Macquarie University, the NSW National Parks and Wildlife Service, and the Australian Research Council. All radiocarbon determinations were performed by the University of Waikato Radiocarbon Dating laboratory. We thank Alan Hogg and his colleagues for their help with the samples. OSL determinations were carried out by Ed Rhodes. Harry Allen, Melinda Allen, Mark Busse, Ben Davies, Judith Littleton, Mark Baumler, Trudy Doelman, and Harold Dibble provided valuable comments on earlier drafts of this chapter.

References Allen, Harry. 1974. “The Bagundji of the Darling Basin: Cereal Gatherers in an Uncertain Environment.” World Archaeology 5(3):309–322. http://dx.doi.org/10.1080 /00438243.1974.9979576. Allen, Harry, Simon J. Holdaway, Patricia C. Fanning, and Judith Littleton. 2008. “Footprints in the Sand: Appraising the Archaeology of the Willandra Lakes, Western New South Wales, Australia.” Antiquity 82(315):11–24. http://dx.doi.org /10.1017/S0003598X0009640X. Allen, Melinda. 2004. “Bet-Hedging Strategies, Agricultural Change, and Unpredictable Environments: Historical Development of Dryland Agriculture in Kona, Hawaii.” Journal of Anthropological Archaeology 23(2):196–224. http://dx.doi .org/10.1016/j.jaa.2004.02.001. Bamforth, Douglas B., and Peter Bleed. 1997. “Technology, Flaked Stone Technology, and Risk.” In Rediscovering Darwin: Evolutionary Theory in Archaeology, ed. C. Michael Barton and Geoffrey A. Clark, 109–140. Archaeological Papers No. 7. Washington, DC: American Anthropological Association. Beaton, John M. 1990. “The Importance of Past Population for Prehistory.” In Hunter Gatherer Demography: Past and Present, ed. Betty Meehan and N. White, 23–40. Oceania Monographs 39. Sydney: University of Sydney. Birdsell, Joseph. 1953. “Some Environmental and Cultural Factors Influencing the Structuring of Australian Aboriginal Populations.” American Naturalist 87(834):171–207. http://dx.doi.org/10.1086/281776. Birdsell, Joseph. 1957. “Some Population Problems Involving Pleistocene Man.” Cold Spring Harbor Symposia on Quantitative Biology 22: 47–69. http://dx.doi.org/10 .1101/SQB.1957.022.01.008.

Cxonte t and Complexity on the Arid Margins of Australia 117 Boone, James L. 2002. “Subsistence Strategies and Early Human Population History: An Evolutionary Ecological Perspective.” World Archaeology 34(1):6–25. http://dx.doi.org/10.1080/00438240220134232. Butlin, Noel G. 1983. Our Original Aggression: Aboriginal Populations of Southeastern Australia 1788–1850. Sydney: Allen and Unwin. Dibble, Harold L., Utsav A. Schurmans, Radu P. Iovita, and Michael V. McLaughlin. 2005. “The Measurement and Interpretation of Cortex in Lithic Assemblages.” American Antiquity 70(3):545–560. http://dx.doi.org/10.2307/40035313. Doelman, Trudy, and Simon J. Holdaway. 2011. “Standardisation and Design: The Tula Adze.” In Keeping Your Edge, ed. Benjamin Marwick and Alexander Mackay. BAR International Series 2273. Oxford: British Archaeological Reports. Douglass, Matthew J. 2010. “The Archaeological Potential of Informal Lithic Technologies: A Case Study of Assemblage Variability in Western New South Wales, Australia.” PhD dissertation, University of Auckland, . Douglass, Matthew J., and Simon J. Holdaway. 2011. “Quantifying Stone Raw Material Size Distributions: An Archaeological Application for Investigating Cortex Proportions and Stone Artefact Curation in Lithic Assemblages.” In Changing Perspectives in Australian Archaeology, Part IV, ed. Jim Specht and Robin Torrence. Technical Reports of the Australian Museum Online, 23. Sydney: Australian Museum. Douglass, Matthew J., Simon J. Holdaway, Patricia C. Fanning, and Justin I. Shiner. 2008. “An Assessment and Archaeological Application of Cortex Measurement in Lithic Assemblages.” American Antiquity 73:513–526. Dunnell, Robert C. 1989. “Aspects of the Application of Evolutionary Theory in Archaeology.” In Archaeological Thought in America, ed. Carl C. Lamberg-Karlovsky, 35–49. Cambridge: Cambridge University Press. http://dx.doi.org/10.1017/CBO978 0511558221.003. Edwards, Doug, and James F. O’Connell. 1995. “Broad Spectrum Diets in Arid Australia.” Antiquity 69(265):769–783. http://dx.doi.org/10.1017/S0003598X00082326. Elston, Robert G. 1990. “A Cost–Benefit Model of Lithic Assemblage Variability.” In The Archaeology of James Creek Shelter, ed. Robert G. Elston and Elizabeth E. Budy, 153–163. Anthropological Papers No. 115. Salt Lake City: University of Utah Press. Fanning, Patricia C., and Simon J. Holdaway. 2001. “Temporal Limits to the Archaeological Record in Arid Western NSW, Australia: Lessons from OSL and Radiocarbon Dating of Hearths and Sediments.” In Australasian Connections and New Directions: Proceedings of the 7th Australasian Archaeometry Conference, ed. Martin Jones and Peter Sheppard, 91–111. Research in Anthropology and Linguistics 5. Auckland: University of Auckland.

118 Holdaway, Shiner, Fanning & Douglass Fanning, Patricia C., and Simon J. Holdaway. 2004. “Artifact Visibility at Open Sites in Western New South Wales, Australia.” Journal of Field Archaeology 29(3–4):255– 271. http://dx.doi.org/10.1179/jfa.2004.29.3-4.255. Fanning, Patricia C., Simon J. Holdaway, and Rebecca Phillipps. 2009a. “Heat Retainer Hearth Identification as a Component of Archaeological Survey in Western NSW, Australia.” In New Directions in Archaeological Science, ed. Andrew Fairbairn and Sue O’Connor, 13–23. Terra Australis 28. Canberra: ANU E-Press. Fanning, Patricia C., Simon J. Holdaway, Edward J. Rhodes, and Tessa Bryant. 2009b. “The Surface Archaeological Record in Arid Australia: Geomorphic Controls on Preservation, Exposure and Visibility.” Geoarchaeology: An International Journal 24(2):121–146. http://dx.doi.org/10.1002/gea.20259. Fanning, Patricia C., Simon J. Holdaway, and Edward Rhodes. 2007. “A Geomorphic Framework for Understanding the Surface Archaeological Record in Arid Environments.” Geodinamica Acta 20(4):275–286. http://dx.doi.org/10.3166/ga.20 .275-286. Fanning, Patricia C., Simon J. Holdaway, and Edward Rhodes. 2008. “A New Geoarchaeology of Aboriginal Artefact Deposits in Western NSW, Australia: Establishing Spatial and Temporal Geomorphic Controls on the Surface Archaeological Record.” Geomorphology 101(3):524–532. http://dx.doi.org/10.1016/j .geomorph.2007.04.027. Flenniken, John, and Peter White. 1985. “Australian Flaked Stone Tools: A Technological Perspective.” Records of the Australian Museum 36(3):131–151. http://dx.doi.org/10.3853/j.0067-1975.36.1985.342. Gould, Richard A. 1969. “Puntutjarpa Rockshelter: A Reply to Messrs. Glover and Lampert.” Archaeology and Physical Anthropology in Oceania 4:229–37. Gould, Richard A. 1980. Living Archaeology. Cambridge: Cambridge University Press. Gould, Richard A. 1991. “Arid-Land Foraging as Seen from Australia: Adaptive Models and Behavioral Realities.” Oceania 62(1):12–33. http://dx.doi.org/10.1002/j .1834-4461.1991.tb02366.x. Griffiths, Tom. 1996. Hunters and Collectors: The Antiquarian Imagination in Australia. Cambridge: Cambridge University Press. Hiscock, Peter. 1986. “Technological Change in the Hunter River Valley and the Interpretation of Late Holocene Change in Australia.” Archaeology in Oceania 21(1):40–50. http://dx.doi.org/10.1002/j.1834-4453.1986.tb00123.x. Hiscock, Peter. 1994. “Technological Responses to Risk in Holocene Australia.” Journal of World Prehistory 8(3):267–292. http://dx.doi.org/10.1007/BF02221051.

Cxonte t and Complexity on the Arid Margins of Australia 119 Hiscock, Peter. 2003. “Prehistoric Populations: Australia and the Pacific.” In Encyclopedia of Population, vol. 2. ed. Paul G. Demeny and Geoffrey McNicoll, 796–798. New York: Macmillan Reference. Hiscock, Peter. 2006. “Blunt and to the Point: Changing Technological Strategies in Holocene Australia.” In Archaeology in Oceania: Australia and the Pacific Islands, ed. Ian Lilley, 69–95. Malden, MA: Blackwell Publishing. http://dx.doi.org/10.1002 /9780470773475.ch4. Holdaway, Simon J., and Harry Allen. 2012. “Placing Ideas on the Land: Practical and Ritual Training Amongst the Australian Aborigines.” In Archaeology and Apprenticeship: Body Knowledge, Identity, and Communities of Practice, ed. W. Wendrich, 79–98. Tucson: University of Arizona Press. Holdaway, Simon J., and Matthew J. Douglass. 2012. “A Twenty-First Century Archaeology of Stone Artifacts.” Journal of Archaeological Method and Theory 19(1):101–131. http://dx.doi.org/10.1007/s10816-011-9103-6. Holdaway, Simon J., and Patricia C. Fanning. 2008. “Developing a Landscape History as Part of a Survey Strategy: A Critique of Current Settlement System Approaches Based on Case Studies from Western New South Wales, Australia.” Journal of Archaeological Method and Theory 15(2):167–189. http://dx.doi.org/10.1007 /s10816-008-9051-y. Holdaway, Simon J., and Patricia C. Fanning. 2010. “Geoarchaeology in Australia: Understanding Human-Environment Interactions.” In Australian Landscapes, vol. 346. ed. Paul Bishop and Brad Pillans, 71–85. Geological Society of London Special Publication. London: Geological Society of London. Holdaway, Simon J., and Patricia C. Fanning. 2014. Geoarchaeology of Aboriginal Landscapes in Semiarid Australia. Collingwood, VIC: CSIRO. Holdaway, Simon J., Patricia C. Fanning, and Matthew J. Douglass. 2013. “A New Ecological Framework for Understanding Human-Environment Interactions in Arid Australia.” In Environment and Technology: Intersections and Transformations, ed. David Frankel, Susan Lawrence, and Jennifer Webb, 51–68. Melbourne: Routledge. Holdaway, Simon J., Patricia C. Fanning, Martin Jones, Justin I. Shiner, Dan Witter, and Geoff Nicholls. 2002. “Variability in the Chronology of Late Holocene Aboriginal Occupation on the Arid Margin of Southeastern Australia.” Journal of Archaeological Science 29(4):351–363. http://dx.doi.org/10.1006/jasc.2002.0719. Holdaway, Simon J., Patricia C. Fanning, and Edward Rhodes. 2008a. “Assemblage Accumulation as a Time Dependent Process in the Arid Zone of Western New South Wales, Australia.” In Time in Archaeology: Time Perspectivism Revisited, ed. Simon J. Holdaway and LuAnn Wandsnider, 110–133. Salt Lake City: University of Utah Press.

120 Holdaway, Shiner, Fanning & Douglass Holdaway, Simon J., Patricia C. Fanning, and Edward Rhodes. 2008b. “Challenging Intensification: Human–Environment Interactions in the Holocene Geoarchaeological Record from Western New South Wales, Australia.” Holocene 18(3):403–412. http://dx.doi.org/10.1177/0959683607087930. Holdaway, Simon J., Patricia C. Fanning, Edward Rhodes, Samuel Marx, Bruce Floyd, and Matthew J. Douglass. 2010. “Human Response to Paleoenvironmental Change and the Question of Temporal Scale.” Palaeogeography, Palaeoclimatology, Palaeoecology 292(1-2):192–200. http://dx.doi.org/10.1016/j.palaeo.2010.03.043. Holdaway, Simon J., Patricia C. Fanning, and Justin I. Shiner. 2005. “Absence of Evidence or Evidence of Absence? Understanding the Chronology of Indigenous Occupation of Western New South Wales, Australia.” Archaeology in Oceania 40(2):33–49. http://dx.doi.org/10.1002/j.1834-4453.2005.tb00584.x. Holdaway, Simon J., Patricia C. Fanning, and Dan Witter. 2000. “Prehistoric Aboriginal Occupation of the Rangelands: Interpreting the Surface Archaeological Record of Far Western New South Wales, Australia.” Rangeland Journal 22(1):44–57. http://dx.doi.org/10.1071/RJ0000044. Holdaway, Simon J., Geoffrey C. P. King, Matthew J. Douglass, and Patricia C. Fanning. 2015. “Human–Environment Interactions at Regional Scales: The Complex Topography Hypothesis Applied to Surface Archaeological Records in Australia and North America.” Archaeology in Oceania 50:58–69. Holdaway, Simon J., Justin I. Shiner, and Patricia C. Fanning. 2004. “Hunter- Gatherers and the Archaeology of Discard Behavior: An Analysis of Surface Stone Artifacts from Sturt National Park, Western New South Wales, Australia.” Asian Perspective 43(1):34–72. http://dx.doi.org/10.1353/asi.2004.0005. Holdaway, Simon J., Justin I. Shiner, and Patricia C. Fanning. 2008c. “Assemblage Formation as a Result of Raw Material Acquisition in Western New South Wales, Australia.” Lithic Technology 23:73–85. Holdaway, Simon J., and LuAnn Wandsnider. 2006. “Temporal Scales and Archaeological Landscapes from the Eastern Desert of Australia and Intermontane North America.” In Confronting Scale in Archaeology, ed. Gary Lock and Brian L. Molyneaux, 183–202. New York: Springer Science and Business Media. Jansen, John D., and Gary J. Brierley. 2004. “Pool-Fills: A Window to Palaeoflood History and Response in Bedrock-Confined Rivers.” Sedimentology 51(5):901–925. http://dx.doi.org/10.1111/j.1365-3091.2004.00643.x. Jones, Rhys. 1977. “The Tasmanian Paradox.” In Stone Tools as Cultural Markers: Change, Evolution and Complexity, ed. Richard Wright, 189–204. Canberra: Australian Institute of Aboriginal Studies.

Cxonte t and Complexity on the Arid Margins of Australia 121 Kirch, Patrick V., and Jean-Louis Rallu. 2007. “Long-Term Demographic Evolution in the Pacific Islands.” In The Growth and Collapse of Pacific Island Societies: Archaeological and Demographic Perspectives, ed. Patrick V. Kirch and Jean-Louis Rallu, 1–13. Honolulu: University of Hawai’i Press. Kuhn, Steven L. 1994. “A Formal Approach to the Design and Assembly of Mobile Toolkits.” American Antiquity 59(3):426–442. http://dx.doi.org/10.2307/282456. Kuhn, Steven L. 1995. Mousterian Lithic Technology: An Ecological Perspective. Princeton: Princeton University Press. http://dx.doi.org/10.1515/9781400864034. Lin, Sam, Matthew J. Douglass, Simon J. Holdaway, and Bruce Floyd. 2010. “The Application of 3D Laser Scanning Technology to the Assessment of Ordinal and Mechanical Cortex Quantification in Lithic Analysis.” Journal of Archaeological Science 37 4):694–702. http://dx.doi.org/10.1016/j.jas.2009.10.030. Littleton, Judith. 2007. “From the Perspective of Time: Hunter-Gatherer Burials in South-Eastern Australia.” Antiquity 81(314):1013–1028. http://dx.doi.org/10.1017 /S0003598X00096083. Lourandos, Harry. 1983. “Intensification: A Late Pleistocene-Holocene Archaeological Sequence from South-Western Victoria.” Archaeology in Oceania 18:81–94. Lourandos, Harry. 1985. “Intensification and Australian Prehistory.” In Prehistoric Hunter-Gatherers the Emergence of Cultural Complexity, ed. T. Douglas Price and James A. Brown, 385–423. New York: Academic Press. Lourandos, Harry. 1997. Continent of Hunter-Gatherers: New Perspectives in Australian Prehistory. Cambridge: Cambridge University Press. Lourandos, Harry, and Anne Ross. 1994. “The Great ‘Intensification Debate’: Its History and Place in Australian Archaeology.” Australian Archaeology 39:54–63. Mackellar, Dorothea. 1908. “My Country.” Accessed April 2016. http://www.dorothe amackellar.com.au/. Madsen, Mark, Carl Lipo, and Michael Cannon. 1999. “Fitness and Reproductive Trade-Offs in Uncertain Environments: Explaining the Evolution of Cultural Elaboration.” Journal of Anthropological Archaeology 18(3):251–281. http://dx.doi.org /10.1006/jaar.1999.0342. Nelson, Margaret. 1991. “The Study of Technological Organization.” In Archaeological Method and Theory, vol. 3. ed. Michael B. Schiffer, 57–100. Tucson: University of Arizona Press. Pennington, Renee. 2001. “Hunter-Gatherer Demography.” In Hunter-Gatherers: An Interdisciplinary Perspective, ed. Catherine Panter-Brick, Robert Layton, and Peter Rowley-Conwy, 170–204. Cambridge: Cambridge University Press. Rhodes, Edward J., Patricia C. Fanning, Simon J. Holdaway, and Cynthia Bolton. 2009. “Ancient Surfaces? Dating Archaeological Surfaces in Western NSW Using

122 Holdaway, Shiner, Fanning & Douglass OSL.” In New Directions in Archaeological Science, ed. Andrew Fairbairn, Sue O’Connor and Benjamin Marwick, 189–200. Terra Australis 28. Canberra: ANU E-Press. Sandweiss, Daniel H., Kirk A. Maasch, and David G. Anderson. 1999. “Climate and Culture: Transitions in the Mid-Holocene.” Science 283(5401):499–500. http://dx .doi.org/10.1126/science.283.5401.499. Shiner, Justin I. 2008. Place as Occupational Histories: An Investigation of the Deflated Surface Archaeological Record of Pine Point and Langwell Stations, New South Wales, Australia. BAR International Series 1763. Oxford: British Archaeological Reports. Shiner, Justin I., Simon J. Holdaway, Harry Allen, and Patricia C. Fanning. 2005. “Understanding Stone Artefact Assemblage Variability in Late Holocene Contexts in Western New South Wales, Australia: Burkes Cave, Stud Creek and Fowlers Gap.” In Lithics Down Under: Australian Perspectives on Lithic Reduction, Use and Classification, ed. Christopher Clarkson and Lara Lamb, 67–80. BAR International Series 1408. Oxford: British Archaeological Reports. Shiner, Justin I., Simon J. Holdaway, Harry Allen, and Patricia C. Fanning. 2007. “Burkes Cave and Flaked Stone Assemblage Variability in Western New South Wales, Australia.” Australian Archaeology 64:35–45. Smith, Mike A. 2004. “TheG rindstone Assemblage from Puritjarra Rock Shelter: Investigating the History of Seed-Based Economies in Arid Australia.” In Archaeology from Australia, ed. Tim Murray, 168–186. Victoria: Australian Scholarly Publishing. Smith, Mike A. 2005. “Desert Archaeology, Linguistic Stratigraphy, and the Spread of the Western Desert Language.” In Desert Peoples: Archaeological Perspectives, ed. Peter Veth, Mike A. Smith, and Peter Hiscock, 222–242. Melbourne: Blackwell Publishing. http://dx.doi.org/10.1002/9780470774632.ch12. Smith, Mike A., Alan N. Williams, Chris M. S. Turney, and Matthew L. Cupper. 2008. “Human-Environment Interactions in Australian Drylands: Exploratory Time-Series Analysis of Archaeological Records.” Holocene 18(3):389–401. http://dx .doi.org/10.1177/0959683607087929. Torrence, Robin. 1989. “Retooling: Towards a Behavioral Theory of Stone Tools.” In Time, Energy and Stone Tools, ed. Robin Torrence, 57–66. Cambridge: Cambridge University Press. Veth, Peter M. 1993. Islands in the Interior: The Dynamics of Prehistoric Adaptations within the Arid Zone of Australia. Archaeological Series 3. Ann Arbor: International Monographs in Prehistory. Veth, Peter M. 2005a. “Between the Desert and the Sea: Archaeologies of the Western Desert and Pilbara Regions, Australia.” In 23° South: Archaeology and

Cxonte t and Complexity on the Arid Margins of Australia 123 Environmental History of the Southern Deserts, ed. Mike A. Smith and Paul Hesse, 132–141. Canberra: National Museum of Australia Press. Veth, Peter M. 2005b. “Cycles of Aridity and Human Mobility: Risk-Minimization Amongst Late Pleistocene Foragers of the Western Desert, Australia.” In Desert Peoples: Archaeological Perspectives, ed. Peter Veth, Mike A. Smith, and Peter Hiscock, 100–115. Melbourne: Blackwell Publishing. http://dx.doi.org/10.1002 /9780470774632.ch6. Veth, Peter M. 2006. “Social Dynamism in the Archaeology of the Western Desert.” In The Social Archaeology of Indigenous Societies, ed. Bruno David, Ian McNiven, and Bryce Barker, 242–253. Canberra: Aboriginal Studies Press. Veth, Peter M., Peter Hiscock, and Alan Williams. 2011. “Are Tulas and ENSO Linked in Australia?” Australian Archaeology 72:7–14. Webb, Cathie. 1993. “The Lithication of a Sandy Environment.” Archaeology in Oceania 28(3):105–111. http://dx.doi.org/10.1002/j.1834-4453.1993.tb00301.x. Williams, Alan N., Sean Ulm, Ian D. Goodwin, and Mike Smith. 2010. “Hunter- Gatherer Response to Late Holocene Climatic Variability in Northern and Central Australia.” Journal of Quaternary Science 25(6):831–838. http://dx.doi.org /10.1002/jqs.1416. Witter, Dan. 2007. “Aboriginal Dreaming and Aridity.” In Animals of Arid Australia: Out on Their Own? ed. C. Dickman, D. Lunney, and S. Burgin, 14–29. Mosman, Australia: Royal Zoological Society of New South Wales. http://dx.doi.org /10.7882/FS.2007.038.

124 Holdaway, Shiner, Fanning & Douglass 6

Although less of an exaggeration than it may have Theoretical Implications been several decades ago, it is still reasonable to claim of Artifact-Scatter Lithic today that archaeologists do not routinely investigate Assemblage Variability for surface scatters of artifacts, often composed princi- Mobility-Based Models of pally of lithic debitage. Despite persistent appeals to Technological Organization investigate lithic artifact scatters systematically (e.g., Chartkoff 1995; Fanning et al. 2009), these phenom- ena typically have been regarded as insignificant, barely Alan P. Sullivan III preservation-worthy, and largely without research merit (Upham 1994:119). On those occasions when lithic arti- fact scatters have been incorporated in regional studies, inferences regarding their behavioral or organizational significance tend to be weakly supported because of problematic research designs, analytical methods, and assertions regarding the meaning of assemblage differ- ences (Tainter and Bagley 2005). These archaeological phenomena rarely have been involved in any signifi- cant or unproblematic way in testing hypotheses and models that fundamentally and directly implicate them (important exceptions include Mehalic [2012], Powell and Klesert [1984], and Tainter [1979]). To moderate this situation, and provide some updates on recent developments, I briefly review the history of debitage studies in Southwest archaeology and evalu- ate two hypotheses (aging but still influential)—the Adaptive Diversity Hypothesis and the Expedient Core Hypothesis—that focus on understanding the causes of regional lithic-assemblage variability. With excava- tion and survey data from five types of archaeological DOI: 10.5876/9781607324942.c006

125 sites located in the Upper Basin, northern Arizona, I show that, with the benefit of interpretation-neutral taxa and unambiguous measures of assem- blage variability, strong inferences regarding regional technological organiza- tion can be developed without referring their validation to mobility models grounded in hunter-gatherer ethnography or ethnoarchaeology. In addition, the results of the Upper Basin study create epistemological space to consider the likelihood that mobility-based models of the organization of technology, which typically have focused on understanding how ecological factors influ- ence problem-solving strategies for making and using lithic artifacts (Nelson 1991), are constrained because of their theoretical dependence on obligate eco- logical paradigms (i.e., those that assume the distribution, abundance, and availability of resources are fixed [e.g.,K elly 1983]). In contrast, I suggest that, with the adoption of a facultative ecological perspective (i.e., one that pre- sumes the distribution, abundance, and availability of resources are not fixed [e.g., Smith 2011]), lithic-artifact-production strategies that enable the flow of resources from anthropogenic acquisition and processing sources (“niches” or patches) to home bases, which is the Upper Basin pattern (cf. Gumerman 1984:103), create a heterogeneous lithic landscape whose characteristics do not align well, if at all, with those stipulated by mobility-based models of tech- nological organization (see Barton and Riel-Salvatore, chapter 2, this volume; Holdaway et al., chapter 5, this volume).

Ana Aw kening of Interest in Southwest Debitage Until the mid-twentieth century, even if debitage artifacts were recovered in archaeological investigations (which cannot be taken for granted [see Haury 1950:238]), they were seldom, if ever, analyzed (Longacre 1967:122). Evidently, these “nondescript lots of flakes” were considered analytically worthless because “there is nothing distinguishing about them” (Roberts 1940:123). At best, assuming that categories such as “utilized flakes” (Rinaldo 1964:88) or “cutting edges” (Roberts 1931:158) actually refer to debitage (i.e., flaked-stone artifacts with positive percussion features only [Hiscock 2007:203]), the inter- pretive significance of these artifacts was not incorporated in any meaningful way for understanding assemblage variation and human behavior. This situation changed abruptly during the mid-1960s when, in an appar- ent case of convergent scientific inspiration, four studies of debitage appeared independently within two years of each other.1 Douglas Osborne (1965) exam- ined nearly 14,000 flakes from seven excavated puebloan sites on Wetherill Mesa in Mesa Verde National Park, southwestern Colorado. His principal

126 Sullivan objective was to track changes in raw material preferences and, secondarily, technological differences, by examining variation in percentages of raw materi- als, platform type (artificial vs. natural “heels”), flake origin (interior or exte- rior), flake shape (blocky or elongate), and weight. Osborne’s major findings were that (1) raw material preferences shifted through time and (2) the preva- lence of platform types and flake weight was affected by material type. Richard P. Wheeler’s collateral study (1965), which was based on selected artifacts from 11 excavated or tested puebloan sites on Wetherill Mesa in Mesa Verde,2 focused on use-wear analysis of 148 flakes, 24 blades, and 15 cores. Even though the ranges of length, width, thickness, and weight of the three afore- mentioned artifact classes were reported, there was no discussion of how such measurements were made on individual artifacts. Moreover, the value ranges did not figure in synthetic technological inferences, other than the observa- tion that certain forms were selected for subsequent use as scrapers and abrad- ers (Wheeler 1965:24). William A. Longacre and John M. Fritz analyzed 5,868 “lithic waste” arti- facts recovered from the excavation of Broken K Pueblo in east-central Arizona (Longacre 1967). Stage categories were used to develop inferences about flaked-stone technology: Stage I, core preparation; Stage II, flake pro- duction; and Stage III, tool manufacture. Longacre argued that patterned spa- tial variation in the density of “lithic waste” across the ruin was indicative of intrasettlement exchange, a practice that may have served to integrate the pueblo during various occupation periods. These three studies were based, obviously, on material recovered from mul- tiroom puebloan structures. In striking contrast is the Arch Lake site, a multi- component artifact-scatter (ad 1100–1500) situated on a low ridge of Holocene aeolian deposits on the east side of Arch Lake, hard against the New Mexico- Texas border about 20 miles southeast of Portales, New Mexico ( Jelinek 1966). The analysis of the Arch Lake site surface material represents the first study of its kind in Southwest archaeology and is groundbreaking in two respects. First, it was designed to ascertain whether variation in flake form could solve prob- lems of cultural classification and chronology for surface assemblages that lack temporally diagnostic projectile points or decorated ceramics. Second, clearly defined variables of artifact length, width, and thickness were used to mea- sure flake form. The study demonstrated that patterned differences among the percentages and metrics of bifacial, normal, and “Blackwater” retouch flake types covary with time. That is, in comparison to surface Archaic assemblages from the nearby Pecos valley ( Jelinek 1967), which disclosed high and low proportions of bifacial and normal end-scraper retouch flakes, respectively, the

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 127 later prehistoric assemblages (Pueblo II–Pueblo III) at Arch Lake had low percentages of both types, and Blackwater end-scraper retouch flakes were associated only with very late (Pueblo IV) prehistoric occupations.

An Expansion of Interest in Southwest Debitage: Exploring the Connections between Artifact Scatters and Mobility Patterns Since the publication of these seminal studies in the mid-1960s, consider- ation of Arch Lake kinds of archaeological phenomena—debitage assemblages and artifact scatters—in Southwest regional studies has accelerated (e.g., Adler 1992; Catlin 1986; Lange 1998; Longacre and Graves 1976; Mehalic 2012; Spurr et al. 2004; Sullivan 1983; Tainter 1979; Wait 1983; Wandsnider and Camilli 1996). Moreover, these phenomena have been linked to two hypotheses that were advanced to reset conventional thinking regarding the relations among mobility, resource productivity, and technological change in the prehistoric Southwest. The first hypothesis pertains to the role of artifact scatters—referred to as Limited Activity Sites (LAS)—in settlement systems. In this case, it was proposed that variation among regions with respect to the number of LAS is a measure of “adaptive diversity,” that is, the degree to which patterns of regional abandonment can be attributed to differences between mobility strategies (Upham 1984; see also Plog 1986:219–221). This hypothesis stipulates that LAS variation registers rates of residential mobility, rather than logisti- cal (non-residential) mobility, as the same populations toggle between mobile hunting-and-gathering and sedentary agriculture settlement-subsistence modes (Rushforth and Upham 1992:63). The second hypothesis—theE xpedient Core Hypothesis—emerged when archaeologists applied mobility-based models of technological organization (Binford 1979, 1980; Kelly 1983; Shott 1986) to the interpretation of South­ western lithic assemblage variability (Parry and Kelly 1987; Nelson 1991). In these cases, it was argued that, with the onset of sedentism, the production of “curated” tools (i.e., bifaces produced from standardized cores) was dis- continued in favor of “expedient flake tools” (Parry andK elly 1987:302–303; see also Torres 2000). Methodologically, this hypothesis stipulated that low biface:core ratios, low percentages of artifacts expressing “facial retouch,” and low percentages of debitage with abraded platforms reflected an emphasis on “expedient core technology” (Parry and Kelly 1987:291–292; see also Young 1994). With respect to variability at the level of site types, Nelson (1991:78–83; see also Magne 1989) distinguished assemblages among residences (i.e., residential

128 Sullivan assemblages should disclose “all stages of manufacturing debris” related to fab- ricating and “repairing tools and preparing cores for use elsewhere,” as well as tool fragments), camps (i.e., camp assemblages should disclose “late-stage manufacturing debris in the absence of tools”), and LAS (i.e., LAS assem- blages should disclose processing/retouch flakes, no primary reduction core debris, and “flakes, as a group, should be fairly homogeneous in form, reflect- ing a portion of a biface reduction sequence”). Since they were first proposed, further work has determined that these hypotheses and their empirical expectations are based on some problem- atic assumptions regarding the fidelity with which variability in mobility is expressed archaeologically (Kelly 1992; McCall 2012) and technologically (Eren and Andrews 2013). In addition, factors other than mobility, such as “processing requirements” (Tomka 2001), tool use-life variability (Surovell 2009:206–210), shifts in weapons technology (Railey 2010), and problematic regional sampling of LAS frequencies (Sullivan 1987) have been shown to entail similar assemblage-composition and site-type data, which are exclusive neither to the Adaptive Diversity nor the Expedient Core hypotheses. In what follows, I explore the consequences of these developments with an analysis of debitage and tool assemblages recovered from artifact scatters in the Upper Basin, which is located in northern Arizona just south of Grand Canyon (figure 6.1). Like many intensively studied areas of the northern American Southwest, the surface archaeology of this portion of the Grand Canyon region is characterized by abundant masonry ruins, artifact scatters, and fire-cracked-rock features (figure 6.2; Sullivan et al. 2002; cf. Spurr and Neff 2005:357). In addition, the Upper Basin and the eastern South Rim of the Grand Canyon are noted for abundant exposures of easily worked lithic raw material, principally Kaibab chert (Brown 1969). The area’s occupants also had unlimited access to sources of obsidian located to the south and southwest (Lesko 1989; Shackley 2005; also Carter et al. 2011). Hence, in this corner of the Southwest, technological strategies were not constrained by raw mate- rial availability, time stress, energy costs, or other factors, often stipulated in mobility-based models of technological organization, that affect assemblage composition and artifact form (Nelson 1991:59–62). Because these archaeo- logical phenomena originated after the establishment of sedentary villages (ca. ad 760; Kohler 1993:277–283), their accompanying assemblages should be indicative of “expedient core” technological strategies. In addition to testing the expectations of these models, however, I investigate the possibility that, by analyzing samples from a variety of contexts and applying standardized meth- ods, such as those used profitably at the Arch Lake site, new understandings

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 129 Figure 6.1. Location of the Upper Basin in northern Arizona.

will emerge regarding the complicated relations among anthropogenic ecol- ogy, technological organization, and assemblage composition.

Psar ing Assemblage Variability For the purposes of this study, I focus on 21 assemblages of flaked-stone artifacts (N = 15,544) recovered from five types of archaeological sites—lithic scatters (n = 12), fire-cracked-rock piles (i.e., mounded accumulations of fire- cracked rock; n = 2), fire-cracked-rock scatters (i.e., surface-level, unmounded concentrations of fire-cracked rock; n = 2), sherd-and-lithic artifact scatters

130 Sullivan Figure 6.2. Locations and spatial distribution of the five most common kinds of prehistoric archaeological phenomena in the Upper Basin, northern Arizona. (n = 2), and masonry ruins (n = 3). Based on calibrated radiocarbon dates and associated diagnostic ceramics, the fire-cracked-rock features (scatters and piles) and the artifact scatters (sherd-and-lithic scatters and “pure” lithic scatters) originated no earlier than ca. ad 775 and were reused, in some cases, until ca. ad 1650 (Cook 1995; Sullivan 1995:56–58; Sullivan et al. 2001:373). Of the three masonry ruins, one multistructure site has multiple tree-ring cut- ting dates spanning the period ad 1049–1060, one single-room structure has a mean ceramic date of ad 1061–1080, and one multiroom structure has tree-ring cutting dates of ad 1070 and ad 1080 (Sullivan and Ruter 2006:189). In view of these chronometric data, the elapsed time during which all 21 assemblages were formed ranges from the late eighth century ad to the mid-seventeenth century ad (cf. Parry and Kelly 1987; Railey 2010).

Methods Lithic artifacts were classified according to the manner in which force (per- cussion) features are expressed on their interior and exterior surfaces (Hiscock 2007): debitage artifacts are those with positive force features only, cores dis- close negative force features only, and tools have both negative and positive force features (note that negative force features [e.g., flake scars] on the exterior surface that predate the origination of the artifact are disregarded; Rozen and Sullivan 1989a:181–182).3 Length, width, and thickness measurements of com- plete flakes (n = 3,365), which are the basis for Relative Thickness calculations (RT = (L + W)/T), follow force-feature protocols (Sullivan 1995:54).4 Flakes were weighed individually to the nearest 0.1 g to enable comparisons of reduc- tion (Shott and Nelson 2008:29). At the assemblage level, various studies have shown that mean values of complete-flake relative thickness and weight vary predictably for tool production (high and low, respectively) and core reduc- tion (low and high, respectively; Railey 2010:267; Sullivan 2001; Sullivan and Rozen 1985; cf. Wilson and Andrefsky 2008). And, to maximize comparability with previous studies, variation in tool production is measured in terms of percentages of unifacially retouched pieces (n = 225) and bifacially retouched pieces (bifaces [n = 360] and projectile points [n = 131]; cf. Andrefsky 2008a:5; Barton and Riel-Salvatore 2014:338).

Results Examination of figure 6.3 reveals that, with the exception of fire-cracked- rock scatters, tool assemblages are dominated by discarded bifaces or biface

132 Sullivan Figure 6.3. Bar charts showing differences among site types with respect to debitage assemblage (Complete Flake Percent) and tool assemblage (Uniface Percent and Biface Percent) composition.

fragments. Of 57 cores recovered from these 21 assemblages, 48 specimens originated from three sites; hence, most assemblages (18 of 21) had fewer than two cores. These data attest to the fact that core reduction (Andrefsky 2008a)—expedient or otherwise—was not a predominant type of lithic arti- fact production at these post–ad 775 sites. In fact, the overall biface:core ratios for these assemblages—8.1 (491:57 for bifaces and projectile points combined) or 6.3 (360:57 for bifaces only)—exceed any reported value for either prece- ramic or ceramic-period assemblages (Parry and Kelly 1987:291–294; Railey 2010:273–274; Young 1994:148–149). For comparison, only 60 cores or core frag- ments were observed at 1,783 mapping units in the Upper Basin project area. Figure 6.3 also reveals a strong inverse relation between percentage of com- plete flakes in the debitage assemblage and the percentage of bifaces in the tool assemblage, as well as a strong positive covariation between complete- flake percentage and uniface percentage in the tool assemblage. In other words, as the percentage of bifaces decreases, there is a corresponding increase in the prevalence of complete flakes and unifaces (it is important to note here that the denominators used to calculate these percentages are independent). These

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 133 Figure 6.4. Scatterplot showing differences among 21 assemblages of complete flakes, distinguished by site type, based on mean values of weight (x-axis) and relative thickness (y-axis).

patterns appear to be related to the duration of the formation intervals of the scatters (conceived of here as “cumulative palimpsests” [Bailey 2007:204–205]), which range from hours (for lithic scatters) to decades (for masonry ruins; Binford 1986; Kuhn 1991:79; see also Surovell 2009:58; cf. Barton and Riel- Salvatore 2014:339–340). Examination of figure 6.4 enables a deeper understanding of the relation between tool assemblage and debitage assemblage variation.5 For instance, the

134 Sullivan relative-thickness and weight values for complete flakes recovered from the masonry-ruin artifact scatters are arguably attributable to the requirements of a wide range of activities (Key and Lycett 2014; Nelson 1991:79–82), including the production of other classes of artifacts, such as ceramics (Sullivan 1988), as well as other types of stone tools, particularly unifacially retouched pieces (Eren and Prendergast 2008) and ground-stone artifacts (Wheeler 1965:27). The formation of the artifact scatters associated with fire-cracked-rock piles is fairly well understood (Sullivan et al. 2001). These thermal features and their assemblages of ceramics, ground-stone artifacts, and flaked-stone artifacts were used to process wild plants, presumably in bulk quantities (Sullivan and Ruter 2006). In addition, debitage assemblages from fire-cracked-rock piles, which were visited repeatedly in many cases, were affiliated with intensive biface production (Purtill 1995). Some of these newly produced tools, particu- larly projectile points, were intended for use elsewhere. Other tools, produced somewhere else, were introduced into the areas surrounding the processing features and remanufactured or discarded at those locations (Cook 1995). Whatever the particulars of the tool production process may have been—newly manufactured or remanufactured bifaces—the composition of the resulting debitage and tool assemblages closely mimics “pure” lithic scatters (see below). Interestingly, much of the processing technology that involved ground-stone artifacts and ceramics had been transported to these locations (Wheeler 1965; C. Osborne 1965); these artifact inventories appear to have been refurbished as well, with laterally cycled artifact fragments obtained, presumably, from nearby perennial residences (Greenberg 2013). The overall effect on the land- scape was the creation of obtrusive non-residential places that were visited episodically for centuries (Wandsnider and Camilli 1996). In contrast to the masonry-ruin and fire-cracked-rock-pile tool and deb- itage assemblages, those associated with fire-cracked-rock scatters reflect a greater emphasis on uniface production in view of their relatively thick but light complete flakes (figure 6.4), attributes that align with the production of what conventionally have been termed scraper retouch flakes (Clarkson 2008; Eren et al. 2005; Frison 1968; see also Shott 1995). Moreover, the scale or inten- sity of biface production was less than that inferred for the fire-cracked-rock pile features. Overall, it is thought that the processing activities conducted at these features (Cook 1995) were relatively short-lived (Nelson 1991:83; C. Osborne 1965). For the 12 “pure” lithic scatters in the study, the low percentage of complete flakes in the debitage assemblage (figure 6.3), the high percentage of bifaces in the tool assemblage (figure 6.3), and the high mean relative-thickness and low

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 135 mean weight values for complete flakes (figure 6.4), signify intensive biface manufacture (Sullivan and Rozen 1985). Because fragmentary bifaces (n = 54) dramatically outnumber complete specimens (n = 3), it is reasonable to infer that the majority of them were broken and discarded during manufacture; of the numerous incomplete projectile points (n = 27; whole projectile points = 4), a number of them may have been introduced to these scatters in fragmentary form (after having been detached from their shafts) and commingled with those that broke during manufacture (Thompson 2003; cf. Andrefsky 2008b). Either way, these concentrations formed quickly (Nelson 1991:83) and were associated with few, if any, of the processing activities inferred for the other site types. What little variation exists among the concentrations with respect to mean complete-flake relative thickness and weight is likely attributable to skill differences or artifact design experimentations of individual knappers (Williams and Andrefsky 2011). Despite extensive analysis of excavated data (Sullivan 1995) and GIS analy- ses (Szeghi 2012), sherd-and-lithic artifact scatters are a challenging class of archaeological phenomena to interpret (Whalen 1986:78). Nevertheless, based on debitage and tool assemblage composition (figure 6.3) and mean relative- thickness values for complete flakes (figure 6.4), the most secure inference regarding the origins of these assemblages is that lithic artifact production seems to have focused on biface manufacture (Sullivan 1995:56, 60). The extent to which these artifacts, as well as others that may have been brought to these locations, were involved in other activities is largely conjectural, but, as the following discussion implies, sherd-and-lithic scatters appear to have materi- alized in ways that do not overlap with other types of artifact scatters.

Discussion In order to contextualize these analyses of assemblage composition based on excavated samples, it is instructive to consider the implications of high- intensity survey data. For instance, the absence of statistically significant associations of bifacial tools (bifaces and projectile points) and ground-stone artifacts (manos and metates) with either fire-cracked-rock piles or fire- cracked-rock scatters (chi-square = .28, df = 1, p = .59) raises the possibility that the principal factor that differentiates these types of thermal-features is the trajectories of their respective formation histories. That is, both types of processing areas were outfitted with basically the same set of “founder” arti- fact classes (figure 6.5). Thereafter, some thermal-feature processing areas had their use-lives extended or rejuvenated, which involved the refurbishing

136 Sullivan Figure 6.5. Bar charts of regional survey-based variation in lithic assemblages at thermal-feature mapping units: of 188 fire-cracked-rock pile mapping units, 24 disclosed bifacial tools (either whole or fragmentary bifaces or projectile points) and 57 disclosed ground-stone artifacts (either whole or fragmentary manos or metates); of 58 fire-cracked- rock scatter mapping units, 19 disclosed bifacial tools (either whole or fragmentary bifaces or projectile points) and 37 disclosed ground-stone artifacts (either whole or fragmentary manos or metates). and reuse of the original founder assemblages and significant scaling up of bifacial-tool production. In marked contrast to thermal-feature artifact scatters, sherd-and-lithic scatters have significantly more ground-stone and fewer bifacial artifacts than expected whereas lithic scatters have significantly more bifacial and fewer ground-stone artifacts than expected (chi-square = 11.19, df = 1, p < .001; figure 6.6). These dramatic patterned differences in assemblage composition indicate

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 137 Figure 6.6. Bar charts of regional survey-based variation in lithic assemblages at non- thermal-feature mapping units: of 936 lithic-scatter mapping units, 375 disclosed bifacial tools (either whole or fragmentary bifaces or projectile points) and 186 disclosed ground- stone artifacts (either whole or fragmentary manos or metates); of 224 sherd-and-lithic scatter mapping units, 92 disclosed bifacial tools (either whole or fragmentary bifaces or projectile points) and 82 disclosed ground-stone artifacts (either whole or fragmentary manos or metates).

that sherd-and-lithic scatters are not simply lithic scatters with ceramics, and that lithic scatters are not sherd-and-lithic scatters without ceramics. With this regional perspective, the clear interpretive challenge raised by sherd-and- lithic artifact scatters is to understand why the incorporation of ceramics at

138 Sullivan non-thermal-feature processing locations influenced the deposition of bifa- cial tools.

Cuoncl sions The Arch Lake and Upper Basin studies exemplify the idea that artifact scatters are essential sources of information about ancient behavior, organiza- tion, and identity (Colton and Colton 1918; Seymour 2010). In the American Southwest, where the course of prehistory is often based on “strong patterns” in the archaeological record (e.g., Arakawa et al. 2013; Gregory and Wilcox 2007), consideration of artifact scatter variability matters because an “accurate prehistory must look beyond the salient features and take into account the people—and the time periods—for whom conspicuous characteristics were not the norm” (Tainter and Plog 1994:166). From an epistemological perspec- tive, therefore, archaeologists can ill afford not to consider the effects of alter- native factors that affect the content and interpretive potential of their arti- fact assemblages; to do otherwise promotes the construction of views of the cultural past that may be ontologically spurious. For instance, terminological ambiguity may have influenced the variables that were involved in measuring the alleged shift from “curated” to “expedient core” technologies. Because the term “flake core” (Parry andK elly 1987:292) is undefined, it could refer to cores in the technical sense (i.e., lithic artifacts that disclose negative force features only) or to flakes from which other flakes have been removed (in which case these artifacts are retouched pieces because they express both positive and negative force features; see Note 3 and Hiscock 2007:210–213). If the term was intended to convey the latter sense, which seems likely, then biface:core ratios are lower than they should be because retouched pieces are being counted as cores, thereby making assemblages look “expedient” (see also Railey 2010:282; Young 1994:148–149). Hence, the hypothesized post-sedentism shift from standardized core technology and “formal tool” manufacture (associated with “curated” technologies) to expedient core technology may never have been as widespread as originally represented, if it occurred at all. Finally, in thinking about the outcomes of examining the Adaptive Diversity and Expedient Core hypotheses, it seems clear that without jointly considering debitage assemblage and tool assemblage data from archaeologi- cal sites that express ranges of occupational and functional variability—from sub-quotidian, bifacial-tool-production locales to multiyear, multiactivity perennial settlements—inferences regarding the relation between mobility and the organization of lithic technology are likely to be skewed, and perhaps

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 139 dramatically so (Olszewski 2007). This issue is particularly relevant for prob- lems involving archaeological phenomena that arose in circumstances where humans controlled the distribution, abundance, and predictability of edible resources. With this facultative ecological perspective, there are theoretically warranted grounds to postulate that the organization of lithic technology in anthropogenic landscapes, particularly those that featured the production and processing of ruderals (Sullivan 2015), was based on problem-solving strate- gies that differ profoundly from those developed in response to resource dis- tributions in non-anthropogenic ecosystems. Resolving what those strategies might have been will promote better understandings of the origins and his- tories of regional archaeological records, especially with regard to theorizing how the “materiality of livelihood” comes to be registered in thin, highly vari- able surface scatters of lithic artifacts.

Acknowledgments The Upper Basin Archaeological Research Project has been supported by grants from the University Research Council (University of Cincinnati), C. P. Taft Research Center (University of Cincinnati), the McMicken College of Arts and Sciences (University of Cincinnati), the , the National Geographic Society/Waitt Grant Program, and the USDA Forest Service (Kaibab National Forest). Sissel Schroeder (University of Wisconsin, Madison) and Deb Olszewski (University of Pennsylvania) provided incred- ibly useful comments on earlier drafts of this chapter, for which I am most grateful. This chapter is dedicated to Kenneth C. Rozen, my friend and colleague, whose untimely death on July 23, 2010, at the age of 56, shocked his vast network of friends, particularly those in the Arizona archaeological com- munity. Ken’s widely cited lithic artifact studies are characterized by jargon- free language, appropriate use of statistics, and clear, concise arguments. In these respects, his AEPCO (Rozen 1979), TEP/St. Johns (Rozen 1981), and ANAMAX/Rosemont (Rozen 1984) investigations are classic contributions to Arizona archaeology, as well as to archaeological method generally. Moreover, whatever impacts our three jointly authored articles (Sullivan and Rozen 1985; Rozen and Sullivan 1989a and 1989b) may have had on the field, they are directly attributable to Ken’s uncompromising attention to detail and to his vision that lithic analysis must be based on objective methods that promote strong assemblage-level behavioral inferences.

140 Sullivan N otes 1. I hesitate to include Robert Ascher and Marcia Ascher’s study of the Leupp site, northern Arizona, which was published in Science in 1965, because it is unclear which lithic artifact class they were investigating (Ascher and Ascher 1965). The artifacts in their figure 2 do not appear to be debitage and are labeled “Typical specimens of the postulated stone-tool industry, collected at the Leupp Site.” Nevertheless, their study is foundational because of its use of sampling methods, integration of experimental and environmental controls (see Rezek et al., chapter 14, this volume), and statistical analysis of measurable percussion features (their “industry angles”). 2. Seven of the 11 Wetherill Mesa sites in Wheeler’s (1965) study also were involved in D. Osborne’s (1965) analysis: Two Raven House, Big Juniper House, Badger House, Long House, Mug House, Step House, and Site 1676. Wheeler also examined material from Adobe House, Site 1205, Site 1230, and Site 1801. 3. These classes of lithic artifacts are distinguished technologically, based on the expression of force (percussion) features. Not uncommonly, definitions of debitage or debris are based on assertions about behavior (e.g., “Detached pieces that are discarded during the reduction process” [Andrefsky 1998:xxii] or “debris and flake debris are used here as collective terms for all waste material generated by humans in lithic reduction” [Shott 1994:70]) and, hence, are equivocal or of little use in differentiating lithic arti- facts empirically—that is, based on their observable properties. Similar problems apply to terms such as informal tools, formal tools, and expedient tools (Andrefsky 1998:xxiii– xxiv; see also Railey 2010:282). In categorizing debitage and retouched pieces, some additional stipulations must be considered (Rozen and Sullivan 1989a:181): “Retouch is defined as negative fea- tures that are at least 3 mm long and that originate from any margin except the edge at the intersection of the platform and the exterior surface, unless flake scars originating from this edge intersect the interior surface. Artifacts with retouch are classified as ‘retouched pieces,’ whereas those on which retouch is absent (those with positive percussion features only, or those with positive features and negative features that originate at the platform and extend across only the exterior surface) are clas- sified as debitage.” 4. Complete flake “length is the distance between the point of impact of the detaching blow and the maximum extent of the flaking axis on the interior surface. Width is the distance between two lateral edges measured perpendicularly at the mid- point of length. Thickness is the distance between interior and exterior surfaces mea- sured at the intersection of length and width” (Sullivan 1995:54). 5. Kruskal-Wallis nonparametric tests indicate statistically significant differences among the five classes of sites with respect to mean values of relative thickness (Chi- square = 13.57, df = 4, p = .009) and weight (Chi-square = 12.58, df = 4, p = .014).

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 141 References Adler, Michael A. 1992. “The Upland Survey.” In The Sand Canyon Archaeological Project: A Progress Report, ed. William D. Lipe, 11–23. Occasional Paper No. 2. Cortez: Crow Canyon Archaeological Center. Andrefsky, William, Jr. 1998. Lithics: Macroscopic Approaches to Analysis. Cambridge: Cambridge University Press. Andrefsky, William, Jr. 2008a. “An Introduction to Stone Tool Life History and Technological Organization.” In Lithic Technology: Measures of Production, Use, and Curation, ed. William Andrefsky, Jr., 3–22. Cambridge: Cambridge University Press. http://dx.doi.org/10.1017/CBO9780511499661.002. Andrefsky, William, Jr. 2008b. “Projectile Point Provisioning Strategies and Human Land Use.” In Lithic Technology: Measures of Production, Use, and Curation, ed. William Andrefsky, Jr., 195–215. Cambridge: Cambridge University Press. http://dx .doi.org/10.1017/CBO9780511499661.010. Arakawa, Fumiyasu, Christopher Nicholson, and Jeff Rasic. 2013. “The Consequences of Social Processes: Aggregate Populations, Projectile Point Accumulation, and Subsistence Patterns in the American Southwest.” American Antiquity 78(1):147– 165. http://dx.doi.org/10.7183/0002-7316.78.1.147. Ascher, Robert, and Marcia Ascher. 1965. “Recognizing the Emergence of Man.” Science 147(3655):243–250. http://dx.doi.org/10.1126/science.147.3655.243. Bailey, Geoff. 2007. “Time Perspectives, Palimpsests, and the Archaeology of Time.” Journal of Anthropological Archaeology 26(2):198–223. http://dx.doi.org/10.1016/j .jaa.2006.08.002. Barton, C. Michael, and Julien Riel-Salvatore. 2014. “The Formation of Lithic Assemblages.” Journal of Archaeological Science 46:334–352. http://dx.doi.org/10 .1016/j.jas.2014.03.031. Binford, Lewis R. 1979. “Organization and Formation Processes: Looking at Curated Technologies.” Journal of Anthropological Research 3:255–373. Binford, Lewis R. 1980. “Willow Smoke and Dogs’ Tails: Hunter-Gatherer Settlement Systems and Archaeological Site Formation.” American Antiquity 45(1):4–20. http://dx.doi.org/10.2307/279653. Binford, Lewis R. 1986. “An Alyawara Day: Making Men’s Knives and Beyond.” American Antiquity 51(3):547–562. http://dx.doi.org/10.2307/281751. Brown, J. William. 1969. “Stratigraphy and Petrology of the Kaibab Formation between Desert View and Cameron, Northern Arizona.” In Geology and Natural History of the Grand Canyon Region, 170–173. Durango, CO: Four Corners Geological Society. Carter, Sidney W., Bettina Wiegand, Gail A. Mahood, Francis O. Dudas, Joseph L. Wooden, Alan P. Sullivan, III, and Samuel A. Bowring. 2011. “Strontium Isotopic

142 Sullivan Evidence for Prehistoric Transport of Gray-Ware Ceramic Materials in the Eastern Grand Canyon Region, USA.” Geoarchaeology: An International Journal 26(2):189–218. http://dx.doi.org/10.1002/gea.20348. Catlin, Mark. 1986. “Intersite Diversity and the Role of Limited-Activity Sites in Subsistence-Settlement Systems on Black Mesa.” In Spatial Organization and Exchange: Archaeological Survey on Northern Black Mesa, ed. Stephen Plog, 169–186. Carbondale: Southern Illinois University Press. Chartkoff, Joseph L. 1995. “A Nested Hierarchy of Contexts: An Approach to Defining Significance for Lithic Scatters.” Journal of California and Great Basin Anthropology 17:28–40. Clarkson, Chris. 2008. “Changing Reduction Intensity, Settlement, and Subsistence in Wardaman Country, Northern Australia.” In Lithic Technology: Measures of Production, Use, and Curation, ed. William Andrefsky, Jr., 286–316. Cambridge: Cambridge University Press. http://dx.doi.org/10.1017/CBO978 0511499661.014. Colton, Mary R. F., and Harold S. Colton. 1918. “The LittleK nown Small House Ruins in the Coconino Forest.” Memoirs of the American Anthropological Association 5:101–126. Cook, Robert A. 1995. “Long-Term Upland Wild-Resource Subsistence Technology: Evidence from Fire-Cracked-Rock Piles in the Upper Basin, Kaibab National Forest, Northern Arizona.” Master’s thesis, University of Cincinnati, Cincinnati, OH. Eren, Metin I., and Brian N. Andrews. 2013. “Were Bifaces Used as Mobile Cores for Clovis Foragers in the North American Lower Great Lakes Region? An Archaeological Test of Experimentally Derived Quantitative Predictions.” American Antiquity 78(1):166–180. http://dx.doi.org/10.7183/0002-7316.78.1.166. Eren, Metin I., Manuel Dominguez-Rodrigo, Steven L. Kuhn, Daniel S. Adler, Ian Le, and Ofer Bar-Yosef. 2005. “Defining the Measuring Reduction in Unifacial Stone Tools.” Journal of Archaeological Science 32(8):1190–1201. http://dx.doi.org/10 .1016/j.jas.2005.03.003. Eren, Metin I., and Mary E. Prendergast. 2008. “Comparing and Synthesizing Unifacial Stone Tool Reduction Indices.” In Lithic Technology: Measures of Production, Use, and Curation, ed. William Andrefsky, Jr., 49–85. Cambridge: Cambridge University Press. http://dx.doi.org/10.1017/CBO9780511499661.004. Fanning, Patricia C., Simon J. Holdaway, Ed J. Rhodes, and Tessa Bryant. 2009. “The Surface Archaeological Record in Arid Australia: Geomorphic Controls on Preservation, Exposure and Visibility.” Geoarchaeology: An International Journal 24(2):121–146. http://dx.doi.org/10.1002/gea.20259.

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 143 Frison, George C. 1968. “A Functional Analysis of Certain Chipped Stone Tools.” American Antiquity 33(2):149–155. http://dx.doi.org/10.2307/278516. Greenberg, April R. 2013. “A GIS-based Spatial Analysis of Factors that Influenced the Placement of Fire-Cracked-Rock Features in the Upper Basin, Northern Arizona.” Master’s thesis, University of Cincinnati, Cincinnati, OH. Gregory, David A., and David R. Wilcox, eds. 2007. Zuni Origins: Toward a New Synthesis of Southwestern Archaeology. Tucson: University of Arizona Press. Gumerman, George J. 1984. A View from Black Mesa: The Changing Face of Archaeology. Tucson: University of Arizona Press. Haury, Emil W. 1950. The Stratigraphy and Archaeology of . Tucson: University of Arizona Press. Hiscock, Peter. 2007. “Looking the Other Way: A Materialist/Technological Approach to Classifying Tools and Implements, Cores and Retouched Flakes.” In Tools versus Cores: Alternative Approaches to Stone Tool Analysis, ed. Shannon P. McPherron, 198–222. Newcastle: Cambridge Scholars Publishing. Jelinek, Arthur J. 1966. “Some Distinctive Flakes and Flake Tools from the Llano Estacado.” Papers of the Michigan Academy of Science, Arts, and Letters 51:399–405. Jelinek, Arthur J. 1967. A Prehistoric Sequence in the Middle Pecos Valley, New Mexico. Anthropological Papers No. 31. Ann Arbor: Museum of Anthropology, University of Michigan. Kelly, Robert L. 1983. “Hunter-Gatherer Mobility Strategies.” Journal of Anthropological Research 39:277–306. Kelly, Robert L. 1992. “Mobility/Sedentism: Concepts, Archaeological Measures, and Effects.” Annual Review of Anthropology 21(1):43–66. http://dx.doi.org/10.1146 /annurev.an.21.100192.000355. Key, Alastair J.M., and Stephen J. Lycett. 2014. “Are Bigger Flakes Always Better? An Experimental Assessment of Flake Size Variation on Cutting Efficiency and Loading.” Journal of Archaeological Science 41:140–146. http://dx.doi.org/10.1016/j.jas .2013.07.033. Kohler, Timothy A. 1993. “News from the North American Southwest: Prehistory on the Edge of Chaos.” Journal of Archaeological Research 1(4):267–321. http://dx.doi .org/10.1007/BF01418108. Kuhn, Steven L. 1991. “‘Unpacking’ Reduction: Lithic Raw Material Economy in the Mousterian of West-Central Italy.” Journal of Anthropological Archaeology 10(1):76– 106. http://dx.doi.org/10.1016/0278-4165(91)90022-P. Lange, Richard C. 1998. Prehistoric Land-Use and Settlement of the Middle Little Colorado River Valley: The Survey of Homolovi Ruins State Park, Winslow, Arizona.

144 Sullivan Arizona State Museum Archaeological Series No. 189. Tucson: University of Arizona. Lesko, Lawrence M. 1989. “A Reexamination of Northern Arizona Obsidian.” Kiva 54:385–399. Longacre, William A. 1967. “Artifacts.” In Chapters in the Prehistory of Eastern Arizona, III, ed. Paul S. Martin, William A. Longacre, and James N. Hill, 56–125. Fieldiana: Anthropology, Volume 57. Chicago: Field Museum of Natural History. Longacre, William A., and Michael W. Graves. 1976. “Probability Sampling Applied to an Early Multi-Component Surface Site in East-Central Arizona.” Kiva 41:277–288. Magne, Martin R. R. 1989. “Lithic Reduction Stages and Assemblage Formation Processes.” In Experiments in Lithic Technology, ed. Daniel S. Amick and Raymond P. Mauldin, 15–31. BAR International Series 528. Oxford: British Archaeological Reports. McCall, Grant S. 2012. “Ethnoarchaeology and the Organization of Lithic Technology.” Journal of Archaeological Research 20(2):157–203. http://dx.doi.org /10.1007/s10814-011-9056-z. Mehalic, David S. 2012. “The Archaeological Geography of Small Architectural Sites of the Mogollon Plateau Region of East-Central Arizona.” PhD dissertation, Department of Anthropology, University of Arizona, Tucson. Nelson, Margaret C. 1991. “The Study of Technological Organization.” In Archaeological Method and Theory, vol. 3. ed. Michael B. Schiffer, 57–100. Tucson: University of Arizona Press. Olszewski, Deborah I. 2007. “Carinated Tools, Cores, and Mobility: The Zagros Aurigancian Example.” In Tools versus Cores: Alternative Approaches to Stone Tool Analysis, ed. Shannon P. McPherron, 91–106. Newcastle: Cambridge Scholars Publishing. Osborne, Carolyn M. 1965. “The Preparation of Yucca Fibers: An Experimental Study.” In Contributions of the Wetherill Mesa Archeological Project, ed. Douglas Osborne, 45–56. Memoirs of the Society for American Archaeology No. 19. Salt Lake City: Society for American Archaeology. Osborne, Douglas. 1965. “Chipping Remains as an Indication of Cultural Change at Wetherill Mesa.” In Contributions of the Wetherill Mesa Archeological Project, ed. Douglas Osborne, 30–44. Memoirs of the Society for American Archaeology No. 19. Salt Lake City: Society for American Archaeology. Parry, William J., and Robert L. Kelly. 1987. “Expedient Core Technology and Sedentism.” In The Organization of Core Technology, ed. Jay K. Johnson and Carol A. Morrow, 285–305. Boulder: Westview Press.

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 145 Plog, Stephen. 1986. “Group Mobility and Locational Strategies: Tests of Some Settlement Hypotheses.” In Spatial Organization and Exchange: Archaeological Survey on Northern Black Mesa, ed. Stephen Plog, 187–223. Carbondale: Southern Illinois University Press. Powell, Shirley, and Anthony L. Klesert. 1984. “A Method for Predicting the Presence of Buried Structures on Unexcavated Artifact Scatters.” In Papers on the Archaeology of Black Mesa, Arizona, vol. II. ed. Stephen Plog and Shirley Powell, 39–46. Carbondale: Southern Illinois University Press. Purtill, Matthew P. 1995. “Analysis and Interpretation of Chipped-Stone Tool Assemblages from the Upper Basin, Kaibab National Forest, Northern Arizona.” Master’s thesis, University of Cincinnati, Cincinnati, OH. Railey, Jim A. 2010. “Reduced Mobility or the Bow and Arrow? Another Look at “Expedient” Technologies and Sedentism.” American Antiquity 75(2):259–286. http://dx.doi.org/10.7183/0002-7316.75.2.259. Rinaldo, John B. 1964. “Artifacts.” In Chapters in the Prehistory of Eastern Arizona, II, ed. Paul S. Martin, John B. Rinaldo, William A. Longacre, Leslie G. Freeman, Jr., James A. Brown, Richard H. Hevly, and M. E. Cooley, 63–109. Fieldiana: Anthropology, Volume 55. Chicago: Field Museum of Natural History. Roberts, Frank H. H. 1931. The Ruins at Kiatuthlanna, Eastern Arizona. Bureau of American Ethnology Bulletin 100. Washington, DC: Smithsonian Institution. Roberts, Frank H. H. 1940. Archeological Remains in the Whitewater District, Eastern Arizona. Bureau of American Ethnology Bulletin 126. Washington, DC: Smithsonian Institution. Rozen, Kenneth C. 1979. “Lithic Analysis and Interpretation.” In The AEPCO Project, Vol. 2, ed. Deborah A. Westfall, Kenneth C. Rozen, and Howard M. Davidson, 209–321. Arizona State Museum Archaeological Series No. 117. Tucson: University of Arizona. Rozen, Kenneth C. 1981. “Patterned Associations among Lithic Technology, Site Content, and Time.” In Prehistory of the St. Johns Area, East-Central Arizona: The TEP St. Johns Project, ed. Deborah A. Westfall, 157–232. Arizona State Museum Archaeological Series No. 153. Tucson: University of Arizona. Rozen, Kenneth C. 1984. “Flaked Stone.” In Hohokam Habitation Sites in the Northern Santa Rita Mountains, by A. Ferg, K. C. Rozen, W. L. Deaver, M. A. Tagg, D. A. Phillips Jr., and D. A. Gregory, 421–604. Arizona State Museum Archaeological Series No. 147 (2). Tucson: University of Arizona. Rozen, Kenneth C., and Alan P. Sullivan, III. 1989a. “The Nature of Lithic Reduction and Lithic Analysis: Stage Typologies Revisited.” American Antiquity 54(1):179–184. http://dx.doi.org/10.2307/281342.

146 Sullivan Rozen, Kenneth C., and Alan P. Sullivan, III. 1989b. “Measurement, Method, and Meaning in Lithic Analysis: Problems with Amick and Mauldin’s Middle Range Approach.” American Antiquity 54(1):169–175. http://dx.doi.org/10.2307/281340. Rushforth, Scott, and Steadman Upham. 1992. A Hopi Social History. Austin: University of Texas Press. Seymour, Deni J. 2010. “Contextual Incongruities, Statistical Outliers, and Anomalies: Targeting Inconspicuous Occupational Events.” American Antiquity 75(1):158–176. http://dx.doi.org/10.7183/0002-7316.75.1.158. Shackley, M. Steven. 2005. Obsidian: Geology and Archaeology in the North American Southwest. Tucson: University of Arizona Press. Shott, Michael J. 1986. “Technological Organization and Settlement Mobility: An Ethnographic Examination.” Journal of Anthropological Research 42:15–50. Shott, Michael J. 1994. “Size and Form in the Analysis of Flake Debris: Review and Recent Approaches.” Journal of Archaeological Method and Theory 1(1):69–110. http://dx.doi.org/10.1007/BF02229424. Shott, Michael J. 1995. “How Much is a Scraper? Curation, Use Rates, and the Formation of Scraper Assemblages.” Lithic Technology 20:52–72. Shott, Michael J., and Margaret C. Nelson. 2008. “Lithic Reduction, Its Measurement, and Implications: Comments on the Volume.” In Lithic Technology: Measures of Production, Use, and Curation, ed. William Andrefsky, Jr., 23–45. Cambridge: Cambridge University Press. http://dx.doi.org/10.1017/CBO9780 511499661.003. Smith, Bruce D. 2011. “Shaping the Natural World: Patterns of Human Niche Construction by Small-Scale Societies in North America.” In The Subsistence Economies of Indigenous North American Societies, ed. Bruce D. Smith, 593–609. Washington, DC: Smithsonian Institution Press. Spurr, Kimberly, Phil R. Geib, and Jim H. Collette. 2004. “Patterns of Human Activity in ‘The Heart of the Desert Wild’: Archaeological Survey and Testing on the Kaiparowits Plateau, Grand Staircase-Escalante National Monument.” In The Colorado Plateau: Cultural, Biological, and Physical Research, ed. Charles van Ripper, III, and Kenneth L. Cole, 19–38. Tucson: University of Arizona Press. Spurr, Kimberley, and L. Theodore Neff. 2005. “Archaeological Survey of New Land Acquired by Walnut Canyon National Monument, Northern Arizona.” In The Colorado Plateau II: Biophysical, Socioeconomic, and Cultural Research, ed. Charles van Ripper, III, and David J. Mattson, 349–365. Tucson: University of Arizona Press. Sullivan, Alan P., III. 1983. “Storage, Nonedible Resource Processing, and the Interpretation of Sherd and Lithic Scatters in the Sonoran Desert Lowlands.” Journal of Field Archaeology 10:309–325.

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 147 Sullivan, Alan P., III. 1987. “Artifact Scatters, Adaptive Diversity, and Southwestern Abandonment: The Upham Hypothesis Reconsidered.” Journal of Anthropological Research 43:345–360. Sullivan, Alan P., III. 1988. “Prehistoric Southwestern Ceramic Manufacture: The Limitations of Current Evidence.” American Antiquity 53(1):23–35. http://dx.doi.org /10.2307/281152. Sullivan, Alan P., III. 1995. “Artifact Scatters and Subsistence Organization.” Journal of Field Archaeology 22(1):49–64. http://dx.doi.org/10.1179/00934699579154 7705. Sullivan, Alan P., III. 2001. “Holmes’s Principle and Beyond: The Case for Renewing Americanist Debitage Analysis.” In Lithic Debitage Analysis: Studies in Context, Form, and Meaning, ed. William Andrefsky, Jr., 192–206. Salt Lake City: University of Utah Press. Sullivan, Alan P., III. 2015. “The Archaeology of Ruderal Agriculture.” In Traditional Arid Lands Agriculture: Understanding the Past for the Future, ed. Scott E. Ingram and Robert C. Hunt, 273–305. Tucson: University of Arizona Press. Sullivan, Alan P., III, Robert A. Cook, Matthew P. Purtill, and Patrick M. Uphus. 2001. “Economic and Land-Use Implications of Prehistoric Fire-Cracked-Rock Piles, Northern Arizona.” Journal of Field Archaeology 28:367–382. Sullivan, Alan P., III, Philip M. Mink, and Patrick M. Uphus. 2002. “From John W. Powell to Robert C. Euler: Testing Models of Grand Canyon’s Prehistoric Puebloan Settlement History.” In Culture and Environment in the American Southwest: Essays in Honor of Robert C. Euler, ed. David A. Philips, Jr., and John A. Ware, 49–68. Anthropological Research Paper No. 8. Phoenix: SWCA. Sullivan, Alan P., III, and Kenneth C. Rozen. 1985. “Debitage Analysis and Archaeological Interpretation.” American Antiquity 50(4):755–779. http://dx.doi.org /10.2307/280165. Sullivan, Alan P., III, and Anthony H. Ruter. 2006. “The Effects ofE nvironmental Fluctuations on Ancient Livelihood: Implications of Paleoeconomic Data from the Upper Basin, Northern Arizona.” In Environmental Change and Human Adaptation in the Ancient American Southwest, ed. David E. Doyel and Jeffrey S. Dean, 180–203. Salt Lake City: University of Utah Press. Surovell, Todd A. 2009. Toward a Behavioral Ecology of Lithic Technology: Cases from Paleoindian Archaeology. Tucson: University of Arizona Press. Szeghi, Shelley A. 2012. “Spatial Distribution and Assemblage Composition Patterns of Sherd-and-Lithic Artifact Scatters in the Upper Basin, Northern Arizona.” Master’s thesis, University of Cincinnati, Cincinnati, OH.

148 Sullivan Tainter, Joseph A. 1979. “The Mountainair Lithic Scatters: Settlement Patterns and Significance Evaluation of Low Density Surface Sites.” Journal of Field Archaeology 6:463–469. Tainter, Joseph A., and Bonnie Bagley. 2005. “Shaping and Suppressing the Archaeological Record.” In Heritage of Value, Archaeology of Renown: Reshaping Archaeological Assessment and Significance, ed. Clay Mathers, Timothy Darvill, and Barbara J. Little, 58–73. Gainesville: University Press of Florida. Tainter, Joseph A., and Fred Plog. 1994. “Strong and Weak Patterning in Southwestern Prehistory: The Formation of Puebloan Archaeology.” In Themes in Southwest Prehistory, ed. George J. Gumerman, 165–181. Santa Fe: School of American Research Press. Thompson, Nicholas C. 2003. “From Data Recovery to Data Analysis: Projectile Points of the Upper Basin, Kaibab National Forest, Northern Arizona.” Master’s thesis, University of Cincinnati, Cincinnati, OH. Tomka, Steve A. 2001. “The Effect of Processing Requirements on Reduction Strategies and Tool Form: A New Perspective.” In Lithic Debitage: Context, Form, Meaning, ed. William Andrefsky, Jr., 207–223. Salt Lake City: University of Utah Press. Torres, John A. 2000. “Changing Lithic Technology during the Basketmaker-Pueblo Transition: Evidence from the Anasazi Heartland.” In Foundations of Anasazi Culture: The Basketmaker-Pueblo Transition, ed. Paul F. Reed, 221–229. Salt Lake City: University of Utah Press. Upham, Steadman. 1984. “Adaptive Diversity and Southwestern Abandonment.” Journal of Anthropological Research 40:235–256. Upham, Steadman. 1994. “Nomads of the Desert West: A Shifting Continuum in Prehistory.” Journal of World Prehistory 8(2):113–167. http://dx.doi.org/10.1007/BF 02220562. Wait, Walter K. 1983. “Alternate Approaches to the Analysis of Low-Density Artifact Scatters.” In The Star Lake Archaeological Project: Anthropology of a Headwaters Area of Chaco Wash, New Mexico, ed. Walter K. Wait and Ben A. Nelson, 59–94. Carbondale: Southern Illinois University Press. Wandsnider, LuAnn, and Eileen L. Camilli. 1996. “Land-Use Histories from the Mesilla Bolson (South-Central New Mexico): Evidence from Spatial Patterning in Surface Artifact Distributions.” In Interpreting Southwestern Diversity: Underlying Principles and Overarching Patterns, ed. Paul R. Fish and J. Jefferson Reid, 211–239. Anthropological Research Papers No. 48. Tempe: Arizona State University. Whalen, Michael E. 1986. “Small-Site Analysis in the Hueco Bolson of Western Texas.” Journal of Field Archaeology 13(1):69–81.

Theoretical Implications of Artifact-Scatter Lithic Assemblage Variability 149 Wheeler, Richard P. 1965. “Edge-Abraded Flakes, Blades, and Cores in the Puebloan Tool Assemblage.” In Contributions of the Wetherill Mesa Archeological Project, ed. Douglas Osborne, 19–29. Memoirs of the Society for American Archaeology No. 19. Salt Lake City: Society for American Archaeology. Williams, Justin P., and William Andrefsky, Jr. 2011. “Debitage Variability among Multiple Flint Knappers.” Journal of Archaeological Science 38(4):865–872. http://dx .doi.org/10.1016/j.jas.2010.11.008. Wilson, Jennifer, and William Andrefsky, Jr. 2008. “Exploring Retouch on Bifaces: Unpacking Production, Resharpening, and Hammer Type.” In Lithic Technology: Measures of Production, Use, and Curation, ed. William Andrefsky, Jr., 86–105. Cambridge: Cambridge University Press. http://dx.doi.org/10.1017/CBO 9780511499661.005. Young, Lisa C. 1994. “Lithics and Adaptive Diversity: An Examination of Limited- Activity Sites in Northeast Arizona.” Journal of Anthropological Research 50:141–167.

150 Sullivan Section II Venerable Sites Revisited

7

Across eastern North America in the nineteenth cen- Timelessness and the tury, individuals with training in engineering and Legacy of Archaeological natural history and a keen interest in antiquities were Cartography directly involved in field surveys of earthen monu- ments. The maps they created from these surveys were developed with a false understanding of a shallow Sissel Schroeder and time depth predating the European occupation of the Lynne Goldstein region and crafted with an emphasis on descriptive elements (e.g., Squier and Davis 1848; Thomas 1894). These nineteenth-century depictions of mound sites helped to establish a timeless view of the past that persists today in some interpretations of the physical layout of mound sites, and many of these sites continue to be depicted in maps and artists’ renderings as the enduring sum of their parts—earthen mounds, plazas, and other structures. As an example of the hegemony of archaeological cartography, we scrutinize the history of interpreta- tions of site organization at Aztalan, a mound center encircled by wooden palisades and situated in southern Wisconsin (figure 7.1), which has been persistently per- ceived and described in terms of a timeless picture of the past (e.g., Barrett 1933; Birmingham and Goldstein 2005; Hyer 1837, 1838; Lapham 1855; Lewis 1894; McKern 1946; Richards 2007). The accumulated data from more than a century of excavations at Aztalan have con- firmed two major archaeological traditions at the site: a Late Woodland presence (ca. ad 800/900–1150/1200) that is evident in some of the ceramics and a Middle DOI: 10.5876/9781607324942.c007

153 Figure 7.1. Location of Aztalan in southern Wisconsin.

Mississippian occupation (ca. ad 1050/1100–1200/1250) that is substantiated by some of the pottery, other forms of material culture, platform mounds, palisades, and various architectural features (Birmingham and Goldstein 2005; Richards 2003; Richards and Jeske 2002). The joint Late Woodland and Mississippian occupation history, which likely spans fewer than two centuries, and a coarse-grained ceramic chronology (with the exception of Powell Plain and Ramey Incised types from Cahokia) that has been challenging to parse into temporal components shorter than the life span of the site, have further contributed to the timeless perception of Aztalan. Drawing on interpretive

154 Schroeder & Goldstein frameworks that make use of time perspectivism (Bailey 2007; Sullivan 2008), migration, coalescence, and hybridity (e.g., Alt 2006; Anthony 1990; Bhabha 1985; Ethridge and Hudson 2002; Kowalewski 2001; Liebmann 2013), and on comparisons with some other stockaded Mississippian sites in the southeast- ern United States, we propose that Aztalan’s site structure developed rapidly through a process of population coalescence. Among the material manifesta- tions of the development of a blended community was the interweaving of architectural conventions with novel ideas that led to recurrent redefinitions, reconstructions, and modifications of place and space.

Antiquarian Encounters with Aztalan In 1836, a man named Nathanial Hyer heard rumors of an “ancient Walled City” about three days travel west of Milwaukee in what was then known as the Wisconsin Territory. After one failed attempt to reach the site in the spring, Hyer succeeded in finding the ruins in October and prepared a crude sketch map of the site that he copied and shared widely among his friends (Hyer 1837). When he saw the earthen pyramids within the “ancient Walled City,” he was reminded of an engraving of the Toltec pyramid at Cholula that he had seen in Baron Alexander von Humboldt’s published account of his travels in Mexico and of descriptions of Aztec pyramids contained elsewhere within the volumes (von Humboldt 1814:I:81; von Humboldt 1814:I and II; see Richards 2007 for a more detailed historical chronology of documents that mention Aztalan). Von Humboldt asserted, “we must look for the first country of the Mexican nations, Aztlan . . . at least north of the 42d degree of latitude” (von Humboldt 1814:II:66). As Hyer’s “ancient Walled City” lies at 43 degrees north, he borrowed the name of the mythical origin of the Aztec empire, thus conferring the name Aztalan on this site. He returned to the site in early 1837 with surveying equipment, which allowed him to produce a map with somewhat greater detail (figure 7.2; Hyer 1837). His descriptions of the earthen platform mounds and walls of so-called brick (baked clay) at Aztalan established an enduring narrative of the site as a unique, timeless, and even mysterious place (Hyer 1837, 1838). A little more than a dozen years following the appearance of Hyer’s map of Aztalan, Increase Lapham, a naturalist and professionally trained engineer who had moved to Milwaukee from New York State in 1836, resurveyed Aztalan and produced a more accurate and detailed map of the site (figure 7.3; Lapham 1855). Without question, the focus of these early explorations was on the visible earthen architecture of the site, particularly the platform mounds and walls.

Timelessness and the Legacy of Archaeological Cartography 155 Figure 7.2. 1837 map of Aztalan by Nathaniel Hyer (Wisconsin Historical Society Image ID 53061).

In 1919, Samuel Barrett of the Milwaukee Public Museum initiated exca- vations into portions of the platforms mounds, the walls, and a knoll in the southeast corner of the site, and conducted some limited excavations of the interior of the site (Barrett 1933). Only a decade earlier, William Henry Holmes’s classification of ceramics in the Mississippi Valley had appeared in print (Holmes 1903), and Barrett recognized that some of the pottery that was being recovered from Aztalan was of the Woodland type,while other pieces were similar to the Middle Mississippian type, which was distinguished on the basis of shell temper, smooth exterior surfaces that were sometimes painted or polished or incised, and certain vessel forms. Barrett also noted that the shell- tempered sherds and vessels from Aztalan closely resembled materials from the site of Cahokia near St. Louis (Barrett 1933).

T he Salient Elements of a Mississippian Site Middle Mississippian sites are concentrated along major waterways and tributary drainages in the southeastern United States (Smith 1986), across

156 Schroeder & Goldstein Figure 7.3. 1850 map of Aztalan by Increase Lapham (Lapham 1855; reprinted with the permission of the University of Wisconsin Press). southwestern Indiana, southern and western Illinois, and into Missouri, but small numbers of them are found along the Mississippi River as far north as Red Wing, Minnesota, and along the Rock and Crawfish Rivers in southern Wisconsin (Schroeder 2004: figure 4). Aztalan, in particular, and the other northern Mississippian sites in Minnesota, Wisconsin, and Iowa have long been viewed as anomalous in their locations, consequences of the migration of people whose origins are commonly attributed to Cahokia (Benden 2004; Finney 2013; Finney and Stoltman 1991; Gibbon and Dobbs 1991; Goldstein 1991; Green and Rodell 1994; Pauketat et al. 2015; Rodell 1991), which is by far the largest of all Mississippian sites and is situated in a broad expanse of the floodplain of the Mississippi River across from St. Louis in present-day Illinois (e.g., Iseminger 2010; Milner 1998, Pauketat 1994, 2004). The atypical location of Aztalan along with the allure of Hyer’s account of Aztalan, which captured the imagination of the public and scholars alike, hindered the rec- ognition that the site was composed of the same basic building blocks that Phillips, Ford, and Griffin listed in 1951 as the traits of a Mississippian site: platform mounds, one or more plazas, and walls, as well as residential and

Timelessness and the Legacy of Archaeological Cartography 157 Figure 7.4. Architectural elements of a Mississippian town and mound center (from Lewis and Stout 1998: Figure 1.2; reprinted with permission of the University of Alabama Press).

ritual structures (Goldstein and Freeman 1997:223; Phillips, Ford, and Griffin 1951; see also contributions to Lewis and Stout 1998). In the horizontal dimension, mounds generally are orthogonally situated along the edges of a centrally located plaza that is usually rectilinear in shape (figure 7.4). One mound in the form of a truncated pyramid is typically larger than others and often is located along the edge of the plaza or, less frequently, in its center. Houses built of perishable materials were once distributed around the perimeter of these mound-and-plaza configurations, although not as densely as presented in figure 7.4, and the boundaries of the town often were demarcated by wooden walls studded with bastions. In the vertical dimension, there is a separation between mound summits, which are associated with elites and rituals, and ground level, where quotidian practices unfolded.

Explaining Site Structure the Old Fashioned Way For decades, archaeologists have sought explanations for the seeming unifor- mity across archaeological cultural and linguistic boundaries in the Southeast

158 Schroeder & Goldstein for how these common elements were arranged to form Mississippian town- and-mound communities that served as the administrative centers of their respective geospatial polities (e.g., Lewis and Stout 1998; Lewis et al. 1998). Under the assumption that all archaeologically visible features at a site were present and in use at the same time, the explanations for the recurrent arrange- ment of these elements range from a shared architectural grammar, to celestial geometry or a physical expression of cosmology, to ritual proscription, to cen- tralized authority within a society characterized by distinct socioeconomic sta- tus differences, to a form of factional competition or political aggrandizement, and to the now-discredited idea of influences from Mexico (e.g., Anderson 1994a; Hall 2004; Kelly and Brown 2014; Knight 1985, 1989; Nassaney 1992; Payne and Scarry 1998; Sherrod and Rolingson 1987; Stout 1984). The absence of a consensus derives from the ambiguity and incompleteness of archaeologi- cal evidence, the diverse theoretical and inferential frameworks that guide the research of different scholars, and variation in how archaeologists think about time as it applies to the spatial organization of a site (see Wilcox, chapter 9, this volume).

“The Loss of Innocence”1 If we start with the assumption that the visible features of the site were not all synchronous, that leads to questions like, “How much of a site was simultaneously active?” “At what point in its history did the site take on its final appearance?” and “How was this final appearance attained?” The effort to address these questions reorients our attention to the formation histories of sites and has the potential to significantly alter our understanding of the temporality of site structure. With the emergence of a new time perspective, some scholars have shifted their inferential frameworks to embrace variability rather than suppress it, and have begun to assemble empirical evidence that challenges theoretical and interpretive orthodoxy, thereby contributing to a dynamic view of the developmental history of cultural landscapes and their formation. For exam- ple, a Bayesian model of radiocarbon dates from Monks Mound has been used to argue that this prominent landscape feature was not part of the ini- tial Mississippianization of Cahokia (Schilling 2010, 2012), a finding that has important implications for the various models of the origins of that site (e.g., Milner 1998; Pauketat 1994). A number of scholars working at Cahokia have made the point that the initial occupation of the site was strung out east- west along the banks of Cahokia Creek, while during its peak florescence as a

Timelessness and the Legacy of Archaeological Cartography 159 Figure 7.5. Samuel Barrett’s map of Aztalan (Barrett 1933; reprinted with permission of the Milwaukee Public Museum).

political center the occupation included a significant expansion to the south and a smaller expansion to the north; later, as the site began to go into decline, areas that had been occupied were abandoned (Beck et al. 2007:842–843; Milner 1998). Elsewhere in the Southeast, major mound sites like Moundville and Etowah each went through their own historical trajectory of shifting site organization and occupation (e.g., King 2001, 2003; Knight 2010; Knight and Steponaitis 1998).

T he Orthodox View of Aztalan Aztalan provides a case example of the influential yet problematic legacy of nineteenth-century archaeological cartography. For instance, Samuel Barrett of the Milwaukee Public Museum approached his excavations at Aztalan with a focus on tracing the numerous walls that were visible to early nineteenth- century visitors to the site (figure 7.5). In his 1933 publication, Barrett force- fully argued that the walls were contemporaneous constructions, a claim that aligns with the timeless view presented in the maps of the site produced by Hyer and Lapham. The strongest evidence that Barrett called upon to support his contention of contemporaneity was that residential architecture had been found only in the excavation units placed within the innermost wall (but see

160 Schroeder & Goldstein Goldstein and Freeman 1997 for an alternative explanation of the concentra- tion of residential architecture within the innermost wall). Barrett’s vision of the site quickly became conventional wisdom, and most archaeologists since that time have pivoted their interpretations of Aztalan around the timeless view (Barrett 1933; Birmingham and Goldstein 2005:53– 76; Goldstein and Freeman 1997; Richards 2007; but see Wittry and Barreis 1958:63 for a rare argument in alignment with the time perspectivism view and Goldstein and Gaff 2002:102). The residential space at Aztalan is found at the lower elevations of the site, adjacent to the riverbank and enclosed by a bastioned palisade that segregated most of the population (figure 7.6). It has been suggested that the two closely spaced walls in this area arose as an added defensive measure or from a slight expansion of the residential sec- tor (Birmingham and Goldstein 2005:53, 58). One platform mound was posi- tioned at the north end, and at the southern end of the residential area is the so-called gravel knoll, a landscape feature with a complex history of surface modification involving human-made deposits of gravel, shell, soil, and arti- facts. To the west of this second wall around the residential district, the eleva- tion rises until another wall is reached. This L-shaped enclosed area is referred to as the public plaza. At the northwestern corner of this plaza is an area called the “sculptuary,” where the land was modified to create three tiers in the glacial gravel that underlies much of the site, each of which contained an estimated one to two dozen very large clay-lined pits that may have been used for com- munal storage, mortuary ritual, and/or trash disposal during their life histories (Goldstein 2010). This area of the site would have been visually notable—a mass of light gravel with dark pits distributed across each tier. Beyond the wall that marks the western limits of the plaza are two platform mounds that anchor the northwest and southwest corners of the site and a stretch of land between them that has been designated as the elite precinct (Birmingham and Goldstein 2005:61–63, 65–74). As one step toward a critique of this orthodox view of the internal organiza- tion of the site and the timelessness of its features, we consider the fact that Aztalan has not been completely, or even systematically, investigated through excavation (Goldstein and Gaff 2002). Barrett’s fieldwork efforts focused on following out the stockades and testing the southwest and northeast mounds; the limited explorations he conducted beyond the palisade lines were mostly situated within the innermost wall. Thus, his “residential zone” may be an artifact of where he conducted his excavations. From the late 1940s into the 1960s the Wisconsin Archeological Survey and the Wisconsin Historical Society examined segments of the palisade, conducted excavations into all

Timelessness and the Legacy of Archaeological Cartography 161 Figure 7.6. Artist’s bird’s-eye view of Aztalan (modified from Birmingham and Goldstein 2005: Figure 4.5; original artwork by Eric Paulson; archived at the University of Wisconsin-Milwaukee, Department of Anthropology).

three mounds, and, significantly, excavated numerous units within the inner wall, finding additional evidence of residential use (Barreis and Bryson 1965; Freeman 1986; Hurley 1977; Maher 1958; Maxwell 1952; Ritzenthaler 1961, 1963;

162 Schroeder & Goldstein Rowe 1958; Wittry and Barreis 1958). The take away message here is that the walls, mounds, and residential district were intensively investigated before the advent of modern methods of excavation and recovery and the availability of AMS radiocarbon dating. Beginning in the 1970s, Lynne Goldstein brought modern question-driven field methods to the site, which confirmed the pres- ence of the plaza, revealed considerable landscape engineering in gravel, and added evidence that diverse economic, ritual, and mortuary activities were undertaken in the sculptuary. Nevertheless, the discoveries made through modern excavations at the Aztalan and the older discoveries remain unified because of their general reliance on the timeless view.

An Unorthodox View of Aztalan In the summer of 2013, we collaborated on a field school at Aztalan, along with Donald Gaff from the University of Northern Iowa. These excavations focused on the knoll (often referred to as the “gravel” knoll) in the southeast portion of the site, and an unusual palisade entrance just west of the southwest mound. In the summer of 2015, Schroeder conducted another field school at the site with excavations focused on the residential part of the site and the plaza. Both of these projects provided us with an opportunity to rethink the timeless interpretations about the emergence of the structure of the site. We started this chapter with a discussion of the walls—those mapped by Hyer and Lapham, then so thoroughly excavated by Barrett. At other Mississippian sites with multiple palisades or stockades, these walls were mostly excavated after the 1930s, when chronology was a priority, and it has been systematically shown in these cases that wall construction proceeded sequentially, either rebuilding in place or through the dismantling of one wall as the next one was erected to accommodate a growing population or more tightly enclose a shrinking community (e.g., Anderson 1969; Anderson 1994b; Anderson and Schuldenrein 1985; Black 1967; Blitz 1993; Butler et al. 2011; Cole 1951; Hammerstedt 2005; Krus 2011; Krus et al. 2013; Polhemus 1987; Schroeder 2006). Taking this perspective that the walls were not synchronous allows us the opportunity to explore a possible dynamic formation history of Aztalan and the interplay between conservatism and innovation expressed through its internal features in a manner that complements and augments recent studies of multiethnic communities in the Southeast that formed through processes of migration and coalescence (e.g., Alt 2002, 2006, 2008; Cobb 2005, 2008; Cobb and Butler 2006; Kidder 1998; Pauketat and Alt 2005; Price et al. 2007;

Timelessness and the Legacy of Archaeological Cartography 163 Schroeder 2011). We propose an alternative hypothesis, informed by a consid- eration of time perspectivism, that the migrants who were involved in estab- lishing the community in collaboration with local residents used Mississippian conventions of public architecture and monumentality to visually assert a new political, social, and ethnic order. Archaeologically, we identify the local peo- ple as part of the Late Woodland Tradition, which in Wisconsin included the construction of effigy mounds and occasionally erecting walls of small- diameter saplings (typically < 15 cm) to protect the perimeter of a compact village (Dirst 1988, 1995; Hall 1962; Overstreet 1995; Salkin 2000). In stark contrast, Mississippian peoples constructed walls out of large-diameter posts (typically > 15 cm); these walls were studded with regularly spaced bastions and enclosed a community that included considerable non-residential space (Milner 1999:118–120). Current archaeological studies of migration posit that the moment when a group of people abandon their homeland and relocate to a wholly new land- scape may provide sufficient rupture in practices that it presents a unique opportunity for a group to reexamine conventions and traditions, thereby cre- ating an opening for innovations (e.g., Alt 2002; Cobb 2008). In selecting a new place to settle, people may choose a place without an obvious history, but they might, as was the case with Aztalan, join an established community, one that already has its own “archaeological” and historical past (see Roth, chapter 3, this volume). The circumstances of establishing a blended community will be highly variable given the different expectations of migrants and people already living at the site, the extent to which the past (an “archaeological record”) is open to different interpretations, and the need for immigrants and residents to develop a common understanding of the place itself. In this con- text the blended community develops very rapidly, new “ways of being” mate- rialize quickly, and notions of place and space are (continually) redefined as architectural conventions are interlayered with novel ideas that emerge from the idiosyncratic and unique configurations of the particular people at these places and in these moments (see Kidder 1998; Cobb 2005, 2008; Cobb and Butler 2006 for similar arguments). Building on archaeological studies in the Southeast where this kind of thinking has been successfully applied to similar kinds of archaeological phe- nomena, it could be postulated that early efforts at monumental construction will emphasize symbols of the newcomers. The construction of a massive wall studded with bastions around a community would have been a particularly dis- tinctive, dramatic, and likely early feature on the landscape, visually proclaim- ing the uniqueness of the newcomers to that setting, while simultaneously

164 Schroeder & Goldstein establishing a method to control access. At the same time, platform mound construction may have initiated an internal reorganization of space.

A Brief Review of Some Evidence from Aztalan Now we turn to a consideration of some of the evidence from Aztalan that could be seen to align with these models. The construction of the northeast mound and possibly the modifications to the knoll in the southeast corner of the site may have been early cooperative events in the formation of the blended community. Based on differences in the ceramic assemblages between submound and mound contexts, Tom Zych (2013) has suggested that the northeast mound may have been a communal construction built by the local long-term Late Woodland residents and Mississippian newcomers. In our 2013 fieldwork at Aztalan we found that the knoll, long accepted to be a natural gravel feature and likely used for mortuary purposes (like a mound), had been subjected to considerable modification in prehistory as well as by past excavations, reconstruction in the 1950s when the site became a public park, and current land use. Our excavations at the apex of the knoll resulted in the identification of multiple pits, mounded features, and distinct fill episodes of gravel and soil, and yielded ceramics of Late Woodland and Mississippian types, rolled copper beads and other objects made from copper, and a large number of whole mussel shells interlayered with gravel—an occurrence that is not natural—that may have been used to cap pit features. Furthermore, the light yellowish-colored glacial gravel at Aztalan would have been a novel material to the Mississippian people who came to the site after having lived much of their lives in an alluvial floodplain that lacks gravel deposits. Based on ethnohistoric and ethnographic analogy, archaeologists have suggested that different colors, like red, black, yellow, and white, had deep symbolic sig- nificance to ancient Native Americans (DeBoer 2005; Pursell 2013; Rodning 2010). The choice to use yellow gravel and white shell in the intentional depos- its on the knoll, along with the exposure of gravel on the ground surface in the sculptuary, would have given a highly distinctive visual appearance to the site. Both the northeast mound and the knoll include the new, Mississippian materials, layered upon and amid the old, Late Woodland materials, in depos- its that seem to have been intentional. In the case of one of the features on the knoll, gravel and shell were intentionally deposited in layers, fragments of at least two ceramic vessels of different types were then laid down on top of each other, and finally objects made of copper were placed before the feature appears to have been capped. Prior to the construction of the northeast mound,

Timelessness and the Legacy of Archaeological Cartography 165 a locally made Late Woodland–type jar was inverted and placed into a pit; as the mound was being built, five vessels were placed upside down in a line that corresponds with the central east-west axis of the mound; four of these ves- sels appear to have been placed while the mound was being constructed while the fifth and western-most of the vessels was placed upside down into a pit excavated into the mound (Zych 2013:114–118, 123, 145, 168, 169, 180, 184). Of the five vessels, two are Cahokia Mississippian types, one is an unidentified Mississippian type, one is too poorly preserved to determine the type, and one ceramic vessel is a hybrid variety—a classic Cahokia Mississippian vessel form tempered with grit instead of shell and likely made by a local potter (Zych 2013:116–117, 169).

Coalescence at Aztalan We argue that it is profitable to consider the formation history of Aztalan from the alternative viewpoint of time perspectivism. The initial calculus of labor investment in monumental architecture may have included conserva- tion of labor to ensure that wall and mound construction would be successful. Early in its establishment, the northeast mound was constructed, anchoring the northern part of the site and formalizing the coalescence of people at the site, while the knoll likely served as ritual space at the southern end of the site. As more members joined the community, some probably local, others coming from places as distant as Cahokia, the first wall may have been dismantled and a new wall erected to encompass the expansion of the town. The town likely evolved quickly, undergoing a major reorganization that included significant outward expansion, perhaps the establishment of a formal plaza, sculpting of the land and associated activities that created the sculptuary, construction of the northwest and southwest mounds, and the erection of a substantial new wall—all of which clearly express the emergent qualities of a coalescent com- munity. The northwest mound contains distinctive Mississippian ceramics and burials of individuals whose strontium isotopes indicate that some were local, including one female, but most came from two isotopically distinctive regions, one consistent with the bedrock geology of the Cahokia area, the other associ- ated with an as yet unknown region (Price et al. 2007). Pivotal moments in the dynamic history of the place—like the construction of the northeast mound, the erection of sequential walls around the community, the interment of indi- viduals from diverse places in the northwest mound, the creation of the sculp- tuary, and the modifications to the knoll—were consequences of an ongoing coalescence process that created a new community and reset the history of

166 Schroeder & Goldstein Aztalan. The end product of this coalescence was an archaeological record that appeared timeless to later antiquarians and scholars and became inscribed in the maps and artists’ renderings of the site. The proposition that Aztalan was a coalescent community aligns better with a dynamic view of Mississippian than it does with the conventional synchronous view.

Neot 1. Borrowed from David L. Clarke (1973).

References Alt, Susan M. 2002. “Identities, Traditions, and Diversity in Cahokia’s Uplands.” Midcontinental Journal of Archaeology 27:217–235. Alt, Susan M. 2006. “The Power of Diversity: Settlement in the Cahokian Uplands.” In Leadership and Polity in Mississippian Society, ed. Paul D. Welch and Brian M. Butler, 289–308. Occasional Paper No. 33. Carbondale: Center for Archaeological Investigations, Southern Illinois University. Alt, Susan M. 2008. “Unwilling Immigrants: Culture, Change, and the ‘Other.” In Invisible Citizens: Captives, Slaves, and Cultural Transmission in Middle Range Societies, ed. Catherine M. Cameron, 205–222. Salt Lake City: University of Utah Press. Anderson, David G. 1994a. “Factional Competition and Political Evolution of Mississippian Chiefdoms in the Southeastern United States.” In Factional Competition and Political Development in the New World, ed. Elizabeth M. Brumfiel and John W. Fox, 61–76. Cambridge: Cambridge University Press. http://dx.doi .org/10.1017/CBO9780511598401.007. Anderson, David G. 1994b. The Savannah River Chiefdoms: Political Change in the Late Prehistoric Southeast. Tuscaloosa: University of Alabama Press. Anderson, David G., and Joseph Schuldenrein. 1985. Prehistoric Human Ecology along the Upper Savannah River: Excavations at Rucker’s Bottom, Abbeville, and Bullard Site Groups. Atlanta: Archaeological Services Branch, Southeast Regional Office, National Park Service. Anderson, James P. 1969. “Cahokia Palisade Sequence.” In Explorations into Cahokia Archaeology, ed. Melvin L. Fowler, 89–99. Illinois Archaeological Survey, Bulletin No. 7. Urbana: University of Illinois. Anthony, David W. 1990. “Migration in Archeology: The Baby and the Bathwater.” American Anthropologist 92(4):895–914. http://dx.doi.org/10.1525/aa.1990.92 .4.02a00030.

Timelessness and the Legacy of Archaeological Cartography 167 Bailey, Geoff. 2007. “Time Perspectives, Palimpsests, and the Archaeology of Time.” Journal of Anthropological Archaeology 26(2):198–223. http://dx.doi.org/10.1016/j.jaa .2006.08.002. Barreis, David A., and Reid A. Bryson. 1965. “Climatic Episodes and the Dating of the Mississippian Cultures.” Wisconsin Archeologist 46:203–220. Barrett, Samuel A. 1933. Ancient Aztalan. Bulletin No. 13. Milwaukee: Museum of the City of Milwaukee. Beck, Robin A., Jr., Douglas J. Bolender, James A. Brown, and Timothy K. Earle. 2007. “Eventful Archaeology: The Place of Space in Structural Transformation.” Current Anthropology 48(6):833–860. http://dx.doi.org/10.1086/520974. Benden, Danielle. 2004. “The Fisher Mounds Site Complex: Early Middle Mississippian Exploration in the Upper Mississippi Valley.” Minnesota Archaeologist 63:7–24. Bhabha, Homi K. 1985. “Signs Taken for Wonders: Questions of Ambivalence and Authority under a Tree Outside Delhi, May 1817.” Critical Inquiry 12(1):144–165. http://dx.doi.org/10.1086/448325. Birmingham, Robert, and Lynne G. Goldstein. 2005. Aztalan: Mysteries of an Ancient Indian Town. Madison: Wisconsin Historical Society Press. Black, Glenn A. 1967. Angel Site: An Archaeological, Historical, and Ethnological Study. Indianapolis: Indiana Historical Society. Blitz, John H. 1993. Ancient Chiefdoms of the Tombigbee. Tuscaloosa: University of Alabama Press. Butler, Brian M., R. Berle Clay, Michael L. Hargrave, Steffan D. Peterson, John E. Schwegman, John A. Schwegman, and Paul D. Welch. 2011. “A New Look at Kincaid: Magnetic Survey of a Large Mississippian Town.” Southeastern Archaeology 30(1):20–37. http://dx.doi.org/10.1179/sea.2011.30.1.002. Clarke, David L. 1973. “Archaeology: The Loss of Innocence.” Antiquity 47(185):6–18. http://dx.doi.org/10.1017/S0003598X0003461X. Cobb, Charles R. 2005. “Archaeology and the ‘Savage Slot’: Displacement and Emplacement in the Premodern World.” American Anthropologist 107:563–574. Cobb, Charles R. 2008. “From Frontier to Border along the Iroquois Southern Door.” Archaeologies 4(1):110–128. http://dx.doi.org/10.1007/s11759-008-9063-y. Cobb, Charles R., and Brian M. Butler. 2006. “Mississippian Migration and Emplacement in the Lower Ohio Valley.” In Leadership and Polity in Mississippian Society, ed. Paul D. Welch and Brian M. Butler, 328–347. Occasional Paper No. 33. Carbondale: Center for Archaeological Investigations, Southern Illinois University. Cole, Fay-Cooper. 1951. Kincaid: A Prehistoric Metropolis. Chicago: University of Chicago Press.

168 Schroeder & Goldstein DeBoer, Warren R. 2005. “Colors for a North American Past.” World Archaeology 37(1):66–91. http://dx.doi.org/10.1080/00438240942000329577. Dirst, Victoria. 1988. Research in Pursuit of the Past at Governor Nelson State Park, Dane County, Wisconsin. Madison: Bureau of Parks and Recreation, Wisconsin Department of Natural Resources. Dirst, Victoria. 1995. Stockbridge Harbor: A Late Woodland Village on Lake Winnebago. Madison: Bureau of Property Management, Wisconsin Department of Natural Resources. Ethridge, Robbie Franklyn, and Charles M. Hudson. 2002. The Transformation of the Southeastern Indians, 1540–1760. Jackson: University Press of Mississippi. Finney, Fred A. 2013. “Intrasite and Regional Perspectives on the Fred Edwards Site and the Stirling Horizon in the Upper Mississippi Valley.” Wisconsin Archeologist 94:3–248. Finney, Fred A., and James B. Stoltman. 1991. “Fred Edwards Site: A Case of Stirling Phase Culture Contact in Southwestern Wisconsin.” In New Perspectives on Cahokia: Views from the Periphery, ed. James B. Stoltman, 229–252. Monographs in World Archaeology No. 2. Madison: Prehistory Press. Freeman, Joan. 1986. “Aztalan: A Middle Mississippian Village.” Wisconsin Archeologist 67:339–359. Gibbon, Guy E., and Clark A. Dobbs. 1991. “Mississippian Presence in the Red Wing Area, Minnesota.” In New Perspectives on Cahokia: Views from the Periphery, ed. James B. Stoltman, 281–306. Monographs in World Archaeology No. 2. Madison: Prehistory Press. Goldstein, Lynne. 1991. “Implications of Aztalan’s Location.” In New Perspectives on Cahokia: Views from the Periphery, ed. James B. Stoltman, 209–227. Monographs in World Archaeology No. 2. Madison: Prehistory Press. Goldstein, Lynne. 2010. “Aztalan Mortuary Practices Revisited.” In Mississippian Mortuary Practices: Beyond Hierarchy and the Representationist Perspective, ed. Lynne P. Sullivan and Robert Mainfort, 90–112. Gainesville: University Press of Florida. http://dx.doi.org/10.5744/florida/9780813034263.003.0006. Goldstein, Lynne, and Joan Freeman. 1997. “Aztalan: A Middle Mississippian Village.” Wisconsin Archeologist 78:223–248. Goldstein, Lynne, and Donald Gaff. 2002. “Recasting the Past: Examining Assumptions about Aztalan.” Wisconsin Archeologist 83:98–110. Green, William, and Roland L. Rodell. 1994. “Mississippian Presence and Cahokia Interaction at Trempealeau, Wisconsin.” American Antiquity 59(2):334–359. http://dx.doi.org/10.2307/281936.

Timelessness and the Legacy of Archaeological Cartography 169 Hall, Robert L. 1962. The Archaeology of Carcajou Point. 2 vols. Madison: University of Wisconsin Press. Hall, Robert L. 2004. “The Cahokia Site and Its People.” In Hero, Hawk, and Open Hand: American Indian Art of the Ancient Midwest and South, ed. Richard F. Townsend and Robert V. Sharp, 93–103. New Haven, CT: Art Institute of Chicago in association with Yale University Press. Hammerstedt, Scott W. 2005 “Mississippian Construction, Labor, and Social Organization in Western Kentucky.” PhD dissertation, Department of Anthropology, Pennsylvania State University, University Park, PA. Holmes, William H. 1903. Aboriginal Pottery of the Eastern United States. Bureau of American Ethnology Twentieth Annual Report. Washington, DC: Smithsonian Institution. Hurley, William M. 1977. “Aztalan Revisited.” Wisconsin Archeologist 58:256–294. Hyer, Nathaniel F. 1837. Letter to Editor, with illustration. Milwaukee Advertiser, 25 February 1837:2. http://www.wisconsinhistory.org/turningpoints/. Hyer, Nathaniel F. 1838. Letter Describing Aztalan, April 28, 1838. Wisconsin Historical Society Archives collection SC 2971. http://content.wisconsinhistory .org/cdm/ref/collection/tp/id/44360. Iseminger, William. 2010. Cahokia Mounds: America’s First City. Charleston, SC: History Press. Kelly, John E., and James A. Brown. 2014. “Cahokia: The Processes and Principles of the Creation of an Early Mississippian City.” In Making Ancient Cities: Space and Place in Early Urban Societies, ed. Andrew T. Creekmore, III, and Kevin Fisher, 292–336. Cambridge: Cambridge University Press. http://dx.doi.org/10.1017/CBO 9781107110274.010. Kidder, Tristram R. 1998. “Mississippi Period Mound Groups and Communities in the Lower Mississippi Valley.” In Mississippian Towns and Sacred Spaces: Searching for an Architectural Grammar, ed. R. Barry Lewis and Charles Stout, 123–150. Tuscaloosa: University of Alabama Press. King, Adam. 2001. “Long-Term Histories of Mississippian Centers: The Developmental Sequence of Etowah and Its Comparison to Moundville and Cahokia.” Southeastern Archaeology 20:1–17. King, Adam. 2003. Etowah: The Political History of a Chiefdom Capital. Tuscaloosa: University of Alabama Press. Knight, Vernon James, Jr. 1985. “Theme and Variation in Mississippian Ritual Expression.” In Indians, Colonists, and Slaves: Essays in Memory of Charles H. Fairbanks, ed. Kenneth W. Johnson, Jonathan M. Leader, and Robert C. Wilson, 105–116. Special Publication 4. Gainesville: Florida Journal of Anthropology.

170 Schroeder & Goldstein Knight, Vernon James, Jr. 1989. “Symbolism of Mississippian Mounds.” In Powhatan’s Mantle, ed. Gregory Waselkov, Peter Wood, and Thomas Hatley, 279–291. Lincoln: University of Nebraska Press. Knight, Vernon James, Jr. 2010. Mound Excavations at Moundville: Architecture, Elites, and Social Order. Tuscaloosa: University of Alabama Press. Knight, Vernon James, Jr., and Vincas Steponaitis, eds. 1998. Archaeology of the Moundville Chiefdom. Washington, DC: Smithsonian Institution Press. Kowalewski, Stephen A. 2001. “Coalescent Societies.” Southeastern Archaeological Conference Bulletin 44:27. Krus, Anthony. 2011. “Refortifying Cahokia: More Efficient Palisade Construction through Redesigned Bastions.” Midcontinental Journal of Archaeology 36(2):227–244. http://dx.doi.org/10.1179/mca.2011.011. Krus, Anthony, Timothy Schilling, and G. William Monaghan. 2013. “The Timing of Angel Mounds Palisade Construction: A Search for the Best Chronological Model.” Midcontinental Journal of Archaeology 38(2):171–182. http://dx.doi.org/10 .1179/mca.2013.006. Lapham, Increase A. 1855. The Antiquities of Wisconsin, as Surveyed and Described. Smithsonian Contributions to Knowledge. vol. 7. Washington, DC: Smithsonian Institution. http://dx.doi.org/10.5962/bhl.title.40358. Lewis, R. Barry, and Charles Stout, eds. 1998. Mississippian Towns and Sacred Spaces: Searching for an Architectural Grammar. Tuscaloosa: University of Alabama Press. Lewis, R. Barry, Charles Stout, and Cameron Wesson. 1998. “The Design of Mississippian Towns.” In Mississippian Towns and Sacred Spaces: Searching for an Architectural Grammar, ed. R. Barry Lewis and Charles Stout, 1–21. Tuscaloosa: University of Alabama Press. Lewis, Theodore H. 1894. “The ‘Aztalan’ Enclosure Newly Described.” American Antiquarian and Oriental Journal 16:205–208. Liebmann, Matthew. 2013. “Parsing Hybridity: Archaeologies of Amalgamation in Seventeenth-Century New Mexico.” In The Archaeology of Hybrid Material Culture, ed. Jeb J. Card, 25–49. Occasional Paper No. 39. Carbondale: Center for Archaeological Investigations, Southern Illinois University. Maher, Robert F. 1958. “Excavation and Reconstruction of the Southwest Pyramidal Mound at Aztalan.” Wisconsin Archeologist 39:77–101. Maxwell, Moreau S. 1952. “Clay Earspools from the Aztalan Site, Wisconsin.” American Antiquity 18(1):61–63. http://dx.doi.org/10.2307/276244. McKern, William C. 1946. “Aztalan.” Wisconsin Archeologist 27:41–52. Milner, George R. 1998. The Cahokia Chiefdom: The Archaeology of a Mississippian Society. Washington, DC: Smithsonian Institution Press.

Timelessness and the Legacy of Archaeological Cartography 171 Milner, George R. 1999. “Warfare in Prehistoric and Early Historic Eastern North America.” Journal of Archaeological Research 7(2):105–151. http://dx.doi.org/10.1007 /s10814-005-0001-x. Nassaney, Michael S. 1992. “Communal Societies and the Emergence of Elites in the Prehistoric American Southeast.” In Lords of the Southeast: Social Inequality and the Native Elites of Southeastern North America, ed. Alex W. Barker and Timothy R. Pauketat, 111–143. Archeological Papers No. 3. Washington, DC: American Anthropological Association. http://dx.doi.org/10.1525/ap3a.1992.3.1.111. Overstreet, David F. 1995. “The Eastern Wisconsin Oneota Regional Continuity.” In Oneota Archaeology – Past, Present and Future, ed. W. Green, 33–64. Report No. 20, Office of the State Archaeologist. Iowa City: University of Iowa. Pauketat, Timothy R. 1994. The Ascent of Chiefs: Cahokia and Mississippian Politics in Native North America. Tuscaloosa: University of Alabama Press. Pauketat, Timothy R. 2004. Ancient Cahokia and the Mississippians. Cambridge: Cambridge University Press. Pauketat, Timothy R., and Susan M. Alt. 2005. “Agency in a Postmold? Physicality and the Archaeology of Culture-Making.” Journal of Archaeological Method and Theory 12(3):213–236. http://dx.doi.org/10.1007/s10816-005-6929-9. Pauketat, Timothy R., Robert F. Boszhardt, and Danielle Benden. 2015. “Trempealeau Entanglements: An Ancient Colony’s Causes and Effects.” American Antiquity 80(2):260–289. http://dx.doi.org/10.7183/0002-7316.80.2.260. Payne, Claudine, and John E. Scarry. 1998. “Town Structure at the Edge of the Mississippian World.” In Mississippian Towns and Sacred Spaces: Searching for an Architectural Grammar, ed. R. Barry Lewis and Charles Stout, 22–48. Tuscaloosa: University of Alabama Press. Phillips, Philip, James A. Ford, and James B. Griffin. 1951. Archaeological Survey in the Lower Mississippi Alluvial Valley, 1940–1947. Papers of the Peabody Museum of American Archaeology and Ethnology. vol. 25. Cambridge, MA: Harvard University. Polhemus, Richard, ed. 1987. The Toqua Site: A Late Mississippian Dallas Phase Town. 2 vols. Report of Investigations, No. 41, Department of Anthropology, University of Tennessee, and Publications in Anthropology, No. 44. Knoxville: Tennessee Valley Authority. Price, T. Douglas, James H. Burton, and James B. Stoltman. 2007. “Place of Origin of Prehistoric Inhabitants of Aztalan, Jefferson County, Wisconsin.” American Antiquity 72(3):524–538. http://dx.doi.org/10.2307/40035859. Pursell, Corin C. O. 2013. “Colored Monuments and Sensory Theater among the Mississippians.” In Making Senses of the Past: Toward a Sensory Archaeology, ed. Jo

172 Schroeder & Goldstein Day, 69–89. Center for Archaeological Investigations Occasional Paper No. 40. Carbondale: Southern Illinois University. Richards, John D. 2003. “Collars, Castellations, and Cahokia: A Regional Perspective on the Aztalan Ceramic Assemblage.” Wisconsin Archeologist 84:139–153. Richards, John D. 2007. “Viewing the Ruins: The Early Documentary History of Aztalan.” Wisconsin Magazine of History 91(2):28–39. Richards, John D., and Robert J. Jeske. 2002. “Location, Location, Location: The Temporal and Cultural Context of Late Prehistoric Settlement in Southeast Wisconsin.” Wisconsin Archeologist 83:32–54. Ritzenthaler, Robert E. 1961. “Radiocarbon Dates for Aztalan.” Wisconsin Archeologist 42:139. Ritzenthaler, Robert E. 1963. “Radiocarbon Dates for Aztalan.” Wisconsin Archeologist 44:18. Rodell, Roland L. 1991. “Diamond Bluff Site Complex and Cahokia Influence in the Red Wing Locality.” In New Perspectives on Cahokia: Views from the Periphery, ed. James B. Stoltman, 253–280. Monographs in World Archaeology No. 2. Madison: Prehistory Press. Rodning, Christopher B. 2010. “Architectural Symbolism and Cherokee Townhouses.” Southeastern Archaeology 29(1):59–79. http://dx.doi.org/10.1179/sea.2010.29.1.005. Rowe, Chandler W. 1958. “A Crematorium at Aztalan.” Wisconsin Archeologist 39:101–110. Salkin, Philip H. 2000. “The Horicon and Kekoskee Phases: Cultural Complexity in the Late Woodland Stage in Southeastern Wisconsin.” In Late Woodland Societies: Tradition and Transformation across the Midcontinent, ed. Thomas E. Emerson, Dale L. McElrath, and Andrew C. Fortier, 525–542. Lincoln: University of Nebraska Press. Schilling, Timothy M. 2010. “An Archaeological Model of the Construction of Monks Mound and Implications for the Development of Cahokian Society (ad 800–1400).” PhD dissertation, Department of Anthropology, Washington University, St. Louis, MO. Schilling, Timothy M. 2012. “Building Monks Mound, Cahokia, Illinois, ad 800– 1400.” Journal of Field Archaeology 37(4):302–313. http://dx.doi.org/10.1179/0093469 012Z.00000000027. Schroeder, Sissel. 2004. “Current Research on Late Pre-Contact Societies of the Midcontinental United States.” Journal of Archaeological Research 12(4): 311–372. http://dx.doi.org/10.1007/s10814-004-0001-2. Schroeder, Sissel. 2006. “Walls as Symbols of Political, Economic, and Military Might.” In Leadership and Polity in Mississippian Society, ed. Brian M. Butler

Timelessness and the Legacy of Archaeological Cartography 173 and Paul D. Welch, 115–141. Occasional Paper No. 33, Center for Archaeological Investigations. Carbondale: Southern Illinois University. Schroeder, Sissel. 2011. “An Investigation of the Origins of Variation in Perishable Architecture at Jonathan Creek.” Southeastern Archaeology 30(2):311–326. Sherrod, P. Clay, and Martha A. Rolingson. 1987. Surveyors of the Ancient Mississippi Valley: Modules and Alignments in Prehistoric Mound Sites. Research Series No. 28. Fayetteville: Arkansas Archeological Survey. Smith, Bruce D. 1986. “Archaeology of the Southeastern United States: From Dalton to de Soto, 10,500–500 B.P.” Advances in World Archaeology 5:1–92. Squier, Ephraim G., and Edwin H. Davis. 1848. Ancient Monuments of the Mississippi Valley: Comprising the Results of Extensive Original Surveys and Explorations. Smithsonian Contributions to Knowledge. vol. 1. Washington, DC: Smithsonian Institution. Stout, Charles B. 1984. “Mississippian Sites in Western Kentucky: Variations on a General Mississippian Theme?” In Late Prehistoric Research in Kentucky, ed. David Pollack, Charles D. Hockensmith, and Thomas Sanders, 167–179. Frankfort: Kentucky Heritage Council. Sullivan, Alan P., III. 2008. “Time Perspectivism and the Interpretive Potential of Palimpsests: Theoretical and Methodological Considerations of Assemblage Formation History and Contemporaneity.” In Time in Archaeology: Time Perspectivism Revisited, ed. Simon Holdaway and LuAnn Wandsnider, 31–45. Salt Lake City: University of Utah Press. Thomas, Cyrus. 1894. Report on the Mound Explorations of the Bureau of Ethnology. 12th Annual Report of the Bureau of American Ethnology, 1890–1891. Washington, DC: United States Government Printing Office. von Humboldt, Alexander. 1814. Researches, Concerning the Institutions and Monuments of the Ancient Inhabitants of America with Descriptions and Views of Some of the Most Striking Scenes in the Cordilleras! Vols. I and II. Originally published in French in 1810, translated into English by Helen Maria Williams. Longman, Hurst, Rees, Orme and Brown. London: J. Murray and H. Colburn. http://dx.doi.org/10.5962 /bhl.title.61789. Wittry, Warren L., and David A. Barreis. 1958. “Domestic Houses at Aztalan.” Wisconsin Archeologist 39:62–77. Zych, Thomas. 2013. “The Construction of a Mound and a New Community: An Analysis of the Ceramic and Feature Assemblages from the Northeast Mound at the Aztalan Site.” Unpublished Master’s thesis, Department of Anthropology, University of Wisconsin, Milwaukee, WI.

174 Schroeder & Goldstein 8

One of the continuing challenges of Bronze Age Sherd Cross-Joins, archaeology is determining the time relations among Ceramic Use-Wear, and the massive architectural features that characterize Depositional History major Minoan settlements on Crete, such as Myrtos- Pyrgos. In view of the duration of occupation of com- Rethinking the Sociopolitical munities like Myrtos-Pyrgos, archaeologists frequently Aftermath of a Collapsed encounter complex sequences of monumental feature Bronze Age Cistern at construction, use, destruction, and ultimately abandon- Myrtos-Pyrgos, Crete ment. To enable precise understandings of the nature and scale of activities at these political centers, it is vital to establish the histories of use, disuse, and reuse of key Emilia Oddo and public architectural features (Letesson 2015). Gerald Cadogan In this regard, this chapter presents the results of the analysis of a large ceramic assemblage recovered from the massive mid-second millennium bc cistern (Cistern 2) at Myrtos-Pyrgos in the early 1970s. Our study involves the stratigraphic analysis of patterns of ceramic cross-joins, that is, sherds from the same once-intact vessel that can be reassembled into larger fragments, and use-wear to refine accounts of the for- mation history of the deposits within and above the cistern. By considering the cistern’s ceramic assemblage in this way, we anticipate that a fuller understanding will emerge of how a monumental feature intended for water storage became a receptacle for discarded pot- tery and other goods, and what these developments imply for the place of Myrtos-Pyrgos in the political economy of ancient Crete. DOI: 10.5876/9781607324942.c008

175 Table 8.1. Cultural chronology of Myrtos-Pyrgos, Crete.

Approximate dates bc Minoan system Myrtos–Pyrgos phases Neopalatial 1750/1700–1460 LM IB Destruction by fire (MM III?)–LM I (IA and IB) Pyrgos IV

Protopalatial 1850–1750/1700 MM IIB Destruction by fire MM IIB Pyrgos III

Prepalatial 2250–1850 EM III/MM IA–MM IB (and IIA) Pyrgos II 2650–2250 EM IIB Destruction by fire EM II Pyrgos I 3100/3000–2650 EM I Pyrgos 0.b 3300–3100/3000 FN IV Pyrgos 0.a

Terminology and Chronology Since Sir Arthur Evans’s excavations at Knossos, starting in 1900, the Cretan Bronze Age (table 8.1) has been generally known as Minoan after the mythical king Minos of Knossos. Evans’s scheme (Evans 1906) for the island’s Bronze Age occupation fol- lowing the Final Neolithic (FN) has three main periods: Early Minoan (EM), Middle Minoan (MM), and Late Minoan (LM), which he divided into three phases each (I, II, III) and, in most cases, into “A” and “B”—thus producing, for instance, MM IIIA. Although there have subsequently been further subdivisions (e.g., LM IIIA1, LM IIIC), and an alternative sys- tem has been proposed based on the construction history of the so-called Minoan palaces, resulting in Prepalatial, Protopalatial, Neopalatial, Final Palatial, and Postpalatial periods, Evans’s Minoan system is still the core framework for arranging and understanding the Bronze Age archaeology of Crete. Table 8.1 also shows the correspondences between the periods at Myrtos-Pyrgos (which we use in this chapter) and the Minoan and palatial chronologies.

176 Oddo & Cadogan Figure 8.1. The eastern Mediterranean, showing the location of Myrtos-Pyrgos (modifi- cation of an original by John Walrodt, Department of Classics, University of Cincinnati).

Moyrt s-Pyrgos The site of Myrtos-Pyrgos is situated on the south coast of Crete, on the east side of the mouth of the Myrtos River, opposite the modern village of Myrtos (on the river’s west side; figure 8.1). First visited by archaeologists in 1962 (Hood et al. 1964), the site was excavated by Cadogan for the British School at Athens between 1970 and 1973, with supplementary campaigns con- tinuing until 1994 (Cadogan 1978, 1992). The center of the site is a small plateau on top of the conical hill, 77.63 m high, called Pyrgos (which means “tower” in Greek and refers to a Venetian and Ottoman beacon-tower on the summit; figure 8.2). Important architec- tural features of Myrtos-Pyrgos include a grand multistoried building (known as the “Country House”) on the flat top of the hill (Cadogan 1978) that is surrounded by a number of smaller buildings, probably houses, on the west, north, and east slopes. In addition, these structures are accompanied by a tomb with two ossuaries, containing many burials and offerings (Cadogan 2011), a defensive system of walls and a bastion-tower, and two water cisterns: a small one in the courtyard in front of the Country House (Cistern 1) and a large one on the north slope (Cistern 2), which is the focus of this study.

Rehint king the Sociopolitical Aftermath of a Collapsed Bronze Age Cistern 177 Figure 8.2. Major architectural features, Myrtos-Pyrgos.

Myrtos-Pyrgos was inhabited for a long time during the Bronze Age, from at least EM II (with a little evidence also of use in FN IV–EM I) to LM I, with a few intermittent breaks of occupation. Much later, we find a Hellenistic shrine (late second–early first centuries bc) and the Venetian/Ottoman tower. The site’s heyday, however, was the Bronze Age, when it witnessed a remark- able amount of monumental building in a small settlement and an intense social and cultural life, as well as three destructions by fire. The latest of these conflagrations in LM IB ended the Minoan life of the settlement.

Cistern 2 Cistern 2 is situated at the north edge of the settlement, about 2.5 m downslope from a defensive terrace wall, 10 m east of the Tower, and 12 m below the hilltop (Cadogan 2007:105). Excavation1 revealed that its northern half, on the edge of the hill, was built up, but the southern part was dug into the hillside (figure 8.3). The cistern itself is a massive, more or less circular struc- ture roughly 5.3 m in diameter and more than 3 m deep that would have held at least 66 tonnes of water when full (Cadogan 1978:74). A layer of white lime

178 Oddo & Cadogan Figure 8.3. Cistern 2 after excavation. North (downslope) is to the left.

plaster coats both the walls, built with large blocks of stone, and its pebble floor. The feature’s position, shape, size, and internal plastering imply that it was designed to hold water for the entire community (Cadogan 2007, 2014).

Cistern 2’s Ceramic Assemblage: Methods and Results Pottery was the predominant type of artifact recovered in the fill of Cistern 2. Approximately 2,000 coarse-ware and fine-ware sherds from the upper lay- ers of the cistern form the analytical assemblage, 99 percent of which can be dated to the LM IA phase, plus a handful of later LM IB sherds.2 The major- ity of these Neopalatial sherds come from the northernmost trenches, F02 and G02 (figure 8.4). During post-excavation study, the pottery from these trenches was heavily selected, a practice that was typical of Cretan archaeology for most of the twentieth century (Momigliano 2007). The selection saved diagnostic sherds and was intended to preserve the proportions of wares, thereby maintaining a representative sample of ware variability. This practice worked to our advan- tage because our focus is the study of the cistern’s depositional history, particu- larly whether its deposits were the product of one or several dumping episodes.

Rehint king the Sociopolitical Aftermath of a Collapsed Bronze Age Cistern 179 Figure 8.4. Frequencies of Neopalatial sherds within Cistern 2. The northern trenches (F02, G02, plus baulk G01/G02) contained the highest concentration of pottery.

The assemblage was examined with specific attention to patterns of cross-joins, degree of completeness, and use-wear traces among the sherds from different levels (cf. Chapman and Gaydarska 2007). As it is easiest to find cross-joins by looking at the fine decorated pottery, the wealth of retained fine ware in Cistern 2 was ideal for our purposes. Analysis showed that about 70 percent of the selected sherds could be mended and, in a few cases, refitted into almost complete vessels. Very few traces of use-wear are present: breaks have sharp edges, and surfaces preserve their glossy, almost pristine finish. These characteristics can be observed on sherds from all levels in the cistern, except for earlier (MM IIB) pottery lying at the bottom of the cistern (Oddo 2015). Figure 8.5 shows the pattern of cross-joins against the cistern’s sections. The upper levels had the most pottery, about 80 percent of what was examined. Lower down in the cistern, the number of sherds decreased sharply; the floor of the cistern yielded Protopalatial oval-mouthed amphorae and other vessels. A few Neopalatial sherds recovered from the bottom of the cistern in all likeli- hood originated from the overlying deposits.

180 Oddo & Cadogan Cistern 2’s Depositional History Clarifying the depositional history of Cistern 2 helps to refine the sequence of its use, disuse, and reuse and, in turn, integrate it into Myrtos-Pyrgos’s overall occupation (cf. Goldberg and Macphail 2006:219–224). From studying the stratigraphy and the cross-joins in the pottery, the history of Cistern 2 can be arranged in a sequence of five phases. Phase 1 covers the initial construction and use of the cistern for storing water during the Pyrgos III period (MM IIB, late Protopalatial), or possibly earlier. Phase 2 witnessed the collapse of the cistern’s outer (downslope) wall. As a result, its capacity for holding water was dramatically reduced, if not elimi- nated altogether. Thereafter, during phase 3, a fairly thick layer of earth and stone accumu- lated, which may be related to the likely abandonment of the site in the last phase of the Middle Bronze Age (MM III; Cadogan 2013). Then, in phase 4, broken pottery was dumped over the earth and stone layer. These ceramics date to the Pyrgos IV period (Neopalatial period, or LM IA and LM IB) and may have originated as a consequence of the removal of material from buildings located somewhere else on the hill, most probably on the hilltop. At about the same time a rough wall (Wall FB in figures 8.6 and 8.7) was erected in the northeast sector of the cistern, where the original cistern wall had fallen away. Like the pottery dump, it rested on the earth and stone layer below. Its pur- pose seems to have been to keep the dump from spreading downhill, where the ground flattens out and fields may have been located. Wall FB could have been put up before dumping began, or while it was going on, or after it was completed: the first two possibilities belong to phase 4, the last one to a probable phase 5. Two other rough walls may be assigned to phase 5. One of them, Wall GT, runs across the line of the collapsed north part of the cistern wall at a much higher level (and found just below the modern surface), suggesting that it was designed to retain a large dump. It is also well above wall FB, and so was emplaced later. The other wall, Wall ET, was also probably intended to stop rubbish from moving downslope. The likely date for these phase 5 walls is that they, too, belong to Pyrgos IV (LM I), because there is no evidence for later occupation of this part of the hill.3

T he Slow-Filling and Fast-Filling Models The circumstances that led to the formation of this dense deposit of pottery during Pyrgos IV require further inquiry to account for the cross-joins and wear traces among the sherds. Here, we explore two possibilities.

Rehint king the Sociopolitical Aftermath of a Collapsed Bronze Age Cistern 181 Figure 8.5. Diagram of ceramic cross-joins among the excavation sections in Cistern 2.

Figure 8.6. Plan of Cistern 2, showing post-collapse wall remains. Figure 8.5. Diagram of ceramic cross-joins among the excavation sections in Cistern 2. The first possibility—the slow-filling model—hypothesizes that material was dumped several times into the disused cistern. If this were the case, it is reasonable to assume that, in the intervals between dumpings, the broken pot- tery would have been exposed to the elements, which would have left traces, such as discolored or dull patches on painted surfaces or worn, rounded edges where the pottery was broken (e.g., Allen 1989). Yet, few sherds from the assemblage have such traces (< 30%). Moreover, if the filling took place over a long period of time, one might expect that some surface-alteration traces (e.g., hardening of unprotected soil, establishment of vegetation) would have been produced (e.g., Leigh 2001). However, none were observed in excavation—but, in the era of the 1970s, neither were they looked for. The fast-filling model, in contrast, posits that material was dumped rap- idly in the disused cistern, perhaps as one episode, and is supported by three observations (cf. Shillito and Matthews 2013). First, ceramic cross-joins are concentrated in the cistern’s upper levels (figure 8.8). Second, most (> 70%) of the sherds in the assemblage could be mended. And third, the sherds’ wear patterns indicate little exposure to the elements, which we interpret to mean that the pottery was buried quickly and so protected from weathering agents after it had been discarded. Additional support for the fast-filling model is the fact that the few LM IB sherds, which were concentrated in the very top levels of the dump that had spread beyond the confines of the cistern, had few joins between them, and consistently showed a high degree of surface wear (figure 8.9).

Rehint king the Sociopolitical Aftermath of a Collapsed Bronze Age Cistern 183 Figure 8.7. (Left) Wall GT (its distal portion is at the end of the horizontal meter stick) and (upper center) Wall FB (against which the vertical meter stick rests), after excavation of Cistern 2.

Cuoncl sions: Beyond Pottery and Stratigraphy Although the Neopalatial artifacts recovered from Cistern 2 represent mate- rial collected from elsewhere on the site and deposited there, they originated as a consequence of a variety of activities that are worth considering. First, the sherds may represent routine garbage disposal, a hypothesis that would favor the slow-filling model. The strength of this argument is the presence in the cistern’s fill of a variety of material besides pottery, including plaster, broken stone vases, and building stones. Second, it is striking to note that about 95 percent of the Neopalatial shapes represented in Cistern 2 correspond to finely decorated vessels that are usually interpreted as shapes intended for purposes such as pouring, drinking, serving, and eating. Although it is tempting to use the selected nature of the assemblage as a counterargument against this hypothesis, it is important to remember that the post-excavation processing procedure was designed to maintain ware/function (shape) proportions between the

184 Oddo & Cadogan Figure 8.8. Section of Cistern 2’s southeastern quarter showing the post-collapse fill sequence. The levels that yielded ceramic cross-joins are highlighted. pre- and postselection assemblages. In other words, the pottery found in the excavation contained a high proportion of decorated (fine-ware) vessels presumably used (functioning) for drinking and eating. Hence, it could be argued that this assemblage represents the remains of banqueting that prob- ably took place higher up the hill, the aftermath of the feasts being depos- ited in the disused cistern. The social and political importance of feasts in ancient societies has been much discussed, particularly in regards to political factionalism (e.g., Brumfiel and Fox 1994). Social gatherings allowed elites to legitimize their power and to engage in conspicuous consumption, with display of prestige objects and, in some cases, large quantities of food (e.g., Dietler and Hayden 2001). As such, ceramic shapes and styles become indicators of social and perhaps political allegiance, where decorative themes played a role in the identification and negotiation of social groups (Hamilakis 1996). To judge from ceramic style and function alone, the LM I pottery in Cistern 2 may well represent the remains of such social gatherings. This hypothesis has the potential to account for the types of pottery and their cross-join patterns found in the cistern, which combined support the fast-fill model. We must keep in mind, however, the difficulty in establishing archaeologically a clear “feasting kit,” and how

Rehint king the Sociopolitical Aftermath of a Collapsed Bronze Age Cistern 185 Figure 8.9. Effects of weathering on decoration and conjoinability expressed on fine- ware sherds recovered from the middle levels of the Neopalatial-period dump (left column) versus those recovered from its surface levels (right column).

the evidence from bones or from pottery can often be misinterpreted (e.g., Dabney et al. 2004; Halstead and Isaakidou 2004, 2011; Isaakidou 2007, 2008, 2011; Isaakidou and Halstead 2013). From a social and political point of view, both the clearance and the feast- ing scenarios may be connected with the remodeling of the Country House

186 Oddo & Cadogan during the LM I period (a phenomenon still being studied), which in turn is attributable to the consolidation of power by the occupants of the building. This scenario parallels that of Cistern 1, a smaller cistern on the southwestern side of the Country House courtyard (figure 8.2). Although built during the Protopalatial period, by the end of the Neopalatial period Cistern 1 ceased to be used for water storage and was filled with pebbles brought up from the river, which may have been a ritual process (Cadogan 2007). It is likely, too, that the filling of Cistern 2 was connected to a ritual through the deposition of domestic or feasting debris. The rebuilding of the Country House on the hilltop might have been not just an architectural enterprise, but equally some- thing triggered by a sociopolitical reorganization, which might well have had implications for Myrtos-Pyrgos’s final destruction. The evidence and the arguments presented above represent an attempt to reconstruct Myrtos-Pyrgos’s history at the end of the Neopalatial period. The completion of the study of Neopalatial Myrtos-Pyrgos will certainly help produce a more fine-grained assessment of the social activities involving the assemblages, features, and buildings discussed above and, more generally, the entire community. Ultimately, this study aims to understand the nature of changing sociopolitical dynamics in southeastern Crete, and the rest of the island, during the Neopalatial period.

Acknowledgments We thank the British School at Athens for permission to present our research in this volume. We should also like to thank Dr. Eleni Hatzaki for stimulating discussion of earlier drafts of this chapter.

N otes 1. Cistern 2 was excavated during 1971 and 1973 as three 4m × 4m trenches (F02, G01, G02), two smaller trenches (F01 and F01X, where X stands for ‘extension’), with four 1 m-wide baulks between the trenches (figure 8.6). 2. Earlier pottery, mainly of Pyrgos III date, is currently being studied by Cadogan and Carl Knappett. 3. The LM IB destruction by fire brought an end to the Bronze Age occupation of the settlement (table 8.1). Myrtos-Pyrgos was then abandoned until Hellenistic times.

Rehint king the Sociopolitical Aftermath of a Collapsed Bronze Age Cistern 187 References Allen, J.R.L. 1989. “A Quantitative Technique for Assessing the Roundness of Pottery Sherds in Water Currents.” Geoarchaeology: An International Journal 4(2):143–155. http://dx.doi.org/10.1002/gea.3340040204. Brumfiel,E lizabeth M., and John W. Fox, eds. 1994. Factional Competition and Political Development in the New World. Cambridge: Cambridge University Press. http://dx.doi.org/10.1017/CBO9780511598401. Cadogan, Gerald. 1978. “Pyrgos, Crete, 1970–77.” Archaeological Reports 24:70–84. Cadogan, Gerald. 1992. “Myrtos-Pyrgos.” In The Aerial Atlas of Ancient Crete, ed. J. Wilson Myers, Eleanor Ellen Myers, and Gerald Cadogan, 202–209. Berkeley: University of California Press. Cadogan, Gerald. 2007. “Water Management in Minoan Crete, : The Two Cisterns of One Middle Bronze Age Settlement.” Water Science and Technology: Water Supply 7:103–111. Cadogan, Gerald. 2011. “Myrtos: From Phournou Koryphi to Pyrgos.” In Stega: The Archaeology of Houses and Households in Ancient Crete, ed. Kevin T. Glowacki and Natalia Vogeikoff-Brogan, 39–49. Hesperia Supplement 44. Princeton: American School of Classical Studies at Athens. Cadogan, Gerald. 2013. “Where Has Middle Minoan III Gone? A Lack at Myrtos- Pyrgos—and Elsewhere? What Does It Mean?” In Intermezzo: Intermediacy and Regeneration in Middle Minoan III Crete, ed. Colin F. Macdonald and Carl Knappett, 179–181. British School at Athens Studies 21. London: British School at Athens. Cadogan, Gerald. 2014. “Water Worries and Water Works in Bronze Age Southern Crete.” In Physis: l’environnement naturel et la relation homme-milieu dans le monde égéen protohistorique, ed. Gilles Touchais, Robert Laffineur, and Françoise Rougemont, 73–78. Aegaeum 37. Liège: Université de Liège, Histoire de l’art et archéologie de la Grèce antique. Chapman, John, and Bisserka Gaydarska. 2007. Parts and Wholes: Fragmentation in Prehistoric Context. Oxford: Oxbow Books. Dabney, Mary K., Paul Halstead, and Patrick Thomas. 2004. “Mycenaean Feasting on Tsoungiza at Ancient Nemea.” Hesperia 73(2):197–215. http://dx.doi.org/10.2972 /hesp.2004.73.2.197. Dietler, Michael, and Brian Hayden. 2001. Feasts: Archaeological and Ethnographic Perspectives on Food, Politics, and Power. Washington, DC: Smithsonian Institution Press. Evans, Arthur J. 1906. Essai de classification des époques de la civilisation minoenne. London: B. Quaritch.

188 Oddo & Cadogan Goldberg, Paul, and Richard I. Macphail. 2006. Practical and Theoretical Goearchaeology. Malden, MA: Blackwell Publishing. Halstead, Paul, and Valasia Isaakidou. 2004. “Faunal Evidence for Feasting: Burnt Offerings from the Palace of Nestor at Pylos.” In Food, Cuisine and Society in Prehistoric Greece, ed. Paul Halstead and John C. Barrett, 136–154. Sheffield Studies in Aegean Archaeology 5. Oxford: Oxbow. Halstead, Paul, and Valasia Isaakidou. 2011. “Political Cuisine: Rituals of Commensality in the Neolithic and Bronze Age Aegean.” In Guess Who’s Coming to Dinner: Feasting Rituals in the Prehistoric Societies of Europe and the Near East, ed. Gonzalo Aranda Jiménez, Sandra Monton-Subias, and Margarita Sánchez Romero, 91–108. Oxford: Oxbow. Hamilakis, Yannis. 1996. “Wine, Oil and the Dialectics of Power in Bronze Age Crete: A Review of the Evidence.” Oxford Journal of Archaeology 15(1):1–32. http://dx.doi.org/10.1111/j.1468–0092.1996.tb00071.x. Hood, Sinclair, Peter Warren, and Gerald Cadogan. 1964. “Travels in Crete, 1962.” Annual of the British School at Athens 59:50–99. http://dx.doi.org/10.1017/S00682 45400006080. Isaakidou, Valasia. 2007. “Cooking in the Labyrinth: Exploring “Cuisine” at Bronze Age Knossos.” In Cooking Up the Past: Food and Culinary Practices in the Neolithic and Bronze Age Aegean, ed. Christopher Mee and Josette Renard, 5–24. Oxford: Oxbow. Isaakidou, Valasia. 2008. “The Fauna and Economy of Neolithic Knossos Revisited.” In Escaping the Labyrinth: The Cretan Neolithic in Context, ed. Valasia Isaakidou and Peter Tomkins, 90–114. Sheffield Studies in Aegean Archaeology 8. Oxford: Oxbow. Isaakidou, Valasia. 2011. “Early Minoan I, the Palace Well: Faunal Remains and Taphonomy; Early Minoan II–III, Area A. Royal Road North: Faunal Remains; Early Minoan II–III, Area B. The Early Houses: Faunal Remains.” In Knossos Excavations, 1957–1961: Early Minoan, by Sinclair Hood and Gerald Cadogan, 63–67, 229–233, 237. British School at Athens Supplementary Volume 46. London: British School at Athens. Isaakidou, Valasia, and Paul Halstead. 2013. “Bones and the Body Politic? A Diachronic Analysis of Structured Deposition in the Neolithic-Early Iron Age Aegean.” In Bones, Behavior and Belief: The Zooarchaeological Evidence as a Source for Ritual Practice in Ancient Greece and Beyond, ed. Gunnel Ekroth and Jenny Wallensten, 87–99. Skrifter Utgivna av Svenska Institutet i Athen 4°55. Stockholm: Swedish Institute at Athens.

Rehint king the Sociopolitical Aftermath of a Collapsed Bronze Age Cistern 189 Leigh, David S. 2001. “Buried Artifacts in Sandy Soils: Techniques for Evaluating Pedoturbation versus Sedimentation.” In Earth Sciences and Archaeology, ed. Paul Goldberg, Vance T. Holliday, and C. Reid Ferring, 269–293. New York: Kluwer Academic/Plenum Publishers. http://dx.doi.org/10.1007/978–1–4615–1183–0_10. Letesson, Quentin. 2015. “Fire and the Holes: An Investigation of Low-Level Meanings in the Minoan Built Environment.” Journal of Archaeological Method and Theory 22(3): 713–750. http://dx.doi.org/10.1007/s10816-014-9206-y. Momigliano, Nicoletta. 2007. “Introduction.” In Knossos Pottery Handbook: Neolithic and Bronze Age (Minoan), ed. Nicoletta Momigliano, 1–8. British School at Athens Studies 14. London: British School at Athens. Oddo, Emilia. 2015. “Cross-Joins, Archaeological Sections, and the Myrtos-Pyrgos Cistern: Reconstructing a Neopalatial Stratigraphy.” In The Great Islands: Studies of Crete and Presented to Gerald Cadogan, ed. Colin F. Macdonald, Eleni Hatzaki, and Stelios Andreou, 58–62. Athens: Kapon Editions. Shillito, Lisa-Marie, and Wendy Matthews. 2013. “Geoarchaeological Investigations of Midden-Formation Processes in the Early to Late Ceramic Neolithic Levels at Catalhoyuk, , ca. 8550–8370 cal BP.” Geoarchaeology: An International Journal 28(1):25–49. http://dx.doi.org/10.1002/gea.21427.

190 Oddo & Cadogan 9

Casas Grandes (figure 9.1) is a site in northwestern Estimating the Population Chihuahua, Mexico, that the priest Obregón assured Size of Casas Grandes the King of Spain, based on his eyewitness visit in 1565, stood then “six and seven stories” high and was Empirical Issues and a “city” (Hammond and Rey 1928:206). He said it was Theoretical Consequences called “Paquimé.” Other observers later thought that parts of the main building were four to five stories high (Bandelier 1892; Brand 1933; Noguera 1930), and based David R. Wilcox on his extensive excavations, Charles Di Peso (1974; Di Peso et al. 1974), argued that the site contained some 2,300 “rooms” and housed nearly 5,000 people at its apogee. No one doubts that it is the largest single late prehistoric site in all of northwestern Chihuahua, per- haps a “primate” node in the political and economic networks of that region (and, indeed, the whole south- ern Southwest and beyond), but just how large was it and how far (on what regional or macroregional scales) its political and economic influence reached are mat- ters of empirical and theoretical debate (Di Peso 1974; McGuire 1980; Schaafsma and Riley 1999; Whalen and Minnis 2001, 2009, 2012; Wilcox 1991, 1996). In a recent paper, Whalen et al. (2010) critique the reports of earlier observers and break down Di Peso’s inferences into four subsets, showing that the facts in evidence require significant revisions of the earlier estimates. They conclude by inferring that only part of the main building stood at most three stories high and that the total “room” count was more like 1,138 rooms, implying that the maximum population was DOI: 10.5876/9781607324942.c009

191 Figure 9.1. Casas Grandes, Chihuahua, Mexico (photo by David R. Wilcox).

only about 2,500 people. Exception can be taken, however, to their model for how the main building collapsed: their assumption that the walls of this coursed-adobe [sic: caliche?]1 multistoried structure “melted” is inconsistent with my studies of the Casa Grande (figure 9.2), a compact coursed-caliche building that still stands four stories high in southern Arizona and that was partly contemporaneous with Casas Grandes (Wilcox and Shenk 1977; Wilcox and Sternberg 1981).2 Its stunning preservation provides a unique opportunity to gain insights into how such buildings were constructed— and how they fell down. In this chapter, after first discussing pertinent aspects of those findings, I revisit some of the observation reports of Obregón (Hammond and Rey 1928), John Russel Bartlett (1854), and Adolph Bandelier (1892), showing that the main building at Casas Grandes probably did, in part, stand four or five stories high. A way to test this finding is then discussed and a few of its theoretical implications are briefly explored.

192 Wlcoi x Figure 9.2. Casa Grande Ruin, Casa Grande Ruins National Monument, Coolidge, Arizona (photo by Stephen Larson in Wilcox and Shenk 1977: Figure 2; courtesy of Arizona State Museum, University of Arizona, Tucson).

T he Significance of the Casa Grande’s Formation History One of the principal findings of my two studies of the CasaG rande (Wilcox and Shenk 1977; Wilcox and Sternberg 1981) is that its three-story and four- story standing walls never were any higher, that is, they did not gradually melt down during the five centuries since people stopped living in or using the buildings. As Whalen et al. (2010:533) note in their recent analysis of Casas Grandes, this finding confirms the inference of Cosmos Mindeleff (1896, 1897), who was dispatched by the Smithsonian’s Bureau of Ethnology in 1891 to sta- bilize and preserve the building. Based on his map (figure 9.3), Mindeleff cal- culated the volume of “fill” and showed that the top of the standing walls had not melted down; this is a method to which I will return my attention shortly. By the 1890s, the Casa Grande was in imminent danger of collapsing. Several processes were in play. First, as Julian Hayden (1957) so well explained, salt erosion undercuts the walls of such buildings and large slabs of the walls can then fall out. One example of the effects of salt erosion at the Casa Grande (figure 9.4) shows a man standing on a fallen wall segment, whose location is also evident in Mindeleff ’s map. Another process that breaks down these

Estimating the Population Size of Casas Grandes 193 Figure 9.3. Topographic map of the Casa Grande by Cosmos Mindeleff (after Wilcox and Shenk 1977; courtesy of Arizona State Museum, University of Arizona, Tucson). Figure 9.4. Casa Grande before 1891, probably in the 1880s (photo by Henry Buehman; courtesy of Arizona State Museum, University of Arizona, Tucson).

buildings is the creation of large vertical cracks (figure 9.2) that I showed were not abutments but drying cracks dating back to the time of initial construc- tion.3 Vertically partitioned by such cracks, and undercut by salt erosion, large vertical wall sections eventually may yield to gravity and fall down. Another erosion process we documented in the Casa Grande is evidenced by large vertical and partially patched gouges in the inside of one of the fourth-story walls (figure 9.5) opposite the east doorway, and in the outer third-story walls (figure 9.6). All had been crudely (figure 9.7) filled in late in the use-history of the building and then had later been subjected to fur- ther erosion. The columnar nature of the erosion scars probably indicates coursing water was the culprit. This, and the patterned placement of the third-story patches to the right of the room’s central door midway along the wall between the door and the side wall, can be explained by postulat- ing that drain vents were regularly placed through the base of the parapet walls to remove rain or snowmelt puddles from the upper roofs (Wilcox and Sternberg 1981:20–21). The leaking of these drains would have concentrated the water in a rushing column of sufficient force to erode the coursed-caliche

Estimating the Population Size of Casas Grandes 195 Figure 9.5. Caliche plug in the inside the fourth-story west wall, Casa Grande Ruin (after Wilcox and Sternberg 1981: figure 3; courtesy of Arizona State Museum, University of Arizona, Tucson).

Figure 9.6. Photo of wall plug, inside the Figure 9.7. Fingerprints on wall plug, third-story east wall, Casa Grande Ruin inside the third-story west wall, Casa (in Wilcox and Shenk 1977: figure 64; photo Grande Ruin (in Wilcox and Shenk 1977: by Lynette O. Shenk; courtesy of Arizona figure 41; photo by Lynette O. Shenk; State Museum, University of Arizona, courtesy of Arizona State Museum, Tucson). University of Arizona, Tucson). walls. The collapse of the upper part of the south room-tier’s right-hand- side third-story wall (figure 9.4) fits the pattern of where we would predict another drain hole was located. Puddling of water on the roofs of multistoried buildings is a universal prob- lem that the builders of Casas Grandes must have tried to solve somehow. What saved the Casa Grande is that the roofs were removed early on, prob- ably by girdling the wood beams with fire so these valuable objects could be reused elsewhere. That stopped the erosive force of leaking drains. At Casas Grandes we were not so lucky.

Fr om the Casa Grande to Casas Grandes Next, I consider the remarks of three of the primary observers of the Casas Grandes, beginning with Obregón’s account (Hammond and Rey 1928).4 Whalen et al. (2010) suggest that Obregón approached the main building from the river side and that his estimate of six to eight stories was an illu- sion caused by mistaking the height of the terrace as part of the building’s height. In fact, however, Obregón (Hammond and Rey 1928:206) says that the Governor Ibarra party camped in the site, and his description of “magnificent patios” and “big pillars of heavy timbers” shows that he entered the build- ing. If it was never more than three stories high, as Whalen et al. (2010) infer, Obregón’s estimate of number of stories makes no sense. If it was at least four or five stories high, however, the fact that he mentions it had towers could explain the additional 1–2 stories he suggests. Obregón also reports that: This large cluster and congregation of houses is not in one place but scattered over a distance of eight leagues down the river, extending northward from the first tableland in the large mountain range. Rodrigo del Rio and I visited and explored this ridge by order of the governor. Houses continued to be found down the river and we did not lose sight of them. This seemed to be the largest and most ancient settlement in those lands. (Hammond and Rey 1928:207; see also Lange and Riley 1970:291, 293–294) This settlement pattern he interprets as what we today would call a “settle- ment system,” that is, a dispersed “city”—not a wholly unreasonable idea. He also told the King of Spain that when, using signs, they asked the local people who lived nearby in straw huts what happened to the former occupants, they were told that warfare5 had driven them away, their enemies being people to the west (Hammond and Rey 1928:207–208), whom we now call the Opata.

Estimating the Population Size of Casas Grandes 197 Figure 9.8. Possible movement of the people of Paquimé to Sonora (see Wilcox et al. 2008a; after Wilcox 2010; map prepared by David R. Wilcox).

Perhaps, then, they did not become the Opata, as Carroll Riley (1987, 2005) has suggested, but they were absorbed by them (figure 9.8).6 Nearly 300 years after Obregón’s visit, John Russel Bartlett (1854), as a member of the US-Mexican Boundary Survey, visited Casas Grandes in

198 Wlcoi x Figure 9.9. John Russel Bartlett’s sketch of Casas Grandes (after Bartlett 1854: vol. 2, facing p. 364).

the early 1850s. Whalen et al. (2010) correctly show that many of his written descriptions do not hold up to critical scrutiny, but they do not consider an image of the site (figure 9.9) he published (Bartlett 1854: Vol. 2, facing p. 364; see Wilcox et al. 2008a). Examination of Bartlett’s image reveals saguaro cactus, presumably express- ing a degree of artistic license, as they do not often grow in the Chihuahuan Desert, which should encourage some skepticism about the accuracy of the portrayed scene. Nevertheless, on top of the main mound, which it is agreed has intact walls at least 2.5 stories high, are columnar wall remnants that are at least an additional two stories high.7 That such wall remnants might be among the last portions of coursed-adobe walls to remain—that are not melted down from the top—is consistent with the findings from the CasaG rande about how such buildings fall down. And they appear to be good evidence that the main building of Casas Grandes once stood—in part—at least 4–5 stories high. Finally we come to the account of Adolph Bandelier, who systematically recorded the site during his May 1884 visit (Bandelier 1892; Lange and Riley 1970). He made one of his remarkable watercolor maps of the site, based on careful pacing, and color-coded it to show different aspects of the architecture (now in the Vatican Library, Rome, Italy). In his famous 1892 Final Report, however, he published a black-on-white image of his map (Bandelier 1892: Pl. VI; figure 9.10). As the first scientific map of the site, it compares very favorably with later maps by Donald Brand and Guevara Sanchez Blanco, as Whalen

Estimating the Population Size of Casas Grandes 199 Figure 9.10. Map of Casas Grandes by Adolph Bandelier, 1884 (after Bandelier 1892: Plate VI).

et al. (2010) show. By Bandelier’s time, the wall-remnant columns pictured by Bartlett had apparently fallen. Based on his intimate, closely observed knowl- edge of the site, Bandelier (1892) inferred cautiously that the main mound may have once stood, in part, 4–5 stories high and that the site may have housed a maximum population of about 3,000–4,000 people.

200 Wlcoi x Resolving an Empirical Issue A way exists to test Bandelier’s inference, using a method first suggested by Cosmos Mindeleff, noted above. Di Peso (1974; Di Peso et al. 1974) appar- ently shoveled out (“cleared”) the fill of room tiers he opened up, and left us, as I have discussed elsewhere (Wilcox 1999), no profiles documenting their depositional structure. Fortunately, a great deal of the main mound remains unexcavated and in the future additional room tiers could be excavated more carefully. The fact that the height of extant wall remnants in the mound are still about 6 m high means that the volume of fill, if largely made up of wall fall, is sufficient to imply original wall heights considerably greater than three stories. At a minimum, enough has been established here to question the con- clusion of Whalen et al. (2010) that the walls were never more than three sto- ries high. Just how many more “rooms” this finding implies should be added to their room count remains uncertain, but I suggest that it could be as many as 200 or so, sufficient to increase a population estimate from 2,500 to the lower end of Bandelier’s (1892:570) range of about 3,000 to 4,000 people.8

T heoretical Consequences As Whalen et al. (2010) well understand, a Casas Grandes population of 3,000 people exceeds an important comparative threshold (see Kosse 1996) with significant implications for assessments of the nature of social complex- ity at the local, regional, and macroregional scales of the Casas Grandes world (Di Peso 1974; Schaafsma and Riley 1999; Wilcox 1991). For instance, results from a comparative study of the Coalescent Communities Database of all known sites 13 rooms or larger throughout the North American Southwest (Wilcox et al. 2007; also Wilcox et al. 2006) show that the first 1,000+ room pueblos in the northern Southwest appear in the Zuni region and at Kin Tiel9 (figure 9.11) aroundad 1275–1300,10 and seem to have been occupied on that scale for only one generation. Arroyo Hondo, one of the first 1,000+ room sites in the Rio Grande, lasted only about 40 years, from ad 1300 to 1340, before being reoccupied on a smaller basis later (Shapiro 2005). It is only in the ad 1400s that numerous 1,000+ room sites were successfully founded in the northern Southwest (figure 9.12) that then persisted for hundreds of years (Wilcox et al. 2007). Before then, Casas Grandes, in the southern Southwest, was a site that prob- ably began in the middle ad 1100s (LeBlanc 1980) and surged into a 1,000+ room site in the later Medio period, perhaps as early as the late ad 1200s (cf. Whalen and Minnis 2012).11 It then persisted as the 1,000+ room primate

Estimating the Population Size of Casas Grandes 201 Figure 9.11. Standing walls of Kin Tiel (after Mindeleff and Mindeleff 1891: Plate LXIII, p. 92). Figure 9.12. Distribution of sites 13 rooms or larger in the North American Southwest, ad 1400–1450, showing Sonoran sites from Sauer 1934 (Wilcox et al. 2007: Figure 12.10; © University of Arizona Press, Tucson.) Figure 9.13. The Casas Grandes macroregional system (after Wilcox 1991: figure 8.1; © University of Arizona Press, Tucson).

center of a regional and macroregional system (figure 9.13)—perhaps the first place in the entire Southwest where a political and socioeconomic multiethnic system emerged that was capable of such adaptive success—at least for several centuries (Wilcox 1991, 1996; Wilcox et al. 2006; Wilcox et al. 2007).12 The space syntax study of Arroyo Hondo, one of the first 1,000+-room pueb- los in the northern Rio Grande region (Shapiro 2005), has shed significant light on how the northern Pueblos reorganized themselves to achieve compa- rable success in the fifteenth century. Just how the people at Paquimé did this remains unknown. The fact that their architectural spaces (e.g., figure 9.14) are often arranged so differently from the repetitive, cellular, or boxy character of

204 Wlcoi x Figure 9.14. Room configurations, Casas Grandes Ruin (after Di Peso et al. 1974:5, figure 85; courtesy of The Amerind Foundation, Inc., Dragoon, AZ; illustrated by Alice Wesche). northern pueblos (figure 9.11) may mean that their political and sociopolitical solutions were dissimilar, as well, from those of the Pueblos (Wilcox 1999; cf. Lekson 1999). Solving that problem remains one of our greatest challenges.

Estimating the Population Size of Casas Grandes 205 Acknowledgments The original images of the CasaG rande elevations were drafted by Charles Sternberg. Thanks go to the institutions holding copyright to the images shown. I also appreciate assistance provided by Curtis and Polly Schaafsma for comments on the essay, and to David Phillips for calling my attention to his excellent poster (Phillips and Bagwell 2001) about the size of Paquimé. For any errors I alone am responsible.

N otes 1. Bandelier (Lange and Riley 1970:293) says: “The walls [of the CasasG randes] are exactly alike to those of Casa Grande on the Gila, not of adobe, but a kind of marly concrete, mixed with pebbles and small stones. This cajón, as they call it here, is in layers of unequal thickness, and when it crumbles, as on the Gila, in almost cubic blocks.” This “marly concrete” is made from what is commonly called “caliche,” calcium carbonate that accumulates in desert soils around “pebbles and small stones” (Wilcox and Shenk 1977). Note that the uneven thickness of the courses is further evidence (Wilcox and Shenk 1977) that both the Casa Grande and Casas Grandes were built by hand without wooden forms, contrary to Di Peso’s (1974) conception. 2. Some support for the Whalen et al. (2010) conception of “melting” appears to come from Bartlett’s (1854: vol. 2., 354) observation that “the walls of the present build- ing are much decayed; in fact one half of their thickness is washed away, and it is only by digging below the surface that their original thickness can be seen”—so perhaps they have a smaller calcium carbonate content than the Casa Grande. However, this melting process is from the sides inwards, not from the top down! 3. A narrow caliche plug was found pressed into a segment of one of these cracks (see Wilcox and Sternberg 1981: figure 6). 4. For discussions of the larger historical context in which studies of Paquimé and Casa Grande have played important roles, see Wilcox 1986; Fowler 2000; Wilcox and Fowler 2002; Hinsley and Wilcox 2002). 5. Whalen and Minnis (2012:421) in a recent statement conclude that there is an “absence of the powerful, pervasive economic, political, and military institutions for which we see little evidence in the archaeological record.” Part of the reason for this common perception is a failure to understand what such evidence would look like in the archaeological record; for attempts to address this problem, see Wilcox and Haas 1994; LeBlanc 1999; Wilcox et al. 2006; Wilcox et al. 2008a, 2008b; Wilcox 2010, 2014). An important distinction that may be helpful in modeling political organization is made by Fish and Fish (2000) when they point out that kinship organization and civic-territorial organization should be modeled independently.

206 Wlcoi x 6. Bandelier (Lange and Riley 1970:273–280) was told by local people that the people of Casas Grandes were Opata who engaged in warfare with their neighbors to the west, who also were Opata. 7. Bartlett (1854: Vol. 2, 350) says that erect walls he saw stood of “varying height from five to thirty feet.” He estimated the height of the main mound at about 20 feet and thus infers that the highest walls still evident probably had a height of “from forty to fifty feet”—which is consistent with what his figure indicates. At 30 cm/foot, 50 feet is 15 meters high, or five stories if we follow Whalen et al. (2010) on how high each story was. 8. In light of Whalen et al.’s 2010 analysis, however, I would probably admit that the population estimate should be closer to 3,000 than 4,000, which would require a revision of the figure we used in the Coalescent Communities Database (Wilcox et al. 2007). A complicating factor is which equation to use to go from “room” counts to people counts, a problem discussed by Wilcox et al. (2007) in some detail. Whereas in the northern Southwest, as Bandelier (1892) showed, the cellular pueblo rooms are quite small, compared to those in pueblos below the Mogollon Rim, the further complication of unusual room-space shapes at Casas Grandes makes any kind of com- parability in such estimates highly problematic (see below). 9. Based on data collected by Victor and Cosmos Mindeleff (Mindeleff and Min- deleff 1891), Dennis Gilpin (personal communication) estimates that the Kin Tiel Ruin once stood three stories high and had about 1,300 rooms. Hargrave (in Haury and Hargrave 1931) excavated two kivas that dated to ad 1276; on ceramic grounds it appears to have been depopulated by about ad 1300, opening up a wide buffer zone between Hopi and Zuni and splitting the further development of White Mountain Redware into two separate traditions (see Wilcox et al. 2007). 10. Paquimé in this figure is shown as a site of 1,000+ rooms. Whalen and Minnis (2012) have recently suggested that a surge of construction to a site of this size may not have occurred until after ad 1300, but they offer no detailed discussion of the chrono- logical facts of tree-ring dates or other data to substantiate this possibility (see next endnote for further discussion). 11. Di Peso’s (1974) chronology has now been superseded. Almost immediately after its publication it was challenged, for example, by Wilcox and Shenk (1977) and Wil- cox (1986), mainly because the dating of Gila Polychrome was of such key compara- tive importance. At the Laboratory of Tree-Ring Research, University of Arizona, I was able to look at the 53 tree-ring specimens from Casas Grandes from Di Peso’s excavations that had been dated, and pointed out to Jeffrey Dean that most of the wood specimens lacked sapwood rings. Dean and John Ravesloot soon used that fact as a basis for a now widely accepted revision of the Di Peso chronology (Dean and Ravesloot 1993). More recently, Whalen and Minnis (2012) report radiocarbon dates

Estimating the Population Size of Casas Grandes 207 from excavation contexts that show the Medio period began without either Gila Poly- chrome or Ramos Polychrome in the period ad 1160–1280 (see also Wilcox 1986). Their preferred dating of Casas Grandes (Whalen et al. 2010; Whalen and Minnis 2012) is ad 1200 to 1475. However, if we also notice that the total sample of tree-ring dates from Casas Grandes is biased, being from a set of room spaces along the perimeter of the main mound, and not from the center (and perhaps oldest part) of the mound, then perhaps LeBlanc (1980) will eventually be vindicated in his suggestion that the Medio period began about ad 1150. They argue for a surge of construction after ad 1300, but that estimate is based on a strategy of using the “round number” from the radiocarbon date range of ad 1280–1330. While not unreasonable as a working hypothesis, per- haps we should entertain an alternative that the transition between the early and late Medio period occurred about ad 1275, when so many other significant changes were taking place in both the northern (LeBlanc 1999; Wilcox et al. 2007) and the southern Southwest (Wilcox et al. 2008b). That still would mean that Gila and Ramos Poly- chrome did not begin until ca. ad 1300. 12. While I admire the careful, detailed studies of the Casas Grandes world pub- lished by Whalen and Minnis (2001, 2009), and the wonderful civility with which they address interpretations different from their own, they have consistently tended to downplay the sophistication and spatial reach of Paquimé compared chiefly to Di Peso’s (1974) elaborate model, but also to my own suggestions (e.g., Wilcox 1991, 1996; Wilcox et al. 2007). Basically, what I think they underestimate is the signifi- cance of Di Peso et al.’s (1974) empirical findings of so much pottery from Paquimé’s neighbors in the “world” defined by Chihuahuan polychrome distributions, as well as El Paso Polychrome from farther east, and bison bone that may have been procured locally, but might as well have come from eastern New Mexico in the late ad 1200s where Speth (2004) has shown specialized bison procurement in a region where cop- per bells and Chihuahuan polychromes and El Paso Polychrome are found. Should we not try to test the hypothesis of the emergence of regional- and macroregional-scale multiethnic divisions of labor into “macroeconomies” (small-scale world systems; see Wilcox 1984; Chase-Dunn and Hall 1997; Chase-Dunn and Mann 1998) across all of the southern Southwest and into the western Plains in the late ad 1200s stimulated by the “market” or “gravitational force” of Paquimé and its neighbors, rather than simply ignore such possibilities?

References Bandelier, Adolph F. 1892. Final Report of Investigations among the Indians of the Southwestern United States, Part II. Papers of the Archaeological Institute of America. American Series. vol. 4. Cambridge: Archaeological Institute of America.

208 Wlcoi x Bartlett, John Russel. 1854. Personal Narrative of Exploration and Incidents in Texas, New Mexico, California, Sonora and Chihuahua. 2 vols. New York: D. Appleton and Co. Brand, Donald D. 1933. “The Historical Geography of Northwestern Chihuahua.” PhD dissertation, University of California, Berkeley. Chase-Dunn, Christopher, and Thomas D. Hall. 1997. Rise and Demise: Comparing World Systems. Boulder, CO: Westview Press. Chase-Dunn, Christopher, and Kelly Mann. 1998. The Wintu and Their Neighbors: A Very Small World-System in Northern California. Tucson: University of Arizona Press. Dean, Jeffrey S., and John C. Ravesloot. 1993. “The Chronology of Cultural Interaction in the Gran Chichimeca.” In Culture and Contact, Charles C. Di Peso’s Gran Chichimeca, ed. Anne I. Woosley and John C. Ravesloot, 83–103. Albuquerque: University of New Mexico Press. Di Peso, Charles C. 1974. Casas Grandes: A Fallen Trading Center of the Gran Chichimeca, vol. 1–3. Amerind Foundation No. 9. Dragoon, AZ: Amerind Foundation. Di Peso, Charles C., John B. Rinaldo, and Gloria Fenner. 1974. Casas Grandes: A Fallen Trading Center of the Gran Chichimeca, vol. 4–8. Amerind Foundation No. 9. Dragoon, AZ: Amerind Foundation. Fish, Suzanne K., and Paul R. Fish. 2000. “Civic-Territorial Organization and the Roots of Hohokam Complexity.” In The Hohokam Village Revisited, ed. David E. Doyel, Suzanne K. Fish, and Paul R. Fish, 373–390. Fort Collins, CO: Southwestern and Rocky Mountain Division of the American Association for the Advancement of Science. Fowler, Don D. 2000. A Laboratory for Anthropology: Science and Romanticism in the North American Southwest, 1846–1930. Albuquerque: University of New Mexico Press. Hammond, George Peter, and Agipito Rey, eds. 1928. Obregón’s History of 16th Century Exploration in Western America. Los Angeles: Wetzel Publishing. Haury, Emil W., and Lyndon Hargrave. 1931. Recently Dated Pueblo Ruins in Arizona. Smithsonian Miscellaneous Collections 82 (11). Washington, DC: Smithsonian Institution. Hayden, Julian D. 1957. Excavations, 1940, University Indian Ruins, Tucson, Arizona. Southwestern Monuments Association Technical Series No. 5. Globe, AZ: Gila Pueblo. Hinsley, Curtis M., and David R. Wilcox. 2002. “Arizona’s First Sacred Site: The Mystique of the Casa Grande, 1848–1889.” The Biennial Review 25:125–145. Tempe: Arizona State University.

Estimating the Population Size of Casas Grandes 209 Kosse, Krisztina. 1996. “Middle Range Societies from a Scalar Perspective.” In Interpreting Southwestern Diversity: Underlying Principles and Overarching Patterns, ed. Paul R. Fish and J. Jefferson Reid, 87–96. Anthropological Research Papers No. 48. Tempe: Arizona State University. Lange, Charles H., and Carroll L. Riley, eds. 1970. The Southwestern Journals of Adolph F. Bandelier, 1880–1882. Albuquerque: University of New Mexico Press. LeBlanc, Steven A. 1980. “The Dating of Casas Grandes.” American Antiquity 45(4):799–806. http://dx.doi.org/10.2307/280150. LeBlanc, Steven A. 1999. Prehistoric Warfare in the American Southwest. Salt Lake City: University of Utah Press. Lekson, Stephen H. 1999. The Chaco Meridian: Centers of Political Power in the Ancient Southwest. Walnut Creek, CA: AltaMira Press. McGuire, Randall H. 1980. “The Mesoamerican Connection in the Southwest.” Kiva 46:3–38. Mindeleff, Cosmos. 1896. Casa Grande Ruin. 13th Annual Report of the Bureau of American Ethnology, 289–319. Washington, DC: US Government Printing Office. Mindeleff, Cosmos. 1897. The Repair of Casa Grande Ruin, Arizona, in 1891. 15th Annual Report of the Bureau of American Ethnology, 315–349. Washington, DC: US Government Printing Office. Mindeleff, Victor, and Cosmos Mindeleff. 1891. A Study of Pueblo Architecture: Tusayan and Cibola. Washington, DC: US Government Printing Office, Bureau of Ethnology. Noguera, Eduardo. 1930. Ruinas Arqueológicas del Norte de México. Casas Grandes (Chihuahua), La Quemada, Chalchihuites (Zacatecas). Mexico City: Publicaciones de la Secretaria de Educatión Pública. Phillips, David A., Jr., and Elizabeth Arwen Bagwell. 2001. “How Big Was Paquimé?” Poster presented at the 66th Annual Meeting of the Society for American Archaeology, New Orleans, LA. Riley, Carroll L. 1987. The Frontier People: The Greater Southwest in the Protohistoric Period. Revised and Expanded Edition. Albuquerque: University of New Mexico Press. Riley, Carroll L. 2005. Becoming Aztlan: Mesoamerican Influence in the Greater Southwest, ad 1200–1500. Salt Lake City: University of Utah Press. Schaafsma, Curtis F., and Carroll L. Riley, eds. 1999. The Casas Grandes World. Salt Lake City: University of Utah Press. Shapiro, Jason S. 2005. Space Syntax Analysis of Arroyo Hondo Pueblo, New Mexico: Community Formation in the Northern Rio Grande. Santa Fe: School of American Research Press.

210 Wlcoi x Speth, John. 2004. Life on the Periphery: Economic Change in Late Prehistoric New Mexico. Memoirs of the Museum of Anthropology. Ann Arbor: University of Michigan. Whalen, Michael E., A. C. MacWilliams, and Todd Pitezel. 2010. “Reconsidering the Size and Structure of Casas Grandes, Chihuahua, Mexico.” American Antiquity 75(3):527–550. http://dx.doi.org/10.7183/0002-7316.75.3.527. Whalen, Michael E., and Paul E. Minnis. 2001. Casas Grandes and Its Hinterlands: Prehistoric Regional Organization in Northwest Chihuahua. Tucson: University of Arizona Press. Whalen, Michael E., and Paul E. Minnis. 2009. The Neighbors of Casas Grandes: Medio Period Communities in Northwestern Chihuahua. Tucson: University of Arizona Press. Whalen, Michael E., and Paul E. Minnis. 2012. “Ceramics and Polity in the Casas Grandes Area, Chihuahua, Mexico.” American Antiquity 77(3):403–423. http:// dx.doi.org/10.7183/0002-7316.77.3.403. Wilcox, David R. 1984. “Multi-Ethnic Division of Labor in the Protohistoric Southwest.” In Collected Papers in Honor of Harry L. Hadlock, ed. Nancy L. Fox, 141–156. Papers of the New Mexico Archaeological Society No. 9. Albuquerque: New Mexico Archaeological Society. Wilcox, David R. 1986. “A Historical Analysis of the Problem of Southwestern- Mesoamerican Connections.” In Ripples in the Chichimec Sea, ed. F. Joan Mathien and Randall H. McGuire, 9–44. Carbondale: Southern Illinois University Press. Wilcox, David R. 1991. “Changing Context of Pueblo Adaptations, ad 1250–1600.” In Farmers, Hunters and Colonists: Interaction between the Southwest and the Southern Plains, ed. Katherine A. Spielmann, 128–154. Tucson: University of Arizona Press. Wilcox, David R. 1996. “The Diversity of Regional and Macroregional Systems in the North American Southwest.” In Debating Complexity, Proceedings of the 26th Annual Chacmool Conference, ed. Daniel A. Meyer, Peter C. Dawson, and Donald T. Hanna, 375–90. Calgary, AB: University of Calgary Archaeological Association. Wilcox, David R. 1999. “A Preliminary Graph-Theoretic Analysis of Access Relationships at Casas Grandes, Chihuahua.” In The Casas Grandes World, ed. Curtis Schaafsma and Carroll Riley, 93–104. Salt Lake City: University of Utah Press. Wilcox, David R. 2010. “Changing Patterns of Resistance and Conflict in West- Central Arizona, ad 1100–1425.” Poster presented at the Southwest Symposium, Hermosillo, Sonora, Mexico. Wilcox, David R. 2014. “Microcosm and Macrocosm in Southwestern Archaeology.” In Archaeology of the Great Basin and American Southwest: Papers in Honor of Don D. Fowler, ed. Nancy J. Parezo and Joel Janetski, 260–279. Salt Lake City: University of Utah Press.

Estimating the Population Size of Casas Grandes 211 Wilcox, David R., and Don D. Fowler. 2002. “The Beginnings of Anthropological Archaeology in the North American Southwest: From Thomas Jefferson to the Pecos Conference.” Journal of the Southwest 44(2):121–234. Wilcox, David R., David A. Gregory, and J. Brett Hill. 2007. “Zuni in the Puebloan and Southwestern Worlds.” In Zuni Origins: Toward a New Synthesis of Southwestern Archaeology, ed. David A. Gregory and David R. Wilcox, 165–209. Tucson: University of Arizona Press. Wilcox, David R., David A. Gregory, J. Brett Hill, and Gary Funkhouser. 2006. “The Changing Contexts for Warfare in the North American Southwest, ad 1200–1700.” In Southwestern Interludes: Papers in Honor of Charlotte J. and Theodore R. Frisbie, ed. Regge N. Wiseman, Thomas C. O’Laughlan and Cordelia T. Snow, 203–232. Papers of the Archaeological Society of New Mexico No. 32. Albuquerque: Archaeological Society of New Mexico. Wilcox, David R., and Jonathan Haas. 1994. “The Scream of the Butterfly: Competition and Conflict in the Prehistoric Southwest.” In Themes in Southwest Prehistory, ed. George J. Gumerman, 211–238. Santa Fe: School of American Research Press. Wilcox, David R., and Lynette O. Shenk. 1977. The Architecture of the Casa Grande and its Interpretation. Arizona State Museum Archaeological Series No. 115. Tucson: University of Arizona. Wilcox, David R., and Charles Sternberg. 1981. Additional Studies of the Architecture of the Casa Grande and Its Interpretation. Arizona State Museum Archaeological Series No. 146. Tucson: University of Arizona. Wilcox, David R., Phil C. Weigand, J. Scott Wood, and Jerry B. Howard. 2008a. “Ancient Cultural Interplay of the American Southwest in the Mexican Northwest.” Journal of the Southwest 50(2):103–206. http://dx.doi.org/10.1353/jsw .2008.0016. Wilcox, David R., and Joseph P. Vogel, Tom Weiss and Sue Weiss, and Neil Weintraub. 2008b. “The Hilltop Defensive and Communication System of the Greater Prescott Area, ad 1100 to 1275, and Their Cohonina Neighbors.” In Prescott to Perry Mesa: 4,000 Years of Adaptation, Innovation, and Change in Central Arizona, ed. Christine K. Robinson, Cory Dale Breternitz and Douglas R. Mitchell, 16.1–16.39. Prescott, AZ: Sharlot Hall Museum Press.

212 Wlcoi x 10

Variability in stone tools has been the focus of pre- Biface Production at Tabun historic archaeology since the inception of the field. A variety of typologies that have been developed as a Manufacture, Maintenance, result of perceived clustering of attribute combinations and Morphological Variability has made for effective communication among prehis- torians, but only in terms of those perceived patterns, or “types.” Trouble can arise when those patterns are Gary O. Rollefson subjected to interpretation in terms of function, style, and time (see also Chase, chapter 13, this volume). Initial analysis of the bifaces from Arthur Jelinek’s Tabun excavations was undertaken between 1973 and 1978 (Rollefson 1978), resulting in, among other things, a biface typology that was based heavily on the one developed by François Bordes for the Lower and Middle Paleolithic of Europe (Bordes 2002), although several modifications were made because of local Levantine biface manifestations. While the typology facilitated information-sharing with colleagues work- ing on biface assemblages elsewhere in the Levant, there was also a sense that the utility of the typology did not extend much farther than the communication of recognized patterns of form and technology of the artifacts. In other words, understanding what bifaces meant for the producers was still elusive. In his widely used typology of bifaces, Bordes defined a cleaver as a bifacial tool with a transverse cutting edge that usually was produced by removing a tranchet flake from the distal edge (Bordes 2002:85–86).E ven so, Bordes often retained the shape of the cleaver’s DOI: 10.5876/9781607324942.c010

213 Figure 10.1. Bifacial cleaver from Lion Spring, eastern Jordan (drawing by Brian Byrd).

armature as an important aspect, referring to an “hachereau sur limande” (Bordes 2002:85) or “grand amygdaloïde, avec ‘coup de tranchet’ lateral” (2002: plate 86:1) and “biface lageniforme à ‘coup de tranchet’ oblique” (Bordes 2002: plate 87:1). Using Bordes’s typology, the 1978 tabulation of Tabun bifaces yielded a bifacial cleaver count of 175 pieces among 1,884 classifiable bifaces, and at 9.3 percent (Rollefson 1978: tables D1 and D2), this frequency was far above the average of 2–5 percent cleavers that Gilead said characterized biface assem- blages in Israel (Gilead 1973); while a notable departure, the spike in cleavers at Tabun was not taken to reflect much more than a sampling error. Continued research at Paleolithic sites in subsequent years expanded famil- iarity with other biface assemblages and introduced considerable diversity compared to the Tabun material (e.g., Rollefson 1981, 1983; Rollefson et al. 1997, 2005). As research extended into Jordan, the Tabun cleaver “anomaly” contin- ued to appear. At Lion Spring, near Azraq, cleavers (figure 10.1) accounted for more than 30 percent of the biface inventory (Rollefson 1981:6 and table 1; 1983:33–34 and table 3). This was seen as a possible consequence of bias on the part of the typologist (L. Copeland and F. Hours, personal communication 1981). Nevertheless, in her later analysis of bifaces from the Azraq area, Copeland also noted high bifacial cleaver presence at C-Spring: 18 of 62 bifaces (29.0%; Copeland 1989a: table 2, 382). Of the bifaces from Kirkbride’s excavations at

214 Rlol efson Lion Spring in the 1950s, Copeland also classed 7 of the 66 bifaces as cleavers (10.6%), although she noted in the description of biface tips that 14 of the 38 preserved distal edges bore tranchet scars (36.8%) (Copeland 1989b: table 2, 204, and table 8, 208). Clearly, biface form still held considerable sway in biface clas- sification despite the awareness of specific technological features. Surface collection and excavations at ‘Ain Soda in South Azraq (“Azraq Shishan”) in 1996 and 1997 continued to disclose high relative frequencies of cleavers (figure 10.2)—more than 90 percent (Quintero et al. 2004:3). It was at this time that Leslie Quintero and Philip Wilke joined the team investi- gating the Lower Paleolithic in Jordan, and it was particularly their insights that brought new illumination on biface production, maintenance, and asso- ciated morphological changes that have essentially made earlier biface typol- ogies obsolete. Starting at the foundation of biface production, it was the purpose of mak- ing bifaces that guided investigations by Quintero and Wilke, and there is little doubt that the principal intent of the biface knappers was to produce butchering tools that relied on fine, sharp slicing edges to enhance the process of removing skins and meat from animal carcasses (Rollefson et al. 2005:59). Only one other “type” of biface was produced in meaningful numbers: long and attenuated piercing tools characterized by sharp points with stout trian- gular or quadrilateral cross sections (e.g., Micoquian, many lanceolates, and ficrons), but their presence at butchering sites was much less remarkable in Jordan (Quintero et al. 2007:162). The same strong focus on cleaver production has also been shown at butchering sites in the Jafr Basin (Rech et al. 2007; Rollefson et al. 2006; Wilke et al. 2010). In 2005, Leslie Quintero, Philip Wilke and I reexamined the biface assem- blage from Tabun in order to reassess the analysis undertaken almost 20 years earlier. Standardized methods to describe qualitative features of bifaces are comparatively rare in the literature of Old World archaeology, for the tradi- tional approach relied heavily on classification based on shape as the primary means of conveying information relevant to technological characteristics of biface manufacture. It is evident from several typologies in earlier and cur- rent use (e.g., Ashton and White 2003; Bordes 2002; Bosinski 1967; Debénath and Dibble 1994; Kleindienst 1962; Wymer 1968) and from reports comparing biface assemblages (e.g., Gilead 1970; Gilead and Ronen 1977; Stekelis et al. 1969) that technological differences among bifaces are treated as being of sec- ondary importance to considerations of morphological variation. In Bordes’s typology, for example, 16 types emphasize the shapes of bifaces, whereas only 4 highlight technological characteristics (Type 13, cleaver; type 14, cleaver on a

Biface Production at Tabun 215 Figure 10.2. (Top) piercer and (bottom) bifacial cleaver from ‘Ain Soda, eastern Jordan (photos by L. Quintero, J. Quintero, P. Wilke, and G. Rollefson). flake; type 20, partial; and type 21, Abbevillian). As Barral and Simon (1971:8) emphasized more than 40 years ago, biface shapes would be better understood if attention were focused on features of the lithic technology that produced them. The focus of the reanalysis of the Tabun bifaces was specifically directed toward the technological features of the tools, with shape playing a decidedly minor role. The 1967–1972 excavations at Tabun Cave, Israel, recovered a total of 1,953 bifaces from in situ deposits ( Jelinek 1982). The bifaces derived from a mini- mum of 29 geological beds, although the relative numbers of these imple- ments varied considerably from one geological deposit to another.1 The original analysis focused on UnitsX –XIV, which contained more than 99 percent of the bifaces recovered from Jelinek’s excavations. The 2005 examination was restricted to analytically complete specimens numbering 921 bifaces in total. Complete bifaces were defined as those specimens for which all linear dimensions could be measured. By standardizing specific landmarks for measuring absolute dimensions of individual bifaces, one can also generate ratios of the dimensions to capture the formal variation of each piece in terms of outline, various cross sections, long section, and edge contour details2.

Results of the Reanalysis of the Tabun Biface Assemblages The Bordes-based classification of 1978 crumbled as a consequence of the 2005 investigation (adding substantiation to the aforementioned analyst’s “bias” observation of Copeland and Hours). Bordes’s morphologically defined types (4–12 and 15–18) disappeared altogether as more than 70 percent of the sample was classified as cleavers or cleavers on flakes in a broad variety of shapes (fig- ure 10.3), with the bulk of the remainder constituted by piercing tools (Bordes types 1–3: lanceolates, ficrons, and Micoquians; figure 10.4). “Diverse” types in the older classification (13% of the classifiable specimens) were highly variable in morphological and technological features, a situation that might have been predictable when classifying tools according to a sys- tem that was developed for assemblages recovered several thousand kilome- ters from Tabun. Most are more or less unique in terms of shape or specific features of retouch, including “gouge-ends,” “end-notches,”3 “proto-bifaces,” “perçoir-tipped” pieces, and burinated bifaces. Many of these were reclassified as cleavers (figure 10.5), although others were left as “other” since they appear to have been one-off tools produced for specific ad hoc uses.

Biface Production at Tabun 217 Figure 10.3. Bifacial cleavers of various shapes from Tabun. Both of the tools in the bottom row have lateral tranchet scars (photos by G. Rollefson).

Discussion Cleavers from biface assemblages from other Acheulian sites in the Levant and Europe have also probably been underreported. At Nadaouiyeh in the el- Kowm Basin of Syria’s eastern steppe, bifacial cleavers are not mentioned at all (Muhesen et al. 1998:112), but Le Tensorer has written that “ovate bifaces often present a very characteristic tranchet blow that transforms the biface into a sort of hachereau à tranchant distal” (Le Tensorer 1997:30). Published figures from the excavation reports indicate that such tools are relatively numerous (e.g., Le Tensorer 1996: figure 5-3; 1997:31). At Boxgrove, “two test pits and area Q2/C produced handaxes, over 90% of which had tranchet sharpening at the distal end” (Roberts et al. 1997:303). From the alluvial terraces of the Sahl Umm Turifa in southern Jordan, while there are huge cleavers classified as such, there are also morphologically defined biface “types” that bear clear tranchet scars at the distal end (Fabiano and Fabiani 2005: figure 15-2; figure 16-1); the same is true for bifaces from southern Jordan at Wadi al-Harad in (Pollarolo 2003), Jebel Issa (Fabiano and Primiceri 2001), and Jebel al-Hittiya (Belmonte et al. 1992), all near Wadi Rum.

218 Rlol efson Figure 10.4. Piercers from Tabun (photos by G. Rollefson).

Even for the recent excavations at Tabun there remains a persistence of mor- phological classification that trumps technological features. Matskevich notes the problem of differing definitions of what constitutes a cleaver (Matskevich 2006:335–336), ranging from Tixier’s 1956 “African” definition requiring that the tool be made on a flake, to Derek Roe’s 1969 contention that cleavers (must) have a ratio where width at the cleaver bit is close to the maximum width (i.e., nearly equal to or greater than 1.0). Matskevich chose a definition that combines the metric relationships as well as the kind of retouch on a flake

Biface Production at Tabun 219 Figure 10.5. Notched bifacial cleavers from Tabun (photo by G. Rollefson).

or a cobble.4 The figures used to illustrate the cleavers he identified clearly are cleavers, and they clearly reflect the combined definition he uses: (1) there are tranchet blows across the distal edge, and the depicted pieces occupy a morphological configuration unique to those pieces, and (2) they are all sub- rectangular and squat. One suspects that among his sample of 214 bifaces, the group he did not consider as cleavers (179, or 83.6%; Matskevich 2006: table 1) might have borne a tranchet blow but did not meet the metric requirement of his definition.

220 Rlol efson Certainly a major element in the reliance on the broad array of morpho- logical definitions of biface types is the long history of associating them as representatives of specific tools to accomplish specific goals. This was certainly a problem that held back interpretation of archaeological sites where Bordes’s flake-tool typology was appropriated, as the debate between Bordes (1966) and the Binfords (Binford and Binford 1969) demonstrated. A strong assump- tion at the time was that each defined type in the inventory was stylistically and functionally specific, and that artifacts as found in the excavations were the originally intended products. For scraper distinctions, at least, Dibble was able to refute this article of faith through experimental replication, showing that the end product of manufacture could have experienced considerable morphological change as the original tool was resharpened (Dibble 1987, 1995). The same approach to unraveling the confusion in biface typologies has begun to emerge in the past couple of decades, but the process continues to meet some objections within the prehistory community. McPherron’s (1999, 2003) reduction model was among the first attempts to approach the rigid formal configurations of traditional biface typology. Much in the vein of Dibble’s scraper demonstration, McPherron argued that biface edges became blunt as a consequence of use, and that resharpening the edges would not only result in a smaller biface as resharpening episodes unfolded, but that the overall shape would also be significantly modified (McPherron 2003:61–62). Acknowledgment of the resharpening of edges is widespread among prehis- torians (e.g., Soressi and Hays 2003), but little attention appears to have been placed on the typological consequences of such activity. McPherron’s model introduced a first glimpse of the “life history” of a biface and the changes a biface underwent, which directly challenged the “mental template” model presented by Ashton and White (2003:119–120; see also Chase, chapter 13, this volume). It might be noted that although resharpening bifaces is an acceptable modification to original bifaces to White, if a tranchet blow creates a cleaver- like shape, it is accidental and not a creation of intentional design (White 2006:377–378, 381). While McPherron’s reduction model has considerable similarity with the model of Quintero and Wilke, there are also some important differences. Like McPherron, Wilke and Quintero strongly support the notion of a “life history” for bifaces (Quintero et al. 2007:162; Wilke et al. 2010:425, 434), with repeated resharpening of the working edge as it dulls through use. Based on replication experiments, there could be up to 10 resharpenings in the life of a cleaver (Quintero et al. 2007:163). Where the models diverge, however, is in the trajectory of morphological change. McPherron (2003:57) contends that

Biface Production at Tabun 221 bifaces are originally shaped into pointed forms, which then become more and more ovate/rounded over time. Recalling that Quintero and Wilke argue that bifaces were overwhelmingly produced as butchering tools with razor- sharp transverse cutting edges (Quintero et al. 2007:162; Wilke et al. 2010:425, 437), they envision the original bifaces to be broad at the distal edge, not pointed, similar in general outline to the cleavers illustrated in Matskevich (2006: figures 1–3, 5–6). The biface design also includes “an expendable mid- section” and a basal, relatively massive grip area that often retained cortex, steep-sided natural méplats, steep unifacial retouch, and intentional batter- ing of the lateral edges (Quintero et al. 2007:162) to facilitate gripping the tool (Gowlett [2006:7–10] also emphasizes the importance of the basal grip area). As resharpening proceeded, the distal end became narrower, becom- ing more pointed over the life history of the tool (cf. Roberts et al. 1997). Once resharpening had reduced the midsection down to the grip area, the cleaver was “exhausted” and discarded. In essence, this is the direct reverse of McPherron’s trajectory. The focus by Wilke and Quintero on the trancheted cleaver edge should be placed in context. While the Paleolithic butcher at a kill-and-butcher site such as ‘Ain Soda might have considered the midsection of the biface to be expendable, the person who produced a cleaver at another kind of site may have viewed the lateral edges of the midsection to be valuable for purposes other than butchering (Soressi and Hays 2003: figures 6.3–6.7). This seems to have been the case at Tabun where, although no use-wear analysis has been undertaken, lateral edges of bifacial cleavers often bore retouch suitable for a variety of purposes. At ‘Ain Soda, on the other hand, lateral edges of the cleaver midsection were almost always blunted by heavy battering, probably to improve handling during the butchering process. The assemblage character relates to the nature of site use, and there are clear distinctions between the use of the Tabun cave and the kill-sites around Azraq in eastern Jordan and Jafr in the southern part of the country. Tabun appears to have witnessed a broad variety of tasks undertaken within its confines—of all the tools from Units XI–XIV, only 8.55 percent were bifaces, while flake tools exceeded 90 percent. ‘Ain Soda, on the other hand, clearly reflects the restricted kinds of activities associated with a butchering site, for more than half the tools there were bifaces while 43.3 percent were flake tools (including Levallois flakes, blades, and points, retouched or not; Rollefson et al. 2006:70– 71). It is likely that the “lower” cleaver percentage at Tabun compared to ‘Ain Soda, for example, is a reflection of the more domestic, less specialized char- acter of the Tabun situation.

222 Rlol efson Table 10.1. Comparison of biface dimensions (in mm) and ratios at Tabun and ‘Ain Soda. “U” refers to Tabun geological units (from Rollefson et al. 2006). Assemblage L W T W/L T/W T/L U XI 76.9 53.9 28.0 0.701 0.519 0.364 U XII 71.9 51.6 27.2 0.718 0.527 0.378 U XIII 80.7 51.1 28.4 0.633 0.556 0.352 U XIV 75.2 53.7 27.2 0.714 0.507 0.362 ‘Ain Soda 102.4 67.2 25.8 0.656 0.384 0.252

Another aspect that relates to the reduction of Tabun bifaces might lie in the proximity of raw material. The butchery site at ‘Ain Soda rested on readily available fine-quality toolstone, while nodules of suitable quality for Tabun’s needs lay 3 km up the adjacent Wadi Mughara ( Jelinek, personal communication, 2005). Overall, ‘Ain Soda bifaces are considerably larger than those from Tabun, although this could simply be a reflection of the original size of nodules available to the respective knappers. Table 10.1 pres- ents dimensions and dimension ratios of bifaces that probably reflect the relative availability of lithic resources and its possible impact on economic behavior at the two sites. The average length and width measurements reflect the potential size of nodules used at ‘Ain Soda and Tabun, although the original sizes were cer- tainly bigger than the L and W values indicate. But the ratios provide a telling story about resource husbandry at Tabun. Resharpening of the distal cutting edge of bifaces reduces length compared to width, so that the ratio becomes larger and larger (squatter in outline shape) as more resharpening episodes are conducted; notably, only Unit XIII at Tabun has a more elongated shape, but this is due to the very high number of piercer types (lanceolates, ficrons, and Micoquian types) in this unit compared to the others and compared to ‘Ain Soda; the W/L ratio might be a strong indicator of more intensive resharpen- ing of the cleaver edge at Tabun. The T/W index relates to the cross section of bifaces. Again, as tranchet blows remove more and more of the distal (thinner) end of cleavers, the T/W ratio increases, reflecting an increasing chunkiness of the biface and a pattern that repeats the situation at Tabun for the W/L index. In the case of Unit XIII, the very bulky ratio is again indicative of the pendulous shape of piercing tools, one that does not change because piercing tools likely were not often resharpened. Finally, the T/L ratio also suggests strongly that the

Biface Production at Tabun 223 Figure 10.6. Bifacial cleavers from Tabun used as flake cores (photos by G. Rollefson).

long section at Tabun is stouter than at ‘Ain Soda, indicating that inten- sive resharpening draws the working edge closer to the point of maximum thickness. The distance that must be traveled to replace toolstone at Tabun is a likely factor in the intensity of resharpening at the cave site. For ‘Ain Soda, where new stone resources were almost at arm’s reach, less conservative economies were necessary. The resource situation at Tabun might also be shown in the higher use of flakes as the blank for biface production (almost 9% at Tabun vs. less than 1% at ‘Ain Soda). Lastly, stone resources at Tabun were more intensively exploited in the sense that when bifaces had lost their utility as tools, they were often used as cores for flake production: 36–41 percent of Tabun bifaces were reused in this way, while only 3.5–12.5 percent of ‘Ain Soda handaxes were used as flake cores (figure 10.6).

Cuoncl sions The attempt to isolate recurrent standardized forms from a continuous array of metric attributes involves the establishment of arbitrary boundaries, and

224 Rlol efson while the use of mathematical formulas to make such distinctions might seem to introduce precision into the process (e.g., Bordes 2002:71–84), the results are still artificial constructs imposed on the archaeological material. For kill-and- butcher sites in areas, such as Azraq and the Jafr Basin, strictures on cleaver shapes were minimal because the functional aspect of the tool was a sharp transverse or oblique edge at the distal end of the tool; how shape changed during the resharpening process was irrelevant. Whatever shapes were recov- ered by the prehistorians, they represent only one of several stages of resharp- ening, but in general, it can be said that as modifications to the working edge increase in intensity during the life history of a cleaver, the shorter and more pointed it will become. The shape of a bifacial cleaver is, as implied by White, “accidental” (White 2006:376), but the cutting edge itself is indeed part of the “purposive design.” The notion that Paleolithic butchers had “mental templates” concerning biface forms cannot be ruled out completely, but that they had “lanceolates” vs. “cordiform” vs. “limande” vs. “ovate” and so on is difficult to demonstrate: such standardization, as mentioned before, is in our minds and not necessarily in theirs (cf. Kuhn 2010; see also Chase, chapter 13, this volume). Instead, we can be reasonably certain that those butchers had an effective cutting edge in mind and the vision of renewing it as necessary, regardless of the forms the tool underwent.

Acknowledgments I am genuinely indebted to a long and fruitful association with Leslie Quintero and Phil Wilke, whose field experience and replication expertise opened new perceptions into Middle and Late Pleistocene behaviors.

N otes 1. Thirteen beds (44.8% of those bearing bifaces) contributed only 21 bifaces (1.1% of the total number of bifaces), whereas Bed 79 produced one-eighth of all the bifaces (12.6%), and Bed 76 yielded a full quarter (25.2%) of these implements. It is evident from these data that many of the geological strata at Tabun contain bifaces that do not form integral parts of the lithic industries of those strata. This especially appears to be the case in geological Units III and V–IX. In at least one case, a heavily rolled handaxe from Bed 59, indicates redeposition from its original discard location. Similarly, bifaces in the other sparsely represented beds may have been found in derived contexts, possi- bly as the result of local erosion and redeposition in the cave or of accidental exposure

Biface Production at Tabun 225 of earlier deposits being disturbed by later inhabitants, for example. Because of the suspicious associations of these implements with the rest of the respective assemblages, they were not considered further. 2. Statistical analysis of metric dimensions is not presented in this chapter. 3. Pieces with “end-notches” occur consistently with edges bearing tranchet scars, and were not recognized originally as bifacial cleavers. It now seems clear that this end-notch represents a catastrophic event associated with a chopping motion of a bifacial cleaver against something hard, such as a bone, during butchering activity. 4. Matskevich notes a “discrepancy” in the definition of cleavers as cited in the report on bifaces from ‘Ain Soda, where he notes that “careful examination of the [two] illustrated items (Rollefson et al. 1997: Plate 2) reveals that these items may be regarded as cleavers only on the basis of their metric proportions, since they do not possess a transverse cutting edge and the general shape of the of the items is rounded throughout the circumference of the tools” (Matskevich 2006:336). First, “roundedness” was never a criterion in the classification as a cleaver. Second, while plate 2c shows a definite tran- chet blow across the distal end, the cleaver in plate 2d is unfortunately depicted upside down, so that the distinct tranchet scar is erroneously shown at the bottom.

References Ashton, N., and Mark White. 2003. “Bifaces and Raw Materials: Flexible Flaking in the British Early Paleolithic.” In Multiple Approaches to the Study of Biface Typologies, ed. Marie Soressi and Harold Dibble, 109–123. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Barral, L., and S. Simon. 1971. “Classification Automatique des Industries Préhistoriques.” Bulletin du Musée d’Anthropologie Préhistorique de Monaco 17:5–38. Belmonte, Angela, Luciana Fabiani, Luciana Mercatanti, Marco Rustioni, and Fabio Vianello. 1992. “L’Acheuleano Medio di Jebel al Hittiya nella Giordania Meridionale.” Studi per L’Ecologia del Quaternario 14:15–31. Binford, Sally, and Lewis R. Binford. 1969. “Stone Tools and Human Behavior.” Scientific American 220(4):70–84. http://dx.doi.org/10.1038/scientificamerican 0469-70. Bordes, François. 1966. “Mousterian Cultures in France.” In New Roads to Yesterday, ed. Joseph Caldwell, 78–89. New York: Basic Books. Bordes, François. 2002 [1961]. Typologie du Paléolithique Ancien et Moyen. Mémoire I: Publications de l’Institut de Préhistoire de l’Université de Bordeaux. Bosinski, Gerhard. 1967. Die Mittelpäläolitischen Funde im Westlichen Mitteleuropa. Köln, : Böhlau Verlag.

226 Rlol efson Copeland, Lorraine. 1989a. “Analysis of the Paleolithic Artifacts from the Sounding of A. Garrard at C-Spring, 1985 Season.” In The Hammer on the Rock: Studies in the Early Paleolithic of Azraq, Jordan, ed. L. Copeland and F. Hours, 325–390. BAR International Series 540 (i–ii). Oxford: British Archaeological Reports. Copeland, Lorraine. 1989b. “The Artifacts from the Sounding of D. Kirkbride at Lion Spring, Azraq, in 1956.” In The Hammer on the Rock: Studies in the Early Paleolithic of Azraq, Jordan, ed. L. Copeland and F. Hours, 172–211. BAR International Series 540 (i–ii). Oxford: British Archaeological Reports. Debénath, André, and Harold Dibble. 1994. Handbook of Paleolithic Typology: Lower and Middle Paleolithic of Europe. vol. 1. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Dibble, Harold. 1987. “The Interpretation of Middle Paleolithic Scraper Morphology.” American Antiquity 52(1):109–117. http://dx.doi.org/10.2307/281062. Dibble, Harold. 1995. “Middle Paleolithic Scraper Reduction: Background, Clarification, and Review of the Evidence to Date.” Journal of Archaeological Method and Theory 2(4):299–368. http://dx.doi.org/10.1007/BF02229003. Fabiano, Marzia, and Luciana Fabiani. 2005. “Il Paleolithico Inferiore sui Terrazzi alluviaonali di Sahl Umm Turifa (Giordania Meridionale).” Studi per L’Ecologia del Quaternario 27:21–39. Fabiano, Marzia, and Pierpaolo Primiceri. 2001. “L’Acheuleano Inferiore di Giabal el Issua nella Giordania Meridionale.” Studi per L’Ecologia del Quaternario 23:9–41. Gilead, David. 1970. “Handaxe Industries in Israel and the Near East.” World Archaeology 2(1):1–11. http://dx.doi.org/10.1080/00438243.1970.9979459. Gilead, David. 1973. “Cleavers in Early Paleolithic Industries in Israel.” Paléorient 1(1):73–86. http://dx.doi.org/10.3406/paleo.1973.901. Gilead, David, and Avraham Ronen. 1977. “Acheulian Industries from ‘Evron on the Western Galilee Coastal Plain.” Eretz-Israel 13:56–86. Gowlett, John. 2006. “The Elements of Design Form in Acheulian Bifaces: Modes, Modalities, Rules and Language.” In Axe Age: Acheulian Tool-Making from Production to Discard, ed. N. Goren-Inbar and G. Sharon, 202–221. London: Equinox. Jelinek, Arthur J. 1982. “The Tabun Cave and Paleolithic Man in the Levant.” Science 216(4553):1369–1375. http://dx.doi.org/10.1126/science.216.4553.1369. Kleindienst, Maxine. 1962. “Components of the East African Acheulian Assemblage: An Analytical Approach.” In Actes du IVème Congres Panafricain de Préhistoire et de l’Etude du Quaternaire, ed. G. Mortelmans and J. Nenquin. Annales Série, Sciences Humaines, No. 40. Tervuren, : Musée Royal de l’Afrique Central. Kuhn, Steven. 2010. “On Standardization in the Paleolithic.” In Stone Tools and the Evolution of Human Cognition, ed. A. Nowell and I. Davidson, 105–134. Boulder: University Press of Colorado.

Biface Production at Tabun 227 Le Tensorer, Jean-Marie. 1996. “Les Cultures Paléolithiques de la Steppe Syrienne: l’Example d’El-Kowm.” Annales Archeologiques Arabes Syriennes 42:43–61. Le Tensorer, Jean-Marie. 1997. Les Premiers Hommes du Désert Syrien. Paris: Éditions du Musée National d’Histoire Naturelle. Matskevich, Zinovi. 2006. “Cleavers in the Levantine Acheulian: The Case of Tabun Cave.” In Axe Age: Acheulian Tool-Making from Production to Discard, ed. N. Goren-Inbar and G. Sharon, 335–346. London: Equinox. McPherron, Shannon. 1999. “Ovate and Pointed Handaxe Assembalges: Two Points Make a Line.” Préhistoire Européenne 14:9–32. McPherron, Shannon. 2003. “Technological and Typological Variability in the Bifaces from Tabun Cave, Israel.” In Multiple Approaches to the Study of Bifacial Technologies, ed. M. Soressie and H. Dibble, 55–76. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Muhesen, Sultan, Rito Jagher, and Jean-Marie Tensorer. 1998. “Nadouiyeh, Ain Askar et Hummal (El-Kowm, Syrie Centrale) 1997.” Chronique Archéologique en Syrie 2:109–120. Pollarolo, Luca. 2003. “L’Acheuleano Medio sui Terrazzi di Uadi el Harad nella Giordania Meridionale.” Studi per L’Ecologia del Quaternario 25:23–36. Quintero, Leslie, Philip Wilke, and Gary Rollefson. 2004. “The Eastern Levant, the Pleistocene and Paleoanthropology.” ACOR Newsletter 16(1):1–3. Quintero, Leslie, Philip Wilke, and Gary Rollefson. 2007. “An Eastern Jordan Perspective on the Lower Paleolithic of the ‘Levantine Corridor.’.” Studies in the History and Archaeology of Jordan 9:157–166. Rech, Jason, Leslie Quintero, Philip Wilke, and Emily Winer. 2007. “The Lower Paleolithic Landscape of ‘Ayoun Qedim, al-Jafr Basin, Jordan.” Geoarchaeology: An International Journal 22(2):261–275. http://dx.doi.org/10.1002/gea.20165. Roberts, M., S. Parfitt, M. Pope, F. Wenban-Smith, R. MacPhail, A. Locker, and J. Stewart. 1997. “Boxgrove, West Sussex: Rescue Excavations of a Lower Palaeo­ lithic Landsurface (Boxgrove Project B, 1989–91).” Proceedings of the Prehistoric Society 63:303–58. http://dx.doi.org/10.1017/S0079497X00002474. Roe, Derek. 1969. “British Lower and Middle Paleolithic Handaxe Groups.” Proceedings of the Prehistoric Society 34:1–82. http://dx.doi.org/10.1017/S0079497 X00013840. Rollefson, Gary. 1978. “A Quantitative and Qualitative Typological Analysis of Bifaces from the Tabun Excavations, 1967–1972.” PhD dissertation, University of Arizona, Tucson. Rollefson, Gary. 1981. “The Late Acheulian Site at Fjaje, Wadi el-Bustan, Southern Jordan.” Paléorient 7(1):5–21. http://dx.doi.org/10.3406/paleo.1981.4284.

228 Rlol efson Rollefson, Gary. 1983. “Two Seasons of Excavations at Ain el-Assad, near Azraq,

Eastern Jordan, 1980–1981.” Bulletin of the American Schools of Oriental Research. American Schools of Oriental Research 252:25–34. http://dx.doi.org/10.2307/1356835. Rollefson, Gary, Leslie Quintero, and Philip Wilke. 2005. “The Acheulian Industry in the al-Jafr Basin of Southeastern Jordan.” Journal of the Israel Prehistoric Society 35:53–68. Rollefson Gary, Leslie Quintero, and Philip Wilke. 2006. “Acheulian Variability in the Southern Levant: A Contrast of the Western and Eastern Wings of the Levantine Corridor.” Near Eastern Archaeology 69:61–72. Rollefson, Gary, Douglas Schnurenberger, Leslie Quintero, Richard Watson, and Rusty Lowe. 1997. “‘Ain Soda and ’Ain Qasiya: New Late Pleistocene and Early Holocene Sites in the Azraq Shishan Area, Eastern Jordan.” In The Prehistory of Jordan II: Perspectives from 1997, ed. H. Gebel, Z. Kafafi, and G. Rollefson, 45–58. Berlin: Ex Oriente. Soressi, Marie, and Maureen A. Hays. 2003. “Manufacture, Transport, and Use of Mousterian Bifaces: A Case Study from the Perigord (France).” In Multiple Approaches to the Study of Bifacial Technologies, ed. Marie Soressi and Harold L. Dibble, 125–147. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Stekelis, Moshe, Ofer Bar-Yosef, and Tamar Schick. 1969. Archaeological Excavations at ‘Ubeidiyah, 1964–1966. Jerusalem: Israel Academy of Sciences and Humanities. White, Mark. 2006. “Axeing Cleavers: Reflections on Broad-Tipped Cutting Tools in the British Earlier Paleolithic.” In Axe Age: Acheulian Tool-Making from Production to Discard, ed. N. Goren-Inbar and G. Sharon, 365–386. London: Equinox. Wilke, Philip, Leslie Quintero, and Jason Rech. 2010. “The Acheulian Occupations at ‘Uyon al-QadΠm, al-Jafr Basin, Jordan: A Progress Summary.” Annual of the Department of Antiquities of Jordan 54:423–441. Wymer, John. 1968. Lower Paleolithic Archaeology in Britain as Represented in the Thames Valley. London: John Baker.

Biface Production at Tabun 229

Section III Cross-Cultural, Conceptual, and Experimental Perspectives

11

Mortuary events were contexts in which ritual practices Celebrating the Dead and celebrated the dead and facilitated an array of social Recrafting Social Identity objectives. In the study of mortuary practices among prehistoric hunter-gatherers, much attention has been Placing Prehistoric focused on political complexity and the identification Mortuary Practices in of leaders and elites. Until recently, considerably less Broader Social Context attention has been placed on aspects of social identity and discerning changes in social interaction as defined by attributes such as age, sex, and group membership. Brian F. Byrd and Diachronic developments in funerary rituals, as Jeffrey Rosenthal depicted in the archaeological record of mortuary events, provide a rare opportunity to gain insight into shifts in social interaction, intragroup dynamics, and ideology. We highlight this research orientation with two prehistoric examples where, based on other lines of evidence and general expectations, political complexity was increasing over time, yet mortuary practices sug- gest the opposite. These prehistoric case studies include the emergence of complex hunter-gatherers and then early agricultural village life in the southern Levant of the Near East, and the rise and persistence of complex hunter-gatherers in the San Francisco Bay area of western North America. Although widely separated in time and space, both took place in Mediterranean-type environmental set- tings, and both were correlated with larger populations, increased settlement permanence, resource intensifica- tion and storage, and rich ideological traditions. These examples were chosen because they share a number of DOI: 10.5876/9781607324942.c011

233 contextual variables in common (both environmental and economic), thereby facilitating comparative analysis. In both examples, unprecedented changes in socioeconomic strategies were correlated with a sudden, initial elaboration in mortuary practices. In each situation, a much larger segment of the society from a wider range of ages was buried with grave goods. Mortuary items were primarily personal adornment, and these practices were concentrated among younger members of society. Thereafter, despite continued intensification of economic and social activities, each case study revealed a striking decline in the number of burials with such goods and in the quantities per individual. These case studies reveal that the tempo of shifts in mortuary behavior can be rapid and multidirectional. Moreover, these changes can best be understood through consideration of the broader social context, rather than attempting to discern elites and status ascription. We argue that these mortuary events were settings in which ritual practices facilitated social integration and group solidarity, as well as active construction of social identities with respect to kin and non-kin, and peers and non-peers. These funerary practices provided an opportunity to reify new forms of community interaction and enhanced eco- nomic cooperation among peers that facilitated emerging cooperative activi- ties (see also Schroeder and Goldstein, chapter 7, this volume).

Bgack round Archaeologists have shown an enduring interest in using mortuary remains to gain insight into the past (Pearson 2001; Rakita et al. 2007; Tarlow and Nillson Stutz 2013). During the last 40 years, mortuary remains also have been a valu- able avenue of research into political complexity. Indeed, mortuary analysis was an important initial line of inquiry for processual archaeology (Chapman 2013; Rakita and Buikstra 2007). Numerous studies worldwide, including promi- nent ones in California and the Near East, analyzed mortuary contexts in the search for chiefs, elites, and the origins of ranked and stratified societies (e.g., Binford 1971; Goldstein 1976; L. King 1969; T. King 1970; Saxe 1970; Wright 1978). These studies emphasized the distinction between achieved status, where an individual’s success was dependent on his or her abilities, and ascribed status, where social position was inherited (Saxe 1970). The general expectation was that, with increasing political complexity, leaders emerged and they and their families had greater access to wealth. This research orientation within mortuary studies is not surprising, given the prominence that the emergence of inequal- ity commands among archaeologists (e.g., Flannery and Marcus 2012).

234 B yrd & Rosenthal In the Near East, some scholars argued that the onset of political complex- ity occurred with the emergence of sedentary hunter-gatherers during the Natufian period. Natufian mortuary practices (notably the sheer numbers of grave goods and the presence of children with grave goods) were cited as evi- dence of hereditary elites and a ranked society (Henry 1989:209–210; Wright 1978). Similarly in California (including the Santa Barbara, San Francisco Bay, and Central Valley areas), scholars asserted that ascribed status and hereditary elites emerged some 2,000 years ago, judging by the presence of burials, including those of children, with considerable numbers of grave goods (Gamble et al. 2001; Gardner 2013:404; C. King 1976; L. King 1969; T. King 1970, 1976). The basic assumption for these studies was that the treatment of an individ- ual at death reflects their status and role in life. This logic is then expanded to an entire burial sample to infer the structure of the society as a whole. Notably, a child buried with numerous grave goods is considered strong evidence for ascribed status and hereditary elites, since it is assumed that a young person could not have accumulated such wealth on his or her own. Subsequently, the use of mortuary assemblages to examine political com- plexity has come under criticism, largely by post-processual archaeologists stressing the difficulty in understanding the symbolic meaning of ancient mortuary practices (e.g., Boyd 2001, 2002; Joyce 2005; Shanks and Tilley 1987). As Dillehay (2012:14722) notes, “an alternative view sees mortuary patterns related to ideology, symbolism, ritual, and history rather than just social orga- nization.” Joyce (2005:139) has highlighted the need to move away from prior perspectives on interpreting mortuary events and burials noting that, “today, the body as a site of lived experience, a social body, and site of embodied agency, is replacing prior static conceptions of an archaeology of the body as a public, legible surface.” An important aspect of this perspective is the impor- tance placed on individual social identity (e.g., Amundsen-Meyer et al. 2011), the recognition that identity was actively constructed, and the awareness that social relationships and positions were multilayered and could crosscut gen- erations of the living and the deceased. Archaeologists have further noted that the cross-cultural ethnographic record is rife with examples where political complexity is not accurately expressed in burial practices; political ranking may be exaggerated, disguised, or denied in death rituals (Metcalf and Huntington 1991; Robben 2005). This is because mortuary practices, including interment of the dead and other ceremonies, were public occasions during which shared social mean- ing and memory were constructed. These were opportunities to reinforce

P lACINg Prehistoric Mortuary Practices in Broader Social Context 235 social order, promote group cohesion, and craft community-wide identities that crosscut kin lines. Burial events and related mourning ceremony activities, such as ritual feast- ing, were important venues of social interaction between and within subgroups of a community comprising families, lineages, and associations such as cults and secret societies (Hayden 2009). As such, changes in mortuary rituals, as depicted in the material record of mortuary events, provide a rare opportu- nity to explore shifts in social interaction, intragroup dynamics, and ideology. This is a particularly appropriate approach for dealing with transegalitarian groups—those best classified as falling between clearly egalitarian and decid- edly ranked societies. In the following two sections we document major changes in mortuary practices for our two examples. In doing so, we interpret these patterns with respect to the construction of social identity and broader aspects of the shift- ing dynamics of social interaction.

Near Eastern Southern Levant Our first example involves the transition to sedentary hunter-gatherers and then agricultural village life in the southern Levant—the portion of the Near East encompassed by modern-day Israel, Jordan, the Palestinian territories of Gaza and the West Bank, , and southern Syria. Archaeological data for this study are derived primarily from well-documented sites in the foot- hills overlooking the Mediterranean Sea, the Jordan Valley, and also, to a lesser extent, the uplands east of the Jordan Valley. This was, at a minimum, a four- step process taking place in the late Pleistocene and early Holocene (table 11.1). Prior to 14,500 years ago, during the Upper Paleolithic and Early-Middle Epipaleolithic (pre-Natufian), hunter-gatherer populations were foragers, most often characterized by small, relatively mobile groups, although larger aggregation sites are also present in certain localities (e.g., Garrard and Byrd 2013; Goring-Morris 1995; Goring-Morris et al. 2009; Henry 1995; Maher et al. 2012). Burials occur infrequently at sites from this 15,000-year time span, with one exception during the Middle Epipaleolithic (just prior to the Natufian) where partial remains of 11 individuals have been documented (table 11.1). Pre- Natufian burials commonly have material remains that are considered by the excavators to be grave associated (figure 11.1). These objects are typically few in number (1–3 per burial), and most often represent either utilitarian artifacts (such as fragments of exhausted ground-stone or bone tools) or potential cere- monial or ritual paraphernalia (e.g., polished pebbles, a fox skeleton, and gazelle

236 B yrd & Rosenthal Table 11.1. Southern Levant cultural sequence and corresponding mortuary data sets. Time Key Range Socioeconomic Period (cal BP) Attributes Burials Sample Studied Burial References Upper 30,000– Semimobile 24 Ayn Qasiyyah Arensburg and Paleolithic– 14,600 foragers (n = 1), Ein Gev I Bar-Yosef 1973; Middle (n = 1), Bar-Yosef 1973; Epipaleolithic (n = 2), Kharaneh Ewing 1947; (pre-Natufian) IV (n = 2), Moghr Garrard and el-Ahwal Cave Yazbeck 2008; 2 (n = 2), Nahal Kaufman 1989; Ein Geve I (n = 1), Maher 2007; Neve David Muheisen 1988; (n = 2), Ohlaho Nadel 1994; II (n = 1), Uyun Richter et al. al-Hammam 2010; Rolston (n = 11), Wadi 1982; Stock et al. Mataha (n = 1) 2005 Early 14,600– Settlement 87 Ain Mallaha Belfer-Cohen Natufian 13,000 perma- I (n = 27), El 1988; Belfer- nence and Wad B2 (n = 34), Cohen et al. initial social Hayonim I–II 1991; Byrd complexity (n = 26) and Monahan 1995; Perrot and Ladiray 1988; Wright 1978 Late Natufian 13,000– Settlement 99 Ain Mallaha Belfer-Cohen 11,700 permanence II–IV (n = 66), El 1988; Belfer- and contin- Wad B1 (n = 11), Cohen et al. ued social Hayonim III–V 1991; Byrd complexity (n = 22) and Monahan 1995; Perrot and Ladiray 1988 Early 11,700– Domestic 221 Hatoula (n = 4), Belfer-Cohen Neolithic - 10,650 plants and Jericho (n = 189), and Arensburg Pre-Pottery first villages Netiv Hagdud 1997; Kurth and Neolithic A (n = 28) Röhrer-Ertl 1981; (PPNA) Kuijt 1996: table 3 Early 10,650– Domestic 270 Ain Ghazal Kuijt 1996: table Neolithic - 8,400 animals (n = 81), Jericho 3; Kurth and Pre-Pottery and larger (n = 189) Röhrer-Ertl 1981; Neolithic B villages Rollefson et al. (PPNB) 1992 Total 701 Figure 11.1. Relative frequency of burials with grave goods as a percentage of all burials in the southern Levant.

horn cores near the head). These grave-associated objects are only present with adults and primarily with males. Items of personal adornment are absent. The start of the Early Natufian was marked by major changes in socio- economic strategies (Bar-Yosef and Valla 1991, 2013; Belfer-Cohen and Goring-Morris 2011; Byrd 2005b; Valla 1995). In the richest environmental settings, larger groups coalesced at sedentary or near-sedentary sites with a collector-based economy focusing on intensive procurement of wild plants

238 B yrd & Rosenthal and animals. This process is considered to represent the emergence of com- plex hunter-gatherers in the Near East (Belfer-Cohen and Goring-Morris 2011:210; Goring-Morris et al. 2009:103; Price and Bar-Yosef 2010:152). The dead were buried in organized cemeteries, often containing group burial areas, potentially indicating a new way of illustrating ties to the land (e.g., Kunzar 2003). The nature of grave-associated goods changes markedly with the Early Natufian—there is a shift away from the presence of few utilitarian and eso- teric items in the graves of adults, and generally males. Instead, Early Natufian graves are dominated by items of personal adornment that were placed in almost a quarter (22.9%) of graves (Byrd and Monahan 1995). Notably, these items are present in the graves of children and infants as well as young adults. Burials with grave goods were also evenly represented between males and females, despite males being significantly more frequent in the burial popu- lations (male:female ratios range from near 1:1 to over 3:1 in these samples) (Belfer-Cohen et al. 1991:414; Hershkovitz and Gopher 1990). Items of personal adornment in Early Natufian graves were dominated by Dentalium marine shell from the Mediterranean Sea. Bone beads and pen- dants (generally partridge and gazelle) and perforated animal teeth (typically fox) are also present, but most often as accessories with Dentalium. The loca- tion of decorative grave goods on the body and their incorporation into gar- ments was highly variable, and included ornamental caps, necklaces, brace- lets, and belts, and as garment leg bands and arm bands (figures 11.2 and 11.3). Notably, there is a lack of items that may symbolize rank. The most striking pattern is that the distribution of grave goods varied among age-grades (figure 11.4). Notably, teenagers (ages 13–19) were most likely to have grave goods, followed by younger adults (ages 20–35, most of whom were under 25 years old) (Byrd and Monahan 1995: table 4). In addition, all older adults lacked grave goods. The quantity of decorative items varied considerably between individuals, ranging from a single bead to almost 400 items (arrayed in a necklace, belt, and bracelet). Those with the most items included teenagers, young adults of both sexes, and some children. During the Late Natufian, substantial communities of complex hunter- gatherers persisted in the southern Levant (although it should be noted that many scholars have emphasized a greater reliance on mobility than in the Early Natufian), and substantial burial areas were present within settlements. Associated mortuary patterns differed markedly from the Early Natufian. Grave goods dramatically declined in number and were only minimally repre- sented (1.1%; see figure 11.1). Where present, they generally consisted of just a

P lACINg Prehistoric Mortuary Practices in Broader Social Context 239 Figure 11.2. Examples of Early Natufian items of personal adornment showing Dentalium shell headgear from (a) El Wad Burial Group 25 and (b) Burial 23 (drawing by Tammara Norton; based on photos in Garrod and Bate 1939: plates VI and VII).

few beads or other items, such as horn cores. Similarly, group burials became less frequent, and individual burials placed under building floors became more common. A notable exception is the so-called Shaman burial from a Late Natufian site that contained a large number of esoteric and ritual materials

240 B yrd & Rosenthal Figure 11.3. Example of Early Natufian item of personal adornment: Ain Mallaha Child Burial 43 with Dentalium belt (drawing by Tammara Norton; based on photo by Marjolaine Barazani in Perrot et al. 1988: planche XVII).

(such as select body parts of an eagle and a leopard) associated with an elderly woman (Grosman et al. 2008). The start of the Early Neolithic (Pre-Pottery Neolithic A [PPNA]) marked another period of major socioeconomic change, including the onset of large

P lACINg Prehistoric Mortuary Practices in Broader Social Context 241 Figure 11.4. Relative frequency within age categories of individuals with grave goods during the Early Natufian. Note: only Ain Mallaha and Hayonim Cave have age data for the entire sample.

agricultural villages and the construction of community-wide public architec- ture (Bar-Yosef 2001; Byrd 2005b; Kuijt and Goring-Morris 2002; Goring- Morris and Belfer-Cohen 2011; Price and Bar-Yosef 2010). These socioeco- nomic developments were associated with greater standardization in mortuary practices (Belfer-Cohen and Arensburg 1997; Kuijt 1996, 2008). Burials were regularly placed under the floors of buildings and uniformly lacked grave goods. Some graves were reopened after initial interment and skulls removed, presumably for display in ceremonial events. Later on, the skulls were reburied separately, often in clusters. This practice was elaborated during the Pre-Pottery Neolithic B (PPNB), with skulls modeled with plaster and decorated with other materials (Verhoeven 2002; Kuijt 2008) and the emergence of special- ized locations where funerary rituals were carried out (Goring-Morris 2005). In summary, the onset of sedentary hunter-gatherer communities—the Early Natufian—was correlated with group cemeteries, abundant grave goods as items of personal adornment, and multifaceted symbolic imagery in part tied to age-grades. As settlement and subsistence strategies were redefined, Late Natufian mortuary practices shifted to individual burials, and grave goods largely disappeared. With the onset of early Neolithic agricultural

242 B yrd & Rosenthal Figure 11.5. Regional trends in the frequency of dated site components per 100-year periods of the Holocene in the San Francisco Bay area (n = 667; data from Milliken et al. 2007: table 8.2).

villages, these mortuary practices were further elaborated upon, no items were interred with burials, and consistently some deceased community members had their skulls removed for use in ritual events.

San Francisco Bay Area, California Our second example examines the emergence of complex hunter-gatherers in the San Francisco Bay area of central California during the late Holocene. This sequence is generally considered to represent a continued upward trend in political complexity. Judging by a large sample of dated occupation components, the regional population increased also over time (figure 11.5). Concurrently, there was a greater reliance on lower-ranked and smaller food resources (including particular species of marine mammals, terrestrial mam- mals, birds, fish, plants, and possibly dogs) indicative of resource intensifica- tion (Broughton 1999, 2002; Broughton et al. 2007; Byrd et al. 2013; Whitaker and Byrd 2014; Wohlgemuth 1996, 2002). Tightly defined territorial packing, active landscape management, and periodic upswings in intergroup violence are also indicated (Andrushko et al. 2010; Lightfoot et al. 2013a, 2013b; Milliken

P lACINg Prehistoric Mortuary Practices in Broader Social Context 243 2006; Schwitalla et al. 2014). Drawing largely on mortuary remains, a num- ber of scholars have argued that community organization entailed nonegali- tarian social status and status ascription (Bellifemine 1997; Fredrickson 1974; Hylkema 2002:258–261; King 1974; Luby 2004; Milliken et al. 2007). Most sug- gest that these changes took place near the start of the late Holocene, although King (1974:38) and Luby (2004:18) suggest these developments occurred earlier. The Bay Area late Holocene sequence begins with the Early period (table 11.2). Middle Holocene antecedents are not as well documented, owing to site burial by sea level rise and alluviation (Bickel 1978; Lightfoot 1997; Rosenthal and Meyer 2004). TheE arly period is represented by a series of moderately sized settlements, including shell mounds well-spaced along the Bay margin (Lightfoot and Luby 2002; Milliken et al. 2007). Many of these sites appear to have been occupied at least seasonally for many years. The subsequent Middle period is generally considered to have been the heyday of mound formation, and larger sites from this time are found throughout the region (Lightfoot 1997). Longer annual stays have been hypothesized and, for near-Bay sites, greater reliance on terrestrial resources (Bartelink 2009). Bay Area popula- tions grew in size through the Late period (Milliken et al. 2007), sedentary sites flourished E( erkens et al. 2013), and material signatures of ritual activity increased (Buonasera 2013; Byrd et al. 2013). Subsequently, early Spanish colonizers documented exceedingly high population densities in the San Francisco Bay and Delta area equaled in California only by the Santa Barbara–area Chumash (Cook and Heizer 1968; Kroeber 1939; Milliken 2006, 2010). Sketchy and sometimes anecdotal ethnohistoric information reinforces the perspective of elaborate ceremo- nialism (including the secret, adult-male-only Kuksu society), and suggests both the presence of a standardized system of exchange based on clamshell beads (Chagnon 1970; Rosenthal 2011) and a system of hereditary authority held by some families (Loeb 1933). We focus on a single setting within this region, the Livermore-Amador Valley situated 20 km from the southeast side of San Francisco Bay (Rosenthal and Byrd 2006). Extensive excavations and detailed chrono- logical investigations (including 66 radiocarbon dates) have documented a series of single-component occupation horizons (typically buried rapidly by alluviation) that have yielded a substantial burial sample (table 11.2). Thus, we can hold constant the setting and ignore potentially mixed assemblages (which are pervasive in California, owing to continued reoccupation of the same locations over hundreds of years), while rigorously exploring changes in mortuary practices.

244 B yrd & Rosenthal Table 11.2. Livermore Valley, Bay Area California, cultural sequence and corresponding mortuary data sets. Key Time Range Socioeconomic Period* (cal BP) Attributes Burials Samples Studied Burial References Early Pre-2450 Initial Bay 30 ALA-483 L Bard et al. 1992; (this sample Area mounds Peak & Assoc. ~3600–2800) 1987; Wiberg 1996 Middle 2100–900 Bay Area 146 ALA-413 Wiberg 1988, (this sample mounds (n = 58), ALA- 1996 ~1700–1300) widespread, 555 L (n = 78) increased annual stays Middle- 900–700 Plant and 343 ALA-42 Tannam et al. Late animal 1993; Wiberg Transition resource 1997 intensifica- tion Late 700–250 Large 158 ALA-483 U Ext. Wiberg 1996 populations (n = 33), ALA- throughout 555 U (n = 25) Total 677

*Dating Scheme D (Groza 2002)

In the Livermore Valley, relatively few burials have been recovered from excavations at Early-period settlements. These burials infrequently con- tain grave items, and, where present, occur only in low numbers (figure 11.6). Notably, items of personal adornment are absent, and grave items were either utilitarian items (e.g., a bone awl), or potential ceremonial or ritual parapher- nalia (one elderly male had 35 esoteric items, including serpentine “splinters,” bone tubes, and a quartz crystal; this burial also had the most grave offerings overall). Only adults, both males and females, have mortuary items. Subsequent Middle-period occupation in the Livermore Valley is correlated with increased settlement intensity. This entailed heavier reliance on nuts (par- ticularly acorns) and large mammals, notably deer, in part acquired through logistical forays (Rosenthal and Byrd 2006; Wohlgemuth 2002). At the same time, the number of burials with grave goods increased to almost 30 percent. A wider range of items were interred, and most (76%) burials with grave items had either utilitarian items (such as projectile points, bifaces, bone tools, and ground stone) or items potentially representative of ceremonial/ritual activities

P lACINg Prehistoric Mortuary Practices in Broader Social Context 245 Figure 11.6. Relative frequency of burials with grave goods as a percentage of all burials in Livermore Valley, California.

(such as quartz crystals, bone whistles and tubes, and wings of birds such as crows). Items of personal adornment were present but occur in only 7 percent of all burials. These were represented by Olivella beads and Haliotis pendants and beads, both Pacific Ocean species that were acquired via trade. A number of Middle-period graves had very high quantities (more than 1,000) of mortuary items. Almost all were adult males in their 20s or 30s, along with one female in her 20s—no teenagers or children had comparable num- bers of grave goods. Olivella shell beads dominate these burials, followed by Haliotis pendants and beads. These burials included the four richest burials for any period within the valley, and the richest in the entire region. All were adult males, with the most impressive having been interred with more than 28,000 Olivella shell beads and numerous other items, including 200-odd Haliotis pendants and beads, along with other items such as bone wands, charmstones, and quartz crystals (Wiberg 1988).

246 B yrd & Rosenthal A significant economic shift took place with the Middle-Late Transition in the Livermore Valley, reflecting economic intensification of terrestrial ani- mals and plants. Large mammals declined in abundance, and small mam- mals including rodents (based on a high frequency of burned remains) were heavily exploited (Rosenthal and Byrd 2006:39–41; Wiberg 1997). Similarly, plant food exploitation entailed a significant increase in the reliance on small- seeded resources, such as grasses and forbs, which are generally considered to take more time to collect and process and have lower energetic return rates (Rosenthal and Fitzgerald 2012: table 4–14; Wohlgemuth 2002). Mortuary practices changed dramatically during this relatively short time segment. Overall, 55 percent of individuals were interred with grave goods, almost double the Middle-period frequency. Moreover, items of personal adornment shift from being infrequent to occurring in 91 percent of all burials with grave goods (i.e., almost half of all burials). Personal adornment was domi- nated by Olivella shell beads, while Haliotis pendants were also ubiquitous. They occurred as necklaces, headgear, and other decorative items (figures 11.7 and 11.8). There is a relatively even distribution in quantities of personal adornment per burial, ranging from a single bead or pendant to around 1,000 items; there is no major gap between those with many items and those with a few items. Most burials (75%), however, had 100 or fewer items. No one, however, had nearly the number of items recovered from the richest Middle-period burials. During the Middle-Late Transition, there was also significant and unprec- edented change in the age of individuals who had grave goods. Previously, during the long Middle period, the likelihood of having grave goods increased with age, peaking between the 20s and 30s and then dropping slightly among older adults (figure 11.9). During the Middle-Late Transition, however, more children (80%) were interred with grave goods than any other age category. The likelihood then decreased with age: first slightly for teenagers, and then more notably for younger adults and then older adults. During the subsequent Late period, the frequency of burials with grave goods declined significantly to near-Middle-period levels (see figure 11.6). Items of personal adornment, however, continue to occur in the vast majority of burials with grave goods (78%). These items are primarily Olivella shell beads, while Haliotis pendants are much less frequent than previously. Moreover, the percentage of burials with more than 100 items and those with at least 1,000 declined to lower than Middle-period levels. Similarly, the distribution of grave goods by age category shifted to more closely resemble Middle-period patterns than those of the Middle-Late Transition: notably, Late-period chil- dren infrequently had grave goods (32%).

P lACINg Prehistoric Mortuary Practices in Broader Social Context 247 Figure 11.7. Examples of Middle-Late Transition items of personal adornment: (a) Ala-42 Burial 112 with Haliotis necklace; (b) Ala-42 Burial 19 with Olivella-beaded headgear (drawing by Tammara Norton; based on photos by Jerry Doty in Wiberg 1997:143, 151). Figure 11.8. Example of Middle-Late Transition items of personal adornment: Ala-42 Burial 277 with Olivella-beaded cloak(?) (drawing by Tammara Norton; based on photo by Jerry Doty in Wiberg 1997:165).

In summary, the Livermore Valley sequence is characterized initially by few burials with grave goods, and when present, these items occur in small quanti- ties and do not include items of personal adornment. Although the frequency and quantity of grave items increased in the Middle period, and some exceed- ingly rich burials have been found, items of personal adornment occurred only

P lACINg Prehistoric Mortuary Practices in Broader Social Context 249 Figure 11.9. Relative frequency within age category of individuals with grave goods during the Middle period and the Middle-Late Transition.

in a small percentage of burials. In the Middle-Late Transition, mortuary prac- tices changed dramatically as many more people were buried with grave goods, and these grave goods were concentrated among children and teenagers at the expense of older adults. Most notably, items of personal adornment dominated. Finally, in the Late period, this practice declined and fewer individuals had grave goods; the average number of grave items also decreased.

Iprnter etation In both case studies, burial practices changed significantly over time. Initially, burials with offerings were infrequent, numbers of items were small, and rarely if ever were the items for personal adornment. Subsequently, grave goods became an extremely important aspect of mortuary practices and this entailed a significant shift toward emphasizing ornamentation and display items. These new developments in each case study were correlated with significant changes in the construction of social identity and in socioeconomic interaction (see also Schroeder and Goldstein, chapter 7, this volume). Shortly thereafter, the

250 B yrd & Rosenthal role of grave goods in mortuary practices declined significantly. These trends in the material evidence from burials provide a basis for interpreting why the ways in which the dead were placed in social memory changed over time. The low frequencies of mortuary offering in baseline temporal contexts from both examples (pre-Natufian southern Levant andE arly-period San Francisco Bay area) demonstrate that grave goods—as personal property or symbolizing ceremonial responsibilities—were not an important aspect of mortuary practices. Subsequently, in the Bay Area Middle period, mortu- ary practices more frequently included grave items, principally as markers of economic roles or ritual/ceremonial association (a parallel development, however, did not occur in the Near East). The corresponding Middle-period appearance of personal adornment, primarily as high numbers of items with adult males, suggests the presence of wealthy, powerful, or highly venerated community leaders. This trend is most consistent with achieved status, rather than ascribed status—indicative of skillful individuals who rose to leadership roles owing to their abilities rather than their ancestry (see also Milliken and Bennyhoff 1993). Subsequently in both settings (Early Natufian and Middle-Late Transition), a large segment of the society, generally younger members, had grave goods, and these were overwhelmingly items of personal adornment. We argue that this is not evidence of hereditary elites, a hierarchical society, and ascribed status. There are too many individuals with grave goods to represent just elites; those with grave goods were not spatially clustered (which might sug- gest high-status extended families); and although children had grave goods, older adults typically lacked them. How could children outrank their elders if hereditary elites were present? Early Natufian and Middle-Late Transition mortuary practices clearly were complex, but we argue that they functioned primarily to reinforce the crafting of new forms of social identity and to mark social roles and memberships tied to age-grade patterns, crosscutting kin lines. This argument is consistent with an emphasis on items of personal display (rather than the earlier focus on util- itarian or esoteric items). That the majority of these decorative items appear to be attached to articles of clothing indicates that these were not offerings made by others during mortuary ceremonies but rather represent personal items of the deceased. Older adults lacked grave goods, since they had moved through all of the age-grade stages and thus had no markers of age-grade membership buried with them. Similar social systems have been noted ethnographically in other contexts, such as among Loikop pastoralists of northern Kenya (Larick 1987) and Arapesh agriculturalists in New Guinea (Tuzin 1980).

P lACINg Prehistoric Mortuary Practices in Broader Social Context 251 The sudden appearance of these age-grade-related mortuary practices can be considered tangible evidence of new social strategies that emerged during times of rapid culture change. They functioned to bind communities together within a set of increasingly multifaceted social roles and obligations. These social obligations were tied in part to new subsistence strategies that entailed greater collective effort in procurement, processing, and preparation (i.e., resource intensification). The ensuing decline in these particular mortuary practices in both contexts suggests that aspects of social identity and community rules had changed fur- ther, and the impetus for marking particular individuals and their age-grades declined in importance. In the Near East, the complete demise in the use of grave goods by the Late Natufian may signify a concerted effort to empha- size similarities and mask differences between individuals throughout society. Subsequent early Neolithic mortuary rituals, with their emphasis on ancestors, linked the living to the dead and undoubtedly served to enhance community- wide harmony and promote an egalitarian ethos (Byrd 2005a:134–135; Kuijt 2008). Skulls subject to special treatment may have denoted representatives of kin groups such as lineages (and included females, young adults, and some children [Bonogofsky 2001]), and undoubtedly played an important role as ritual furniture during mortuary-related events. They would have been highly evocative anchors that aided the creation of new aspects of social identity and meaning through the collective veneration of ancestors. The corresponding decline in grave goods during the Late period in the Livermore Valley is somewhat more difficult to interpret, and investigation of other Bay Area settings is clearly required to fully understand these shifts in mortuary practices. Part of the challenge for the Bay Area is that these devel- opments were occurring within a very compressed time frame (especially as compared to the Near East), and throughout the sequence, mortuary practices entailed multiple lines of symbolic imagery. In the Late period, however, other aspects of ritual activity appear to have become more elaborate, perhaps indi- cating how other facets of community interaction played a more important role in recrafting of social identity and reinforcing social rules (e.g., Buonasera 2013; Byrd et al. 2013). The overall pattern (from the Early to the Late period), however, is supported by results from the northeast portion of the Bay Area (Atchley 1994; Fredrickson 1974). In both of our case studies, changes in mortuary practices were tied to important society-wide changes (larger populations, increased settlement per- manence, and economic intensification and extensification) and the more com- plex social relationships that ensued. Mortuary rituals may have also emerged

252 B yrd & Rosenthal as a stabilizing force—or a leveling mechanism—that constrained the devel- opment of centralized power and authority. Yet throughout both prehistoric sequences, political complexity undoubtedly increased. Community leaders undoubtedly emerged, and possibly elites as well, but these developments may have been de-emphasized during mortuary activities. The documented changes in mortuary patterns do not demonstrate increased political complex- ity. Instead they provide insight into developments in the construction of new aspects of social identity, and novel, yet successful, systems of social interaction that reinforced group cohesion. Such mortuary-related ritual events were a stabilizing force during periods of unprecedented culture change, reinforc- ing a community-wide identity initially through the use of items of personal adornment within younger age-grades.

Cuoncl sions We have argued that mortuary practices among transegalitarian groups can serve a variety of social roles and can vary independently of political com- plexity. In our Near Eastern example, political complexity increased mark- edly over time, while mortuary practices initially became more elaborate with ubiquitous grave goods, and then became materially less intricate and lacked grave goods. The relationship between political complexity and mortuary practices in our Bay Area example follows a similar trajectory, although to a lesser degree. Both examples are contrary to the typical expectation of more grave goods with increased political complexity and the emergence of ascribed status and wealthy elites. This disjunction between grave-good richness and political complexity occurred because other aspects of social interaction dominated the symbolism embedded in rituals celebrating and memorializing the dead. We have argued that social identity typically dictated how individuals were treated in death. Notably, the prominence of age-grade membership marking—through the use of items of personal adornment—took place during periods of increased social interaction, and this functioned to reinforce new social roles and rules, and to bind communities together. Mortuary rituals have always been about the living, and they do not change in isolation. Mortuary practices were less elaborate at the start of the Near Eastern and Bay Area sequences because in each example social interaction was less complex, in part due to lower population densities, and economic interaction was more often focused among closely related individuals. Social relations may have been structured more by familial relationships than by

P lACINg Prehistoric Mortuary Practices in Broader Social Context 253 larger community or supracommunity-structured interaction. Subsequently, increased settlement permanence, resource intensification, storage, and ter- ritorial compression entailed greater interfamily and intergroup social interac- tion. For these adaptive strategies to be successful, new forms of social identity, more social roles, and further rules of interaction were needed to successfully navigate this novel social landscape, maintain harmony, resolve disputes, and organize cooperative economic enterprises. Initial elaboration in mortuary practices tied to recrafting social identity was a logical corollary of these socio- economic developments. We also suspect that the ensuing decline in the scale of mortuary symbolism, particularly with respect to the quantity of material interred in graves, is probably a frequent trend cross-culturally. Increasingly, archaeologists are exploring aspects of social identity (e.g., Amundsen-Meyer et al. 2011), often focusing on mortuary data to gain more nuanced insights into this complex topic ( Joyce 2005). Temple et al.’s (2011) study of mortuary remains of complex Jomon hunter-gatherers explored the role of age-grades and other factors in the construction of social iden- tity. Others have considered the role of ornamentation in social identity, and the concurrent construction of individual and group identity through funer- ary rituals (McCafferty and McCafferty 2011; Potter and Perry 2011). Notably, Dillehay (2012) has stressed that mortuary patterns may reveal selective com- plexity that can best be understood within a larger social and economic con- text (Marquet et al. 2012). We have argued that considerable insight into social interaction can be gleaned from a detailed analysis of mortuary remains. This study has high- lighted the dynamic nature of mortuary practices and the value of diachronic studies that can identify sudden and often short-lived changes. Moreover, mortuary practices cannot be studied in isolation but must be interpreted in the broader ideological, social, and economic contexts of a prehistoric society. By studying mortuary practices in a wider context and undertaking cross- cultural comparisons, our insight into the dynamics of social complexity and interaction will be greatly enhanced.

References Amundsen-Meyer, Lindsay, Nicole Engel, and Sean Pickering, eds. 2011. Identity Crisis: Archaeological Perspectives on Social Identity. Proceedings of the 42nd (2010) Annual Chacmool Archaeology Conference. Calgary, AB: University of Calgary. Andrushko, Valerie A., Al W. Schwitalla, and Phillip L. Wilke. 2010. “Trophy-Taking and Dismemberment as Warfare Strategies in Prehistoric Central California.” American Journal of Physical Anthropology 141:83–96.

254 B yrd & Rosenthal Arensburg, Baruch, and Ofer Bar-Yosef. 1973. “Human Remains from Ein Gev I, Jordan Valley, Israel.” Paléorient 1(2):201–206. http://dx.doi.org/10.3406/paleo .1973.4165. Atchley, Sara M. 1994. “A Burial Analysis of the Hotchkiss Site (CA-CCO-138).” Master’s thesis, Sonoma State University, Rohnert Park, CA. Bard, James C., Colin I. Busby, Michael R. Fong, Robert M. Harmon, Melody E. Tannam, Donna M. Garaventa, and Angela M. Banet, S. A. Jarvis, Steven J. Rossa, and Ranbir S. Sidhu. 1992. “Archaeological Site Testing Report, CA-ALA-483, Laguna Oaks Project, Pleasanton, Alameda County, California.” Report on file, Northwest Information Center. Rohnert Park, CA: Sonoma State University. Bartelink, Eric J. 2009. “Late Holocene Dietary Change in the San Francisco Bay Area: Stable Isotope Evidence for an Expansion in Diet Breadth.” California Archaeology 1(2):227–251. http://dx.doi.org/10.1179/cal.2009.1.2.227. Bar-Yosef, Ofer. 1973. “Nahal Ein Geve I, Preliminary Report.” Mitekufat Haeven 11:1–7. Bar-Yosef, Ofer. 2001. “From Sedentary Foragers to Village Hierarchies: The Emergence of Social Institutions.” In The Origin of Human Social Institutions, ed. W. G. Runciman, 1–38. Proceedings of the British Academy 110. Oxford: Oxford University Press. Bar-Yosef, Ofer, and François Valla, eds. 1991. The Natufian Culture in the Levant. Ann Arbor: International Monographs in Prehistory. Bar-Yosef, Ofer, and François Valla, eds. 2013. Natufian Foragers in the Levant: Terminal Pleistocene Social Changes in Western Asia. Archaeological Series 19. Ann Arbor: International Monographs in Prehistory. Belfer-Cohen, Anna. 1988. “The Natufian Graveyard in Hayonim Cave.” Paléorient 14(2):297–308. http://dx.doi.org/10.3406/paleo.1988.4476. Belfer-Cohen, Anna, and Baruch Arensburg. 1997. “The Human Remains from Netiv Hagdud.” In An Early Neolithic Village in the Jordan Valley, Part I: The Archaeology of Netiv Hagdud, ed. Ofer Bar-Yosef and Avi Gopher, 201–208. American School of Prehistoric Research Bulletin 43. Cambridge: Peabody Museum of Archaeology and Ethnology, Harvard University. Belfer-Cohen, Anna, and A. Nigel Goring-Morris. 2011. “Becoming Farmers: The Inside Story.” Current Anthropology 52(S4):S209–220. http://dx.doi.org/10.1086 /658861. Belfer-Cohen, Anna, Lynne A. Schepartz, and Baruch Arensburg. 1991. “New Biological Data for the Natufian Populations in Israel.” In The Natufian Culture in the Levant, ed. Ofer Bar-Yosef and Francois Valla, 411–424. Ann Arbor: International Monographs in Prehistory.

P lACINg Prehistoric Mortuary Practices in Broader Social Context 255 Bellifemine, Viviana. 1997. “Mortuary Variability in Prehistoric Central California: A Statistical Study of the Yukisma Site, CA-SCL-38.” Master’s thesis, San Jose State University, San Jose, CA. Bickel, Polly M. 1978. “Changing Sea Levels along the California Coast: Anthropological Implications.” Journal of California Anthropology 5:6–20. Binford, Lewis R. 1971. “Mortuary Practices: Their Study and Their Potential.” In Approaches to the Social Dimension of Mortuary Practices, ed. James A. Brown, 6–29. Memoirs of the Society for American Archaeology 25. Menasha, WI: Society for American Archaeology. Bonogofsky, M. 2001. “Cranial Modeling and Neolithic Bone Modification at ‘Ain Ghazal: New Interpretations.” Paléorient 27(2):141–146. http://dx.doi.org/10.3406 /paleo.2001.4734. Boyd, Brian. 2001. “The Natufian Burials from el-Wad, Mount Carmel: Beyond Issues of Social Differentiation.” Mitekufat Haeven 31:185–200. Boyd, Brian. 2002. “Ways of Eating/Ways of Being in the Later Epipaleolithic (Natufian) Levant.” In Thinking Through the Body: Archaeologies of Corporeality, ed. Y. Mailakis, M. Pluciennik, and S. Tarlow, 137–152. New York: Kluwer Academic/ Plenum Press. http://dx.doi.org/10.1007/978-1-4615-0693-5_8. Broughton, Jack M. 1999. Resource Depression and Intensification during the Late Holocene, San Francisco Bay: Evidence from the Emeryville Shellmound Vertebrate Fauna. Anthropological Records 32. Berkeley, CA: University of California. Broughton, Jack M. 2002. “Prey Spatial Structure and Behavior Affect Archaeological Tests of Optimal Foraging Models: Examples from the Emeryville Shellmound Vertebrate Fauna.” World Archaeology 34(1):60–83. http://dx.doi.org /10.1080/00438240220134269. Broughton, Jack M., Daniel Mullins, and Tamara Ekker. 2007. “Avian Resource Depression or Intertaxonomic Variation in Bone Density? A Test with San Francisco Bay Avifaunas.” Journal of Archaeological Science 34(3):374–391. http://dx .doi.org/10.1016/j.jas.2006.05.013. Buonasera, Tammy Y. 2013. “More than Acorns and Small Seeds: A Diachronic Analysis of Mortuary Associated Ground Stone from the South San Francisco Bay Area.” Journal of Anthropological Archaeology 32(2):190–211. http://dx.doi.org /10.1016/j.jaa.2013.01.003. Byrd, Brian F. 2005a. Early Village Life at Beidha, Jordan: Neolithic Spatial Organization and Vernacular Architecture. British Academy Monographs in Archaeology 12. Published for the Council for British Research in the Levant. Oxford: Oxford University Press.

256 B yrd & Rosenthal Byrd, Brian F. 2005b. “Reassessing the Emergence of Village Life in the Near East.” Journal of Archaeological Research 13(3):231–290. http://dx.doi.org/10.1007/s10814 -005-3107-2. Byrd, Brian F., Anna Cornellas, Jelmer W. Eerkens, Jeffrey Rosenthal, Tim R. Carpenter, Alan Leventhal, and Jennifer A. Leonard. 2013. “The Role of Canids in Ritual and Domestic Contexts: New Ancient DNA Insights from Complex Hunter-Gatherer Sites in Prehistoric Central California.” Journal of Archaeological Science 40(4):2176–2189. http://dx.doi.org/10.1016/j.jas.2012.12.020. Byrd, Brian F., and Christopher Monahan. 1995. “Death, Mortuary Ritual, and Natufian Social Structure.” Journal of Anthropological Archaeology 14(3):251–287. http://dx.doi.org/10.1006/jaar.1995.1014. Chagnon, Napoleon A. 1970. “Ecological and Adaptive Aspects of California Shell Money.” Annual Reports of the University of California Archaeological Survey 12:1–25. Chapman, Robert. 2013. “Death, Burial, and Social Representation.” In The Oxford Handbook of the Archaeology of Death and Burial, ed. Sarah Tarlow and Liv Nillson Stutz, 47–58. Oxford: Oxford University Press. Cook, Sherbourne F., and Robert F. Heizer. 1968. “Relationship among Houses, Settlement Areas, and Population in Aboriginal California.” In Settlement Archaeology, ed. K. C. Chang, 79–116. Palo Alto, CA: National Press Books. Dillehay, Tom D. 2012. “Selective Complexity and Adaptive Mortuary Behavior.” Proceedings of the National Academy of Sciences of the United States of America 109(37):14722–14723. http://dx.doi.org/10.1073/pnas.1212756109. Eerkens, Jelmer W., Brian F. Byrd, Howard J. Spero, and Anna Fritschi. 2013. “Stable Isotope Reconstructions of Shellfish Harvesting Seasonality in an Estuarine Environment: Implications for Late Holocene San Francisco Bay Settlement Patterns.” Journal of Archaeological Science 40:2014–2024. http://dx.doi.org/10.1016/j .jas.2012.12.018. Ewing, J. Franklin. 1947. “Preliminary Note on the Excavations at the Paleolithic Site of Ksar ‘Akil, Republic of Lebanon.” Antiquity 21(84):186–196. http://dx.doi.org/10 .1017/S0003598X00016677. Flannery, Kent, and Joyce Marcus. 2012. The Creation of Inequality: How Our Prehistoric Ancestors Set the Stage for Monarchy, Slavery, and Empire. Cambridge: Harvard University Press. http://dx.doi.org/10.4159/harvard.9780674064973. Fredrickson, David A. 1974. “Social Change in Prehistory: A Central California Example.” In Antap: California Indian Political and Economic Organization, ed. Lowell J. Bean and Thomas F. King, 57–73. Anthropological Papers 2. Menlo Park, CA: Ballena Press.

P lACINg Prehistoric Mortuary Practices in Broader Social Context 257 Gamble, Lynn H., Phillip L. Walker, and Glenn S. Russell. 2001. “An Integrative Approach to Mortuary Analysis: Social and Symbolic Dimensions of Chumash Burial Practices.” American Antiquity 66(2):185–212. http://dx.doi.org/10.2307/2694605. Gardner, Karen Smith. 2013. “Diet and Identity among the Ancestral Integrating Stable Isotope Analysis Mortuary Context at the Yukism Mound (CA-SCL-38).” Master’s thesis, California State University, Chico, CA. Garrard, Andrew N., and Brian F. Byrd. 2013. Project Background and the Late Palaeolithic (Geological Context and Technology), vol. 1. Beyond the Fertile Crescent: Late Palaeolithic and Neolithic Communities of the Jordanian Steppe: The Azraq Basin Project. Council for British Research in the Levant Supplemental Series. Oxford: Oxbox Books. Garrard, Andrew, and Corine Yazbeck. 2008. “Qadisha Valley Prehistoric Project (Lebanon): The 2004–2008E xcavations at Moghr El-Ahwal.” Bulletin d’Archeologie et d’Architecture Libanaise 12:5–15. Garrod, Dorothy, and Dorothea M. A. Bate. 1939. The Stone Age of Mount Carmel: Excavations at the Wady el-Mughara. vol. 1. Oxford: Clarendon Press. Goldstein, Lynne. 1976. “Spatial Structure and Social Organization.” PhD disserta- tion, Northwestern University, Evanston, IL. Goring-Morris, A. Nigel. 1995. “Complex Hunter/Gathers at the End of the Paleolithic (20,000–10,000 BP).” In The Archaeology of Society in the Holy Land, ed. T. E. Levy, 141–168. London: Leicester University Press. Goring-Morris, A. Nigel. 2005. “Life, Death and the Emergence of Differential Status in the Near Eastern Neolithic: Evidence from Kfar HaHoresh, Lower Galilee, Israel.” In Archaeological Perspectives on the Transmission and Transformation of Culture in the Eastern Mediterranean, ed. Jeffery Clark, 89–105. Oxford: Levant Supplementary Series CBRL and Oxbow Books. Goring-Morris, A. Nigel, Erella Hovers, and Anna Belfer-Cohen. 2009. “The Dynamics of Pleistocene and Early Holocene Settlement Patterns and Human Adaptations in the Levant: An Overview.” In Transitions in Prehistory: Essays in Honor of Ofer Bar-Yosef, ed. John J. Shea and Daniel E. Lieberman, 185–252. Oxford: Oxbow Books. Goring-Morris, A. Nigel, and Anna Belfer-Cohen. 2011. “Neolithization Processes in the Levant: The Outer Envelope.” Current Anthropology 52(S4):S195–208. http:// dx.doi.org/10.1086/658860. Grosman, Leore, Natalie S. Munro, and Anna Belfer-Cohen. 2008. “A 12,000-Year- Old Shaman Burial from the Southern Levant (Israel).” Proceedings of the National Academy of Sciences of the United States of America 105(46):17665–17669. http://dx.doi .org/10.1073/pnas.0806030105.

258 B yrd & Rosenthal Groza, Randall Gannon. 2002. “An AMS Chronology for Central California Olivella Shell Beads.” Master’s thesis, San Francisco State University, San Francisco, CA. Hayden, Brian. 2009. “Funerals as Feasts: Why Are They So Important?” Cambridge Archaeological Journal 19(01):29–52. http://dx.doi.org/10.1017/S095977430900002X. Henry, Don O. 1989. From Foraging to Agriculture: The Levant at the End of the Ice Age. Philadelphia: University of Pennsylvania Press. Henry, Don O. 1995. Prehistoric Cultural Ecology and Evolution: Insights from Southern Jordan. New York: Plenum Press. http://dx.doi.org/10.1007/978-1-4757-2397-7. Hershkovitz, Israel, and Avi Gopher. 1990. “Paleodemography, Burial Customs, and Food Producing Economy at the Beginning of the Holocene: A Perspective from the Southern Levant.” Mitekufat Haeven 23:9–47. Hylkema, Mark G. 2002. “Tidal Marsh, Oak Woodlands, and Cultural Florescence in the Southern San Francisco Bay Region.” In Catalysts to Complexity: Late Holocene Societies of the California Coast, ed. Jon M. Erlandson and Terry L. Jones, 205–232. Perspectives in California Archaeology 6. Los Angeles: Institute of Archaeology, University of California. Joyce, Rosemary A. 2005. “Archaeology of the Body.” Annual Review of Anthropology 34(1):139–158. http://dx.doi.org/10.1146/annurev.anthro.33.070203.143729. Kaufman, Daniel. 1989. “Observations on the Geometric Kebaran: A View from Neve David.” In Investigations in Southern Levantine Prehistory, ed. Ofer Bar-Yosef and B. Vandermeersch, 275–286. BAR International Series 497. London: British Archaeological Reports. King, Chester D. 1976. “Chumash Inter-Village Economic Exchange.” In Native Californians: A Theoretical Perspective, ed. Lowell Bean and Thomas Blackburn, 289–318. Ramona, CA: Ballena Press. King, Linda B. 1969. “The Medea Creek Cemetery (LAN–243): An Investigation of Social Organization from Mortuary Practices.” Reports of the University of California Archaeological Survey 11:23–68. Los Angeles: Department of Anthropology, University of California. King, Thomas F. 1970. The Dead at Tiburon. Occasional Paper 2. Marin County, California: Northwestern California Archaeological Society. King, Thomas F. 1974. “The Evolution of Status Ascription around San Francisco Bay.” In ?Antap: California Indian Political and Economic Organization, ed. Lowell J. Bean and Thomas F.K ing, 35–54. Anthropological Papers 2. Menlo Park, CA: Ballena Press. Kroeber, Alfred L. 1939. Cultural and Natural Areas of Native North America. Berkeley, CA: University of California Press.

P lACINg Prehistoric Mortuary Practices in Broader Social Context 259 K uijt, I. 1996. “Negotiating Equality through Ritual: A Consideration of Late Natufian and PrePottery Neolithic A Period Mortuary Practices.” Journal of Anthropological Archaeology 15(4):313–336. http://dx.doi.org/10.1006/jaar.1996.0012. Kuijt, I. 2008. “The Regeneration of Life: Neolithic Structures of Symbolic Remembering and Forgetting.” Current Anthropology 49(2):171–197. http://dx.doi .org/10.1086/526097. Kuijt, Ian, and Nigel Goring-Morris. 2002. “Foraging, Farming, and Social Complexity in the Pre-Pottery Neolithic of the Southern Levant: A Review and Synthesis.” Journal of World Prehistory 16(4):361–440. http://dx.doi.org/10.1023 /A:1022973114090. Kunzar, L. A. 2003. “Sacred Sites and Profane Conflicts: The Use of Burial Facilities and Other Sacred Locations as Territorial Markers – Ethnographic Evidence.” In Theory, Method, and Practice in Modern Archaeology, ed. Robert J. Jeske and Donald K. Charles, 270–286. London: Praeger. Kurth, Gottfried, and Olav Röhrer-Ertl. 1981. “Appendix B. On the Anthropology of Mesolithic to Chalcolithic Human Remains from the Tell es-Sultan in Jericho, Jordan.” In The Architecture and Stratigraphy of the Tell: Excavation at Jericho, vol. III. ed. Kathleen M. Kenyon, 407–500. London: British School of Archaeology in Jerusalem. Larick, R. 1987. “The Circulation of Spears among Loikop Cattle Pastoralists of Samburu District, Kenya.” Research in Economic Anthropology 9:143–166. Lightfoot, Kent G. 1997. “Cultural Construction of Coastal Landscapes: A Middle Holocene Perspective from San Francisco Bay.” In Archaeology of the California Coast during the Middle Holocene, ed. Jon M. Erlandson and Michael A. Glassow, 129–142. Perspectives in California Archaeology 4. Los Angeles: Institute of Archaeology, University of California. Lightfoot, Kent G., Rob Q. Cuthrell, Chuck J. Striplen, and Mark G. Hylkema. 2013a. “Rethinking the Study of Landscape Management Practices among Hunter-Gatherers in North America.” American Antiquity 78(2):285–301. http://dx .doi.org/10.7183/0002-7316.78.2.285. Lightfoot, Kent G., and Edward M. Luby. 2002. “Late Holocene in the San Francisco Bay Area: Temporal Trends in the Use and Abandonment of Shell Mounds in the East Bay.” In Catalysts to Complexity: Late Holocene Societies of the California Coast, ed. Jon M. Erlandson and Terry L. Jones, 263–281. Perspectives in California Archaeology 6. Los Angeles: Institute of Archaeology, University of California. Lightfoot, Kent G., Lee M. Panich, Tsim D. Schneider, Sara L. Gonzalez, Matthew A. Russell, Darren Modzelewski, Theresa Molino, and Elliot H. Blair. 2013b. “The

260 B yrd & Rosenthal Study of Indigenous Political Economies and Colonialism in Native California: Implications for Contemporary Tribal Groups and Federal Recognition.” American Antiquity 78(1):89–103. http://dx.doi.org/10.7183/0002-7316.78.1.89. Loeb, Edwin M. 1933. “The Eastern Kuksu Cult.” University of California Publications in Archaeology and Ethnology 33:139–232. Luby, Edward M. 2004. “Shell Mounds and Mortuary Behavior in the San Francisco Bay Area.” North American Archaeologist 25(1):1–33. http://dx.doi.org/10.2190/N8B9 -TWPB-QFP9-74BD. Maher, Lisa A. 2007. “2005 Excavations at the Geometric Kebaran Site of ‘Uyun al-Hammam, al-Koura District, Northern Jordan.” Annual of the Department of Antiquities of Jordan 51:263–272. Maher, Lisa A., Tobias Richter, and Jay T. Stock. 2012. “The Pre-Natufian Epipaleolithic: Long-Term Behavioral Trends in the Levant.” Evolutionary Anthropology 21(2):69–81. http://dx.doi.org/10.1002/evan.21307. Marquet, Pablo A., Calogero M. Santoro, Claudio Latorre, Vivien G. Standen, Sebastián R. Abades, Marcelo M. Rivadeneira, Bernardo Arriaza, and Michael E. Hochberg. 2012. “Emergence of Social Complexity among Coastal Hunter- Gatherers in the Atacama Desert of Northern .” Proceedings of the National Academy of Sciences of the United States of America 109(37):14754–14760. http://dx.doi .org/10.1073/pnas.1116724109. McCafferty, Geoffrey, and Sharisse McCafferty. 2011. “Bling and Things: Ornamentation and Identity in Pacific Nicaragua.” In Identity Crisis: Archaeological Perspectives on Social Identity, ed. Lindsay Amundsen-Meyer, Nicole Engel and Sean Pickering, 178–188. Proceedings of the 42nd Annual Chacmool Archaeology Conference. Calgary, AB: University of Calgary. Metcalf, Peter, and Richard Huntington. 1991. Celebrations of Death: The Anthropology of Mortuary Ritual. Cambridge: Cambridge University Press. http://dx.doi.org/10 .1017/CBO9780511803178. Milliken, Randall. 2006. “The Central California Ethnographic Community Distribution Model, Version 2.0, with Special Attention to the San Francisco Bay Area—Cultural Resources Inventory of Caltrans District 4 Rural Conventional Highways.” Submitted to Office of Cultural Resource Studies, California Department of Transportation District 4. Environmental Division. Milliken, Randall. 2010. “Contact-Period Native California Community Distribution Model: A Dynamic Digital Atlas and Wiki Encyclopedia.” Report on file. Rohnert Park, CA: Northwest Information Center, Sonoma State University. Milliken, Randall T., and James A. Bennyhoff. 1993. “Temporal Changes in Beads as Prehistoric California Grave Goods.” In There Grows a Green Tree: Essays in Honor

P lACINg Prehistoric Mortuary Practices in Broader Social Context 261 of D. A. Fredrickson, ed. Greg White, Pat Mikkelsen, William R. Hildebrandt, and Mark E. Basgall, 381–395. Center for Archaeological Research at Davis 11. Davis: University of California. Milliken, Randall, Richard T. Fitzgerald, Mark G. Hylkema, Randy Groza, Tom Origer, David G. Bieling, Alan Leventhal, Randy S. Wiberg, Andrew Gottsfield, Donna Gillette, et al. 2007. “Punctuated Culture Change in the San Francisco Bay Area.” In California Prehistory: Colonization, Culture, and Complexity, ed. Terry L. Jones and Kathryn Klar, 99–124. Walnut Creek, CA: Altamira Press. Muheisen, Mujahed. 1988. “Epipaleolithic Phases of Kharaneh IV.” In The Prehistory of Jordan, ed. Andrew N. Garrard and Hans G. Gebel, 353–367. BAR International Series 396. Oxford: British Archaeological Reports. Nadel, Dani. 1994. “Levantine Upper Palaeolithic-Early Epipalaeolithic Burial Customs: Ohalo II as a Case Study.” Paléorient 20(1):113–121. http://dx.doi.org/10 .3406/paleo.1994.4605. Pearson, Mike P. 2001. The Archaeology of Death and Burial. Gloucestershire, UK: Stutton Publishing. Perrot, Jean, and Daniel Ladiray. 1988. Les Hommes de Mallaha (Eynan) Israel, les Sepultures. Memoires et Travaux du Centre de Recherche Française de Jerusalem 7. Paris: Association Paléorient. Potter, James M., and Elizabeth M. Perry. 2011. “Mortuary Features and Identity Construction in an Early Village Community in the American Southwest.” American Antiquity 76(3):529–46. http://dx.doi.org/10.7183/0002-7316.76.3.529. Price, T. Douglas, and Ofer Bar-Yosef. 2010. “Traces of Inequality at the Origins of Agriculture in the Ancient Near East.” In Pathways to Power, ed. T. Douglas Price and Gary M. Feinman, 147–168. NY: Springer. http://dx.doi.org/10.1007/978-1 -4419-6300-0_6. Rakita, Gordon F. M., and Jane E. Buikstra. 2007. “Introduction.” In Interacting with the Dead: Perspectives on Mortuary Archaeology for the New Millennium, ed. Gordon F. M. Rakita, Jane E. Buikstra, Lane A. Beck, and Sloan R. Williams, 1–11. Gainesville: University Press of Florida. Rakita, Gordon F. M., Jane E. Buikstra, Lane A. Beck, and Sloan R. Williams. 2007. Interacting with the Dead: Perspectives on Mortuary Archaeology for the New Millennium. Gainesville: University Press of Florida. Richter, Tobias, Jay Stock, Lisa Maher, and C. Hebron. 2010. “An Early Epipalaeolithic Sitting Burial from the Azraq Oasis, Jordan.” Antiquity 84(324): 321–334. http://dx.doi.org/10.1017/S0003598X00066606. Robben, Antonius C. G. M. 2005. Death, Mourning, and Burial: A Cross-Cultural Reader. New York: Wiley-Blackwell.

262 B yrd & Rosenthal Rollefson, Gary O., Alan H. Simmons, and Zeidan Kafafi. 1992. “Neolithic Cultures at ‘Ain Ghazal, Jordan.” Journal of Field Archaeology 19(4):443–470. http://dx.doi .org/10.2307/530427. Rolston, Scott L. 1982. “Two Prehistoric Burials from Qasr Kharaneh.” Annual of the Department of Antiquities of Jordan 16:221–229. Rosenthal, Jeffrey S. 2011. “The Function of Shell Bead Exchange in Central California.” In Perspectives on Prehistoric Exchange in California and the Great Basin, ed. Richard E. Hughes, 83–113. Salt Lake City: University of Utah Press. Rosenthal, Jeffrey S., and Brian F. Byrd. 2006. “Archaeological Survey Report for the I–580 Eastbound High Occupancy Vehicle Lane Project, East of Greenville Road to Hacienda Drive, Livermore Valley, Alameda County, California.” 4-ALA-580, K.P. R12.6/30.7 (P.M. R7.8/19.1), EA 04258-90810. Report on file. Rohnert Park, CA: Northwest Information Center, Sonoma State University. Rosenthal, Jeffrey S., and Richard T. Fitzgerald. 2012. “The Pleistocene-Holocene Transition in Western California.” In From the Pleistocene to the Holocene: Human Organization and Cultural Transformations in Prehistoric North America, ed. C. Britt Bousman and Bradley J. Vierra, 67–103. College Station: Texas A&M Press. Rosenthal, Jeffrey S., and Jack Meyer. 2004. Landscape Evolution and the Archaeological Record: A Geoarchaeological Study of the Southern Santa Clara Valley and Surrounding Region. Center for Archaeological Research at Davis Publication 14. Davis: University of California. Saxe, Arthur A. 1970. “Social Dimension of Mortuary Practices.” PhD dissertation, University of Michigan, Ann Arbor. Schwitalla, Al W., Terry L. Jones, Marin A. Pilloud, Brian F. Codding, and Randy S. Wiberg. 2014. “Violence among Foragers: The Bioarchaeological Record from Central California.” Journal of Anthropological Archaeology 33:66–83. http://dx.doi .org/10.1016/j.jaa.2013.11.004. Shanks, Michael, and Christopher Tilley. 1987. Re-constructing Archaeology: Theory and Practice. New York: Cambridge University Press. Stock, Jay, S. K. Pfeiffer, M. Chazan, and J. Janetski. 2005. “F–81 Skeleton from Wadi Mataha, Jordan, and Its Bearing on Human Variability in the Epipalaeolithic of the Levant.” American Journal of Physical Anthropology 128(2):453–465. http://dx.doi .org/10.1002/ajpa.20163. Tannam, Melody E., Michael R. Fong, Steven J. Rossa, Colin I. Busby, James C. Bard, Robert M. Harmon, David G. Brittin, Deborah M. Di Pasqua, Donna M. Garaventa, Angela M. Banet, et al. 1993. “Archaeological Emergency Data Recovery/CEQA Documentation, CA-ALA-42, Pleasanton, Alameda County,

P lACINg Prehistoric Mortuary Practices in Broader Social Context 263 California.” Report on file, Northwest Information Center. Rohnert Park, CA: Sonoma State University. Tarlow, Sarah, and Liv Nillson Stutz. 2013. The Oxford Handbook of the Archaeology of Death and Burial. Oxford: Oxford University Press. Temple, Daniel H., S. Kusaka, and Paul W. Sciulli. 2011. “Patterns of Social Identity in Relation to Tooth Ablation among Prehistoric Jomon Foragers from the Yoshigo Site, Aichi Prefecture, .” International Journal of Osteoarchaeology 21(3):323–335. http://dx.doi.org/10.1002/oa.1146. Tuzin, Donald. 1980. The Voice of the Tambaran: Truth and Illusion in Ilahita Arapesh Religion. Berkeley: University of California Press. Valla, François R. 1995. “The First Settled Societies—Natufian (12,500–10,200 BP).” In The Archaeology of Society in the Holy Land, ed. Tom E. Levy, 170–187. London: Leicester University Press. Verhoeven, M. 2002. “Ritual and Ideology in the Pre-Pottery Neolithic B of the Levant and Southeast Anatolia.” Cambridge Archaeological Journal 12(2):233–258. http://dx.doi.org/10.1017/S0959774302000124. Whitaker, Adrian R., and Brian F. Byrd. 2014. “Social Circumscription, Territoriality, and the Late Holocene Intensification of Small-Bodied Shellfish along the California Coast.” Journal of Island & Coastal Archaeology 9(2):150–168. http://dx .doi.org/10.1080/15564894.2014.881934. Wiberg, Randy S. 1988. The Santa Rita Village Mortuary Complex (CA-ALA-413): Evidence and Implications of a Meganos Intrusion. Archives of California Prehistory, 18. Salinas, CA: Coyote Press. Wiberg, Randy S. 1996. “Archaeological Excavations and Burial Removal at Sites CA-ALA-483, CA-ALA-483 Extension, and CA-ALA-555, Pleasanton, Alameda County, California.” Report on file, Northwest Information Center. Rohnert Park, CA: Sonoma State University. Wiberg, Randy S. 1997. “Archaeological Investigations at Site CA-ALA-42, Alameda County, California.” Report on file, Northwest Information Center. Rohnert Park, CA: Sonoma State University. Wohlgemuth, Eric. 1996. “Resource Intensification in Prehistoric Central California: Evidence from Archaeobotanical Data.” Journal of California and Great Basin Anthropology 18:81–103. Wohlgemuth, Eric. 2002. “Late Prehistoric Plant Resource Intensification in the Eastern San Francisco Bay Area: Plant Remains from ALA-42 and ALA-55, Pleasanton, California.” In Hunter-Gatherer Archaeobotany: Perspectives from the Northern Temperate Zone, ed. Sarah L. R. Mason and Jon G. Hather, 28–43. London: Institute of Archaeology, University of London.

264 B yrd & Rosenthal Wright, Gary A. 1978. “Social Differentiation in the Early Natufian.” In Social Archaeology: Beyond Subsistence and Dating, ed. Charles L. Redman, Mary Jane Berman, Edward V. Curtis, William T. Langhorne Jr., Nina M. Versaggi, and Jeffrey C. Wanser, 201–223. New York: Academic Press.

P lACINg Prehistoric Mortuary Practices in Broader Social Context 265

12

I shall probably never completely escape from Flint from the Ancestors an early training that identified as ceremonial every queer or inexplicable object. (Judd Ritualized Use of Stone Tools 1954:262) in the Prehistoric Southwest

This has of course become an old archaeological joke, and a tiresome one. The most extended riff upon the John C. Whittaker and theme is Macaulay’s (1979) classic Motel of the Mysteries, Kathryn A. Kamp in which a future “Howard Carson” excavates a twenti- eth-century motel, interpreting all the mundane finds as ritual funerary objects. But even the cleverest lam- poons, such as Motel of the Mysteries, while amusing, are also faintly irritating. They annoy because in most cases archaeologists can in fact do better. A holistic approach to the full range of evidence makes the functions of even unusual artifacts less obscure than they often seem at first. Furthermore, much ritual, ceremonial, or religious behavior is patterned behavior, which over time will leave patterns in the material remains, poten- tially interpretable to the discerning archaeologist, or at least comprehensible enough to avoid the absurdities of Carson’s partner decked out in lavatory “jewelry.” Nevertheless, we want to argue that the threadbare old joke has some truth to it, that there are going to be patterns in the archaeological record that cannot be readily interpreted in terms of what our materialistic world view considers utilitarian “function,” but that many of these patterns can be interpreted in terms of human activities that attempt to influence the DOI: 10.5876/9781607324942.c012

267 nonmaterial world. This is hardly a new view (e.g., Flannery 1976: 329–333), although the so-called post-processual archaeologies have given new promi- nence to the archaeological study of symbolic behaviors. All artifacts have the potential to be invested with meaning and used symbolically for social or religious purposes. Archaeological concepts like Binford’s (1962) technomic, sociotechnic, and ideotechnic (or utilitarian, social, and symbolic) functions are arbitrary classifications that prove far too tidy in practice, especially in societies that may not make the strong distinc- tions between real and intangible, sacred and profane, material and spiritual that we do. What we are really interested in is identifying instances when an artifact is used symbolically, and often the best way to do so is to find it used outside its expected function. A classic example in Southwestern archaeology is Watson Smith’s discussion of “when is a Kiva?” (Smith 1952:154–165). Smith concluded that in much of the Southwest there are no certain markers for recognizing ceremonial architecture, only patterns of unusual form, atypical but recurring features, and association with other rooms. He felt that “the ascription of ceremonial function in a given instance must rest largely on an inferential basis and will be only a more or less convincing hypothesis.” The same is true with stone tools. For instance, what do we do with two large black bifaces, exceptionally well-made of unusual nonlocal stone, from Grasshopper Pueblo, which were broken in complex and unusual ways? In the first case (figure 12.1a), after the biface snapped, the short section of the piece was set on end and smashed, resulting in bipolar impact scars on the break surfaces. The other biface was also snapped, then struck on the break, thereby burinating one whole edge; the spall was found with it (figure 12.1b). These artifacts could represent the helpless rage of a knapper who has just spoiled a couple of hours of hard work, but in our experience the bifaces’ attributes are too elaborate in response to a common set-back—a broken biface usually elicits a simple “curse and toss” reaction. These enigmatic biface fragments embody a number of aspects that are commonly used in identifying ritually used artifacts. April Sievert (1994), also working with stone tools, has usefully articulated a set of principles that of course can also be applied to other artifact classes. She suggests that ritually important stone tools are likely to be characterized by (1) exaggerated size or proportions, sometimes beyond what is functionally acceptable; (2) fine and ornate crafting; (3) presence of unusual residues such as pigments; (4) inten- tional breakage or destruction; (5) eccentric, unusual forms; and (6) apparent use for display, lacking evidence of normal kinetic use. To this list we must add

268 Whittaker & Kamp Figure 12.1. Two intentionally damaged bifaces from Grasshopper Pueblo (AZ P:14:1 [ASM]): (a) From Room 13, about 65 mm long, and damaged by bipolar percussion; (b) From Room 183, 163 mm long, and burinated down one edge (photos by J. Whittaker).

unusual contextual patterns or associations, as seen for instance in two points found at the bottom of postholes at Lizard Man Village, a Sinagua site in northern Arizona (figure 12.2). We think it is clear that these traits exemplify the archaeological joke— anomalous forms and treatments of artifacts are expected to reflect “ceremo- nial” use. Can we do any better than this? As anthropologists who do archaeology, the authors like to think we sometimes can. Archaeologists should be inter- ested in the meanings of artifacts used symbolically—what, in the minds of the people who used them, were they supposed to do? What was the meaning of the behavior that resulted in a shattered biface or a stone tool in a post- hole? Much of this meaning, especially the details, will always be lost with the deaths of people, minds, and cultures. However, repeated patterns and associa- tions, interpreted through ethnographic analogies, can allow us to probe some

R iTUALIZed Use of Stone Tools in the Prehistoric Southwest 269 Figure 12.2.Unusual lithic items from Sinagua sites (d, g, and h are from Fortress Hills [NA6612]; all others are from Lizard Man Village [NA17957]): (a [an early type] and b) obsidian points recovered from postholes; (c) fragment of an old point, laid on its side and damaged by bipolar percussion, from kiva fill (R19/539); (d) patinated old point with parallel flaking, reflaked to make a small arrow point; (e) small triangular arrow point with heavily ground edges and worn faces; (f ) older point with heavy grinding (R19/233) from kiva fill; (g) impact-damaged old point with heavily ground edges and worn faces; and (h) obsidian spicule (photo by J. Whittaker). of the more general kinds of meaning to be expected in stone tools (see also Rollefson, chapter 10, this volume). The Southwest seems an ideal place to attempt this kind of study. There are libraries full of well-reported archaeological work, and many detailed ethnographic accounts of descendant cultures. However, as we encountered instances of abnormal stone tool use in our own sites and elsewhere, we came to realize that there has been little systematic interest either in the common but scattered archaeological occurrences of ritualized stone-tool use or in the role of stone tools among recent Southwestern people that might allow us to interpret them. The situation has improved somewhat since we first prepared this paper in 2008 (e.g., Holly 2010; Sedig 2014). There are a number of obvious reasons why neither the ethnography nor the archaeology has received the attention we think it deserves. First, anomalous stone tools occur at many sites all over the Southwest, but they are usually relatively rare, making patterns hard to see. Second, stone tools have always been somewhat neglected among ceramocentric Southwestern archaeologists (see also Sullivan, chapter 6, this volume). Similarly, ethnographic accounts of stone-tool use in the Southwest are common but scattered, and stone tools had largely been replaced in native technologies by the time ethnographers got to work. To make the case that ritualized stone-tool use is present and can be inter- preted, it is worth assembling some examples to show the kind of things that have been and are being found. Many of these examples are taken from our own encounters beginning as graduate students at the large fourteenth-­ century pueblo of Grasshopper in central Arizona, and more recently in some- what earlier and much smaller Sinagua culture sites around Flagstaff. Personal examples help to make the point that, while we do not have the range of experience in time or geography of some of our mentors, we can think of quite a few instances of apparent ritual use of stone tools, and suspect that our peers likewise are aware of a large body of disparate and uncollated data. No one seems to have pursued the question of unusual uses of stone tools very far yet, so what we would like to do is argue that there are patterns in the unusual finds in the Southwest, and plenty of ethnographic reasons why we should expect them to be meaningful.

Finds and Contexts Burials are a situation where the unusual context of normal artifacts pro- vides the argument for ritual behavior. Artifacts intentionally included in

R iTUALIZed Use of Stone Tools in the Prehistoric Southwest 271 burials are pretty universally assumed to be there for symbolic purposes. Even if, in the minds of the mourners, tools in a grave were intended to serve their normal function in the otherworld existence of the dead, the archaeologist feels justified in reading those tools as saying something about belief systems and the expected activities of the individual in the grave. Projectile points are occasionally found in prehistoric Southwestern burials. At Grasshopper and elsewhere in the Southwest (e.g., Chaco Canyon [ Judd 1954; Lekson 1997]), they are sometimes found in groups, presumably the remains of a quiver of arrows. They may be there as necessary equipment for the deceased, reflecting an individual’s role as hunter or warrior, or they may suggest more abstract meanings. At Grasshopper, the points usually occur in sets that were apparently made by one knapper (figure 12.3; Whittaker 1984, 1987a, 1987b). They are often unusually large and carefully flaked, sometimes too fragile to be practical, and rarely show signs of use. Only a few of the adult male burials, always those well-endowed with other goods, have point sets, suggesting that a quiver of arrows is not necessary funerary equipment for everyone. Whittaker (1984) argued that they should be interpreted not as a man’s normal hunting gear, but as objects included in the grave to symbolize something about the social status of the deceased; for example, as Reid and Whittlesey (1999) more specifically suggest, point sets, Conus shell tinklers, and Glycymeris shell bracelets all mark membership in different sodalities within the Grasshopper community. At Grasshopper, one male burial accom- panied by many goods had also 128 points, representing some 16 sets or dif- ferent knappers (Griffin 1967; Whittaker 1984, 1987a), and in the Sinagua area, the “Magician Burial” (McGregor 1943) included 420 points, many of which were unusual forms. On the basis of his associated goods, this individual was interpreted by McGregor’s Hopi informants as a ritual leader whose multiple roles (O’Hara 2008) included warlike magic. Lithic artifacts of high value whose primary purpose seems to be display are less obvious in the Southwest than in some other areas. Maya eccentrics are the model: unusual form, extravagant workmanship, large and visually striking, but often lacking any possible utilitarian cutting function, they are generally found intentionally deposited in burials and foundations and sometimes bear- ing unusual residues of paint or even clothing (Guderjan 1998; Iannone 1992; Meadows 2001). In the Southwest, large bifacial knives are the closest analog, although many of them were, or could have been, useful tools. Bifaces larger than 10 cm long are found at many sites, but always as rare artifacts. They are often of exotic material such as obsidian. The specimens in figure 12.4, from central Arizona, are made of a distinctive chert from Wyoming (Whittaker

272 Whittaker & Kamp Figure 12.3. Point sets from two well-endowed male burials at Grasshopper Pueblo (AZ P:14:1 [ASM]) (photo by J. Whittaker). Figure 12.4. Large bifaces of Wyoming chert from central Arizona sites: Q-Ranch, Safford area, and Gipe site (Arizona State Museum; see Whittaker et al. 1988; photo by J. Whittaker).

et al.1988). Making a large biface requires a large piece of high-quality mate- rial, a rare commodity in many areas. Large, well-made bifaces also reflect an amount of work and skill that is beyond the norm for Southwestern stone tools. Not uncommonly, they are found in contexts like caches, niches, and burials that also signal a special function. Typical examples occur in published reports from Pecos (Kidder 1932) and Pueblo Bonito ( Judd 1954; Lekson 1997). Grasshopper again provides further examples (Whittaker 1984; Whittaker and Kaldahl 2001). One room of the 100 or so excavated in a pueblo of around 500 rooms contained an extensive midden deposit largely composed of lithic debitage. Much of it came from the manufacture of large bifaces, and included many that were broken in manufacture (figure 12.5). Similar bifaces did not

274 Whittaker & Kamp Figure 12.5. Broken bifaces from Room 246, Grasshopper Pueblo (AZ P:14:1 [ASM]; photo by J. Whittaker). occur elsewhere in the site. Arrow points were also being made in this room, and other male crafts were evidenced. The room had high concentrations of unusual artifacts and bird remains as well, and we interpret it as a ritually oriented structure analogous to Hopi clan houses. The bifaces could have been intended as tools or trade goods, but their limited distribution argues other- wise. As they are confined to a single context with other apparently ceremo- nial aspects, it seems likely that the manufacture of such bifaces was ritually meaningful, and the disposal of waste from them was also bound by unusual rules, which Walker (1995) and others would argue is a common pattern. Similarly, points and other tools are sometimes found in caches or other intentional deposits. Some of these may simply be stored items, but an arrow point buried at the bottom of a posthole, like those from Lizard Man Village (see figure 12.2), suggests an offering of some sort. It is unclear how com- mon this phenomenon is—it is difficult to search old reports for such small details, even when they were recorded. Foundation offerings of all kinds are common in many cultures, including all over Mesoamerica. In the Southwest, Judd (1954:156) reports offerings of “ceremonial sticks,” shell and bone beads, turquoise chips, and at least one arrow point in Pueblo Bonito walls, as well as many small deposits (none of which seem to have included flaked-stone artifacts), on the tops of kiva pilasters. In the projectile-point assemblages of most sites where we have worked in the Southwest, there are points that were already old when the site was occu- pied. In some cases, this reflects an earlier use of the site area, but this is not always the case. For instance, Lizard Man Village, a small Sinagua site near Flagstaff (Kamp and Whittaker 1999; Kamp 1998), was occupied between ad 1060 and 1260. The assemblage of 261 projectile points, which was dominated by small triangular arrow-point forms (n = 157), also included 55 large points. Some of these may have been knives or other tools, but most are recognizable earlier point forms, at a site with no other evidence of such earlier occupa- tion. This is true also of the three other sites in the area that we have exca- vated. Most of the large points are early forms, which appear to have been collected by later Sinagua people. These points even include a couple of late Paleo-Indian types (figure 12.6). At Lizard Man Village, less than one-tenth of the large points were whole, compared to almost half the small triangular points. Half the large points have identifiable impact damage, twice the rate of the small points. This suggests to us that early points were usually damaged and lost in the field, to be retrieved by later people, while more of the late arrow points were lost on site, damaged and abandoned during manufacture, or returned to the village after use.

276 Whittaker & Kamp Figure 12.6. Large old points from later Sinagua sites; photo by J. Whittaker). Some of these older collected points may have been simply recycled as tools, but they are most likely to show signs of unusual use as well. At Lizard Man Village, the points in postholes mentioned before included one old large point, as well as one later small point. Another large old point was included with the burial of a young woman. This is a pattern also seen atG rasshopper, where in burial contexts large old points are almost exclusively with females (4 to 10 instances), while sets of small triangular arrow points are exclusively male (n = 15) as are most single arrow-point associations (Whittaker 1984:287–291). In Ventana Cave, for another example, Haury (1950:290) reports a burial dat- ing after ad 1000 with associated arrow points, but also with a necklace of a shell, wooden nose plug, and two San Pedro type dart points, a type of point expected to date a thousand years earlier. At Lizard Man and elsewhere, we find a few points, usually older ones, whose edges have been heavily dulled—far beyond use-wear—and these often have abraded or polished surfaces as well (see figure 12.3). This suggests that they have been modified and carried in contact with other items, perhaps as charms or components of medicine kits. Again, this pattern is found elsewhere: of 14 groups of odd minerals and other objects found with burials at Pecos, seven of them included points, mostly old and damaged (Kidder 1932:106). At Rudd Creek Pueblo, a Tularosa-phase site in Eastern Arizona, a sealed floor pit in a habitation room contained three preceramic dart points, a small biface, and a red stone bead (Clark et al. 2006:416). A final example, again from the Sinagua region, is a small site we helped to record after it was found by a Coconino National Forest survey project. It consisted entirely of a surface scatter of projectile points and fragments. Many different types and ages of points were represented, and almost no other artifacts, nor was there a habitation site nearby. The most plausible explanation remains that the scatter represents some kind of shrine or offering.

E thnographic Parallels The archaeological examples above serve to show some of the patterns that we suspect are even more common than is apparent from the published sources. The suggested interpretations of these finds come of course from ethnographic analogies. Here too, the Southwestern literature has not been thoroughly explored, but we can readily present a few examples to show that the ritual use of stone tools, both old and new, is widespread and diverse, and supports the archaeological picture.

278 Whittaker & Kamp However, we are repeatedly reminded that careless use of ethnographic analogy chains us to particular models of the past, and dismisses the indi- vidual histories of past cultures. Even when there are descendant communi- ties that claim continuity with ancestral cultures, and archaeological evidence of long strains of cultural similarity, both of which are true in the Southwest (Dozier 1970), we cannot assume that the mutable meanings of symbols have remained the same. Today’s Hopi are not Sinagua, any more than the authors are Anglo-Saxons of Norman Conquest times. Today’s Christian sects show visible symbolic continuities with King Harold’s Christianity, but result from 1,000 years of schism, war, debate, proselytizing, inspiration, revival, and sup- pression. Likewise, current archaeological views suggest that the parts of mod- ern Pueblo religion most visible to outsiders, the public aspects of katsina cults, reflect practices and beliefs that most likely spread through the Southwest in the fourteenth century (Adams 1991; Adams and Lamotta 2006; Hays-Gilpin 2006; Schaafsma and Schaafsma 1974), perhaps incorporating elements from Mesoamerica and elsewhere. Pueblo religion was further modified by 500 years of contact with Christianity and the religious persecutions that came with it. Nevertheless, the existing religious forms seem to have incorporated at least parts of ancient thematic complexes that have been characterized as ideolo- gies of warfare (Schaafsma 2000) and “the Flower World” (Hays-Gilpin 2006; Hill 1992; Hill and Hays-Gilpin 1999), a spiritual landscape of beauty. The former, as we shall see, seems to connect better to stone tool symbology than the latter, although both share underlying concerns with water and fertility. Ethnographic accounts of Southwestern stone-tool use are spotty partly because by the time most of them had been written stone tools had largely been supplanted by metal ones. Nevertheless, stone projectile points and axes can hardly have been the mystery to historic Southwestern tribes that they were to most Europeans during the period of early contact. After all, until the nineteenth century, science and folklore in Europe mostly considered stone tools to be the natural products of lightning and petrifaction or the magical work of elves or other supernaturals (Blinkenberg 1987). It is striking that only a short period of obsolescence is enough to create an ignorance of a simple technology, to the point where it requires some sort of fanciful explanation. The folk idea that stone tools are made by dripping cold water on heated rocks can still be found in the United States; Neolithic points and axes were explained as elf stones and thunderbolts in Europe even while masons and gunflint knappers continued to flake stone, and mythical beings or ancestors were soon credited with making arrowheads by Southwestern tribes only a generation or two removed from knapping. Rather than suggest that this is

R iTUALIZed Use of Stone Tools in the Prehistoric Southwest 279 simply ignorance and superstition, we would like to argue that stone tools are of wide symbolic importance in the Southwest (and elsewhere) because they provide easy symbolic links to a number of common cultural concerns. After briefly demonstrating this, we will argue that many of the themes can be connected by a pervading perception that stone tools are linked to the land and to ancestors. A similar situation has been well documented in Australia (e.g., Akerman et al. 2002; Flood 1983:188; Tindale 1985), and parallels in other cultures around the world suggest that this is a common symbology (see Holdaway et al., chapter 5, this volume). Lightning is one of the symbolic connections most commonly made with stone tools, not just in the Southwest, but almost all over the world. It is easy to suggest, as did Cushing (1883b), that the motion and force of lightning create the mental connection between thunderbolts and arrows. Among the late-nineteenth-century Zuni, Cushing (and Parsons 1939 similarly) reports a preference for fetishes collected as naturally shaped stones or ancient effi- gies; projectile points collected from ruins were commonly added (figure 12.7). “Although fashioned by man, it [an arrowhead] is regarded as originally the gift or “flesh” of lightning and rendered more effective by connections with the dread element” (Cushing 1883b:9). Many representations of lightning in the Pueblo world end in arrow-shaped points because each flash is believed to be tipped with a stone knife (Roediger 1941), and a belief in lightning as the source of stone points, or of their power, is common (Parsons 1939:106, 333). In Europe, stone tools were often regarded as protective charms against lightning, apparently on the theory that as they were the product of lightning, and lightning never strikes twice in the same place, a stone axe under the eaves protected a house or the like (Blinkenberg 1987). Southwestern ethnog- raphy also suggests anti-electrical effects of stone tools (Hill 1982). However, in Navajo ritual, the power of lightning as represented by stone points may also be harnessed to protect against other dangers (Frisbie 1987:61–66). Projectile points were weapons, so it is not surprising to find symbolic uses connected to hunting as in many of the Zuni fetishes, or to protection from enemies during travel, also reported by Cushing (1883b) in connection with the points attached to fetishes. Stone points or knives appear in myth as instruments of both benevolent and malevolent power. The Soyok ogres, man-eating monsters of Hopi rit- ual, carry knives and saws, which in precontact times would probably have been stone knives and in this context represent their threat to butcher and eat human children (Fewkes 1897:288; Kealinohonoka 1980). In one story involv- ing a Soyok ogre (Malotki and Gary 2001:265), the vanquished monster is cut

280 Whittaker & Kamp Figure 12.7. Zuni fetishes with attached stone projectile points (from Cushing 1883b). open with its own knife to examine and destroy its evil heart. Darling (1998) suggests that stone knives are part of anti-witchcraft ceremonial and perhaps executions. Bunzel (1932b:873) noted in her study of Zuni katsinas, that some katsina accouterments included hafted stone knives. Stevenson (1904:173–174) described a ceremonial bundle that contained two obsidian knives, one of which was 10 cm in length. The hero twins of Zuni myth are armed with a great stone knife, which serves as a weapon against the monsters of the world and opens the way for humans to pass from one world to the next (Cushing 1883b). Cushing in 1879 documented a ceremonial dance at Zuni where participants carried “huge, leaf-shaped, blood-stained knives of stone” (Cushing 1883a:47). The knives

R iTUALIZed Use of Stone Tools in the Prehistoric Southwest 281 and points attached to Zuni fetishes represent supernatural weaponry and its power to protect the owner from enemies and provide success in hunting (Cushing 1883b:39, 40). Stevenson (1904:41) and Bunzel (1932a:528) both described the Zuni Stone Knife Society, whose principal “deity” is a figure with wings like knives. A parallel figure at Acoma is Flint Bird, who wears a coat of stone knives (Thompson 1999; White 1932:172). These creatures are sometimes monsters, sometimes subdued by the power of heroes, but always powerful and danger- ous, the spiritual force of nature in earth and sky.

Pos wer of the Past There are many close physical parallels between prehistoric and ethno- graphic practices that probably imply similar symbolic content. Arrows, for instance are commonly used as offerings placed at shrines and other sacred places (Haury 1945; Hibben 1938; Parsons 1918), or are included in ritual cos- tume (Hibben 1975; Roediger 1941). Arrow points are included in ritual para- phernalia such as medicine pouches (Frisbie 1987), attached to ritual costumes, or used as components of fetishes and other medicine objects (Cushing 1883b). However, it is the meaningful content that makes these interesting. In eth- nographic Southwestern society we can discern several common symbolic associations for stone tools. These are usually projectile points but include also knives and sometimes other tools. While women may use stone points and knives (and make them for that matter), stone tools are consistently associated with maleness around the world. Elaborate stone tools like projectile points and knives are often used for activities generally performed by men, particu- larly warfare and hunting, which are thus often referenced by stone tools. The uses and properties of stone tools can also symbolize death and destruction, either magical or real, or sometimes protection from supernatural dangers. Inimical supernatural beings may be feathered or clothed with flint, heroes destroy monsters with arrows and knives, sacrifice or ritual execution may involve them, and some otherworld beings are averted by them. These patterns likewise are not confined to the Southwest. Stone points are the products of thunder and lightning, and protect against it. Worldwide this is probably the most common association of stone tools in societies that no longer use them. Even where stone tools are no longer current, and are sometimes given fan- tastical explanations, it is not uncommon to recognize them as belonging to the past, the ancestors. In the prehistoric Southwest, where people saw and used stone tools every day, we have no way of knowing exactly how older

282 Whittaker & Kamp projectile points were explained. Were they seen as antiques from slightly older generations, objects from distant and mythical pasts, or the magical products of supernatural beings? As all these apply in ethnographic societ- ies, we can expect them in the prehistoric Southwest as well. Since old stone tools were collected and used for non-cutting purposes, we should suspect that some ritual significance was attached. The spiritual power of ancestors and of places is called upon by such objects, and forms a common thread in many of the ethnographic symbols. As time proceeds, as each generation is supplanted by the next, as settle- ments move from old locales and establish new ones, the landscape is enriched. Ruined settlements, artifact scatters, and other landscape features bear mute witness to the past and imbue places with both power and meaning. Places are associated not only with history and homeland, but also with meaningful sto- ries and oral history (Basso 1996; Naranjo 2008), with the spirits of the ances- tors, and with sacred beliefs and traditions (Enote and McLerran 2011; Snead 2008). The relationship between place, memory, meaning, and power has been well-documented cross-culturally (Bradley 2000; Deacon 1988; Moore 2005; Myers 2002; Schmidt 2006; Van Dyke and Alcock 2003) and the American Southwest is no exception. For Hopi, Zuni, Apache, and Tohono O’odham alike, attachment to the past and to the archaeological remains that symbolize the past is strong (Colwell-Chanthaphonh 2003). The landscapes of archaeological sites create imagined communities that simultaneously include the living inhabitants of the present and the residents of the past, both mythic and real. These sites of memory may have meaning to small social units such as families or to larger ones, even surpassing entire nations. Nevertheless, whether the locus is on the scale of a farmstead or the Statue of Liberty, the collective meanings of place have the power to act on a wealth of further symbols that are used to forge or solidify identity, reify cultural concepts, and provide spiritual union with the past (see also Roth, chapter 3, this volume). Individuals or groups may seek to partake of the power of place not only with stories and visits, but also by acquiring objects from sites of specific meaning, what Bradley (2000) calls “pieces of places.” The value of an arti- fact, sometimes the raw material itself, may be partly in its association with a meaningful place, and this kind of value may be enhanced or protected by dif- ficulty of access to the place (Bradley 2000; Edmonds 1995; Spielmann 2002). From the modern analyst’s point of view, this concept is difficult to apply to reused ancient artifacts. The location of the original raw material or even a cultural source area may be traceable, but we can usually say little about how

R iTUALIZed Use of Stone Tools in the Prehistoric Southwest 283 an older artifact came into the hands that deposited it in the archaeological record. However, “distance” and “difficulty of access” in many cases may have been the distance of time, and the difficulty of obtaining such objects without the proper spiritual preparation or qualifications. The Aztecs looted Tula (Diehl 1983:27) and commemorated a real or assumed heritage of greatness by owning Toltec art. By so doing, they asserted their relationship to the Toltecs and enhanced both their own prestige and that of their presumed ancestors. Today, tourists and collectors replay a similar scenario when they collect natural or manufactured objects, symbolizing among other things a connection with both time and place. Owning a piece of a place invests the owner with a power that derives from it. Artifacts can act as a proxy for place. It is this phenomenon that is seen with the anomalous stone tools that prehispanic Southwesterners sometimes collected and saved. Much as petro- gylphs are viewed by modern Zuni as “signs from the ancestors” (Colwell- Chanthaphonh 2003:9), stone tools and other artifacts have similar saliency. For the Apache, ancient places are imbued with power and consequently so are the artifacts found there. Both medicine men and others search for and collect from sites, although people are warned not to handle artifacts with their bare hands (Colwell-Chanthaphonh 2003:17). Points included in Navajo medicine bundles, like many other ritual components, are found objects. Significantly, according to some practitioners, they must be disinterred naturally, by erosion or animals (Frisbie 1987), a gift of the place or its spiritual occupants. Such objects may also be handed down for generations (Cushing 1883b:12). The effects of “curation” and heirlooming become important as objects are passed from owner to owner and acquire their own histories (Spielmann 2002), which enhances their value. However, these relatively recent histories may be mythologized, even working to obscure the specific origins of objects, and increase their association with the ancestral past. Points found in ruins, like tools handed down for generations, are evidence of connection to the land and its eternal spirituality. This is not the past of archaeologists, who subdivide to find order and sequential history, but that of ritual, which collectivizes the evidence of the past and makes it timeless. Old stone tools speak of the con- nections between people and their ancestors, mediating between humans and the powers of nature. Even today, the modern collector and also the archae- ologist, whether we admit it or not, feels an emotional connection to the past in such attractive and individual artifacts. While we may not believe in the same magical powers that the ancients probably attributed to the stone tools of their even older ancestors, to possess them makes concrete a link of kinship, of spiritual power, or of scientific understanding.

284 Whittaker & Kamp References Adams, E. Charles. 1991. The Origin and Development of the Pueblo Katsina Cult. Tucson: University of Arizona Press. Adams, E. Charles, and Vincent Lamotta. 2006. “New Perspectives on an Ancient Religion: Katsina Ritual and the Archaeological Record.” In Religion in the Prehispanic Southwest, ed. Christine VanPool, Todd VanPool, and David A. Phillips Jr., 53–66. Lanham, MD: Altamira Press. Akerman, Kim, Richard Fullager, and Annelou van Gijn. 2002. “Weapons and Wunan: Production, Function, and Exchange of Kimberley Points.” Australian Aboriginal Studies 1:13–42. Basso, Keith H. 1996. Wisdom Sits in Places: Landscape and Language among the Western Apache. Albuquerque: University of New Mexico Press. Binford, Lewis R. 1962. “Archaeology as Anthropology.” American Antiquity 28(2):217–225. http://dx.doi.org/10.2307/278380. Blinkenberg, Christopher. (Original work published 1911) 1987. The Thunderweapon in Religion and Folklore: A Study in Comparative Archaeology. New Rochelle, NY: Caratzas Publishing. Bradley, Richard. 2000. An Archaeology of Natural Places. London: Routledge. Bunzel, R. L. 1932a. Introduction to Zuni Ceremonialism. 47th Annual Report of the Bureau of American Ethnology. Washington, DC: U.S. Government Printing Office. Bunzel, R. L. 1932b. Zuni Katsinas: An Analytical Study. 47th Annual Report of the Bureau of American Ethnology. Washington, DC: U.S. Government Printing Office. Clark, Tiffany C., Gregson Schachner, Suzanne L. Eckert, Todd L. Howell, and Deborah L. Huntley. 2006. “Rudd Creek Pueblo: A Late Tularosa Phase Village in East Central Arizona.” Kiva 71(4):397–428. http://dx.doi.org/10.1179/kiv.2006 .71.4.003. Colwell-Chanthaphonh, Chip. 2003. “Signs in Place: Native American Perspective of the Past in the San Pedro Valley of Southeastern Arizona.” Kiva 69:5–29. Cushing, Frank H. 1883a [1970]. My Adventures in Zuni. Palo Alto, CA: America West Publishing Company. Cushing, Frank H. 1883b. Zuni Fetiches. Second Annual Report of the Bureau of Ethnology for 1880–1881, 1–45. Washington, DC: US Government Printing Office. Darling, J. Andrew. 1998. “Mass Inhumations and the Execution of Witches in the American Southwest.” American Anthropologist 100(3):732–752. http://dx.doi.org /10.1525/aa.1998.100.3.732.

R iTUALIZed Use of Stone Tools in the Prehistoric Southwest 285 Deacon, Janette. 1988. “The Power of a Place in Understanding Southern San Rock Engravings.” World Archaeology 20(1):129–140. http://dx.doi.org/10.1080/00438243.1 988.9980061. Diehl, Richard A. 1983. Tula: the Toltec Capital of Ancient Mexico. London: Thames and Hudson. Dozier, Edward P. 1970. “Making Inferences from the Present to the Past.” In Reconstructing Prehistoric Pueblo Societies, ed. William A. Longacre, 202–213. Albuquerque: University of New Mexico Press. Edmonds, Mark. 1995. Stone Tools and Society: Working Stone in Neolithic and Bronze Age Britain. London: BT Batsford Ltd. Enote, Jim, and Jennifer McLerran, eds. 2011. A:shiwi A:wan Ulohnanne: The Zuni World. Flagstaff: Museum of Northern Arizona. Fewkes, Jesse Walter. 1897. The Group of Tusayan Ceremonials Called Katcinas. Bureau of American Ethnology Annual Report No. 15 (for 1893–94), 245–313. Washington, DC: Smithsonian Institution. Flannery, Kent V., ed. 1976. The Early Mesoamerican Village. New York: Academic Press. Flood, Josephine. 1983. Archaeology of the Dreamtime: The Story of Prehistoric Australia and Her People. Sydney: Collins. Frisbie, Charlotte J. 1987. Navajo Medicine Bundles or Jish: Acquisition, Transmission, and Disposition in the Past and Present. Albuquerque: University of New Mexico Press. Griffin, P. Bion. 1967. “A High Status Burial from Grasshopper Ruin, Arizona.” Kiva 33:37–53. Guderjan, Thomas H. 1998. “The Blue Creek Jade Cache: Early Classic Ritual in Northwestern .” In The Sowing and the Dawning: Termination, Dedication, and Transformation in the Archaeological and Ethnographic Record of Mesoamerica, ed. S. B. Mock, 100–111. Albuquerque: University of New Mexico Press. Haury, Emil W. 1945. The Excavation of Los Muertos and Neighboring Ruins in the Salt River Valley, Southern Arizona. Papers of the Peabody Museum of American Archaeology and Ethnology 24. Cambridge, MA: Harvard University. Haury, Emil W. 1950. The Stratigraphy and Archaeology of Ventana Cave, Arizona. Albuquerque: University of New Mexico Press. Hays-Gilpin, Kelley A. 2006. “Icons and Ethnicity: Hopi Painted Pottery and Murals.” In Religion in the Prehispanic Southwest, ed. Christine VanPool, Todd VanPool, and David A. Phillips Jr., 67–80. Lanham, MD: Altamira Press. Hibben, Frank C. 1938. “A Cache of Wooden Bows from the Mogollon Mountains.” American Antiquity 4(1):36–38. http://dx.doi.org/10.2307/275360.

286 Whittaker & Kamp Hibben, Frank C. 1975. Kiva Art of the Anasazi at Pottery Mound. Las Vegas: KC Publications. Hill, Jane. 1992. “The Flower World of Old Uto-Aztecan.” Journal of Anthropological Research 48:117–144. Hill, Jane, and Kelley A. Hays-Gilpin. 1999. “The Flower World in Material Culture: An Iconographic Complex in the Southwest and Mesoamerica.” Journal of Anthropological Research 55:1–37. Hill, Willard Williams. 1982. An Ethnography of Santa Clara Pueblo, New Mexico. Albuquerque: University of New Mexico Press. Holly, Lance E. 2010. “Beyond the Point: Arrow Shaft Technology of the Prehistoric Southwest.” Master’s thesis, University of Colorado, Boulder. Iannone, Gyles John. 1992. “Ancient Maya Eccentric Lithics: A Contextual Analysis.” Master’s thesis, Trent University, Peterborough, ON. Judd, Neil. 1954. The Material Culture of Pueblo Bonito. Smithsonian Miscellaneous Collections 124. Washington, DC: Smithsonian Institution. Kamp, Kathryn. 1998. Life in the Pueblo: Understanding the Past through Archaeology. Prospect Heights, IL: Waveland Press. Kamp, Kathryn, and John Whittaker. 1999. Surviving Adversity: The Sinagua of Lizard Man Village. Anthropological Papers No. 120. Salt Lake City: University of Utah. Kealinohonoka, J. 1980. “The Drama of the Hopi Ogres.” In Southwestern Indian Ritual Drama, ed. Charlotte Frisbie, 37–69. Albuquerque: University of New Mexico Press. Kidder, Alfred Vincent. 1932. The Artifacts of Pecos. New Haven: Yale University Press. Lekson, Stephen H. 1997. “Points, Knives, and Drills of Chaco Canyon.” In Ceramics, Lithics, and Ornaments of Chaco Canyon: Analyses of Artifacts from the Chaco Project, 1971–1978, Volume 2, ed. Francis J. Mathien, 531–609. Publications in Archaeology 18G, Chaco Canyon Studies. Santa Fe: National Park Service, U.S. Department of the Interior. Macaulay, David. 1979. Motel of the Mysteries. Boston: Houghton Mifflin Company. Malotki, Ekkehart, and Ken Gary. 2001. Hopi Stories of Witchcraft, Shamanism, and Magic. Lincoln: University of Nebraska Press. McGregor, John C. 1943. “Burial of an Early American Magician.” Proceedings of the American Philosophical Society 88:270–298. Meadows, Richard Keith. 2001. “Crafting K’awil: A Comparative Analysis of Maya Symbolic Flaked Stone Assemblages from Three Sites in Northern Belize.” PhD dissertation, University of Texas, Austin. Moore, Jerry D. 2005. Cultural Landscapes in the Ancient Andes: Archaeologies of Place. Gainesville: University Press of Florida.

R iTUALIZed Use of Stone Tools in the Prehistoric Southwest 287 Myers, Fred R. 2002. “Ways of Place-Making.” La Ricerca Folklorica 45: 101–119. http://dx.doi.org/10.2307/1480159. Naranjo, Tessie. 2008. “Life as Movement: A Tewa View of Community and Identity.” In The Social Construction of Communities: Agency, Structure, and Identity in the Prehispanic Southwest, ed. Mark D. Varien and James M. Potter, 251–262. Lanham, MD: Alta Mira Press. O’Hara, Michael. 2008. “The Magician of Ridge Ruin: An Interpretation of the Social, Political, and Ritual Roles Represented.” Paper presented 73rd Annual Meeting of the Society for American Archaeology, Vancouver. Parsons, Elsie Clews. 1918. “War God Shrines of Laguna and Zuni.” American Anthropologist 20(4):381–405. http://dx.doi.org/10.1525/aa.1918.20.4.02a00030. Parsons, Elsie Clews. 1939. Pueblo Indian Religion. Chicago: University of Chicago Press. Reid, Jefferson, and Stephanie Whittlesey. 1999. Grasshopper Pueblo: A Story of Archaeology and Ancient Life. Tucson: University of Arizona Press. Roediger, Virginia More. 1941. Ceremonial Costumes of the Pueblo Indians: Their Evolution, Fabrication, and Significance in the Prayer Drama. Berkeley: University of California Press. Schaafsma, Polly. 2000. Warrior, Shield, and Star: Imagery and Ideology of Pueblo Warfare. Santa Fe: Western Edge Press. Schaafsma, Polly, and Curtis Schaafsma. 1974. “Evidence for the Origins of the Pueblo Katchina Cult as Suggested by Southwestern Rock Art.” American Antiquity 39(4):535–545. http://dx.doi.org/10.2307/278903. Schmidt, Peter R. 2006. Historical Archaeology in Africa: Representation, Social Memory, and Oral Traditions. Lanham, MD: Alta Mira Press. Sedig, Jakob W. 2014. “An Analysis of Non-utilitarian Stone Point Function in the US Southwest.” Journal of Anthropological Archaeology 34:120–132. http://dx.doi.org /10.1016/j.jaa.2014.03.003. Sievert, April Kay. 1994. “The Detection of Ritual Tool Use through Functional Analysis: Comparative Examples from the Spiro and Angel Sites.” Lithic Technology 19:146–156. Smith, Watson. 1952. Excavations in Big Hawk Valley, Wupatki National Monument, Arizona. Bulletin 24. Flagstaff: Museum of Northern Arizona. Snead, James E. 2008. Ancestral Landscapes of the Pueblo World. Tucson: University of Arizona Press. Spielmann, Katherine. 2002. “Feasting, Craft Specialization, and the Ritual Mode of Production in Small-Scale Societies.” American Anthropologist 104(1):195–207. http://dx.doi.org/10.1525/aa.2002.104.1.195.

288 Whittaker & Kamp Stevenson, Matilda C. 1904. The Zuni Indians: Their Mythology, Esoteric Fraternities and Ceremonies. 23rd Annual Report of the Bureau of American Ethnology. Washington, DC: US Government Printing Office. Thompson, Marc. 1999. “Knife-Wing: A Prominent Mesoamerican, Mimbres, and Pueblo Icon.” In Sixty Years of Mogollon Archaeology: Papers from the Ninth Mogollon Conference, Silver City, New Mexico, 1996, ed. Stephanie M. Whittlesey, 145–150. Tucson: Statistical Research Institute. Tindale, Norman B. 1985. “Australian Aboriginal Techniques of Pressure-Flaking Stone Implements: Some Personal Observations.” In Stone Tool Analysis: Essays in Honor of Don E. Crabtree, ed. M. G. Plew, J. C. Woods, and M. G. Pavesic, 1–33. Albuquerque: University of New Mexico Press. Van Dyke, Ruth, and Susan E. Alcock, eds. 2003. Archaeologies of Memory. Malden, MA: Blackwell Publishing. http://dx.doi.org/10.1002/9780470774304. Walker, William H. 1995. “Ceremonial Trash?” In Expanding Archaeology, ed. James M. Skibo, William H. Walker, and Axel E. Nielsen, 67–79. Salt Lake City: University of Utah Press. White, L. A. 1932. The Acoma Indians. 47th Annual Report of the Bureau of American Ethnology. Washington, DC: U.S. Government Printing Office. Whittaker, John C. 1984. “Arrowheads and Artisans: Stone Tool Manufacture and Individual Variation at Grasshopper Pueblo.” PhD dissertation, University of Arizona, Tucson. Whittaker, John C. 1987a. “Individual Variation as an Approach to Economic Organization: Projectile Points at Grasshopper Pueblo, Arizona.” Journal of Field Archaeology 14:465–480. Whittaker, John C. 1987b. “Making Arrowpoints in a Prehistoric Pueblo.” Lithic Technology 16:1–12. Whittaker, John C., Alan Ferg, and John Speth. 1988. “Arizona Bifaces of Wyoming Chert.” Kiva 53:321–334. Whittaker, John C., and Eric Kaldahl. 2001. “Where the Waste Went: A Knapper’s Dump at Grasshopper Pueblo, Arizona.” In Lithic Debitage: Context, Form, Meaning, ed. William Andrefsky, 32–60. Salt Lake City: University of Utah Press.

R iTUALIZed Use of Stone Tools in the Prehistoric Southwest 289

13

The term “mental template” has been used frequently by Form, Function, and lithic analysts in recent years, especially in discussions Mental Templates in of Lower Paleolithic biface and large-cutting-tool vari- Paleolithic Archaeology ability, of Middle Paleolithic stone tool typology, and of differences between Middle and Upper Paleolithic industries in Europe (e.g., Ambrose 1998; Ashton and Philip G. Chase White 2003; Barton 1990; Bisson 2001; Gowlett 1984, 1996, 2006; Marks et al. 2001; McNabb et al. 2004; McPherron 2000; Mellars 1996b; Monnier 2006, 2007; Nowell et al. 2003; White and Dibble 1986). The concept is linked to a number of issues that are of special interest to Paleolithic archaeologists—the origins and evolution of modern human intelligence, style, symbolism, lan- guage, and cultural norms or emic categories. The fol- lowing analysis of the term is made from this perspective. Arguments that employ the concept of mental tem- plates to make inferences about fundamentally impor- tant issues, such as intelligence or symbolism, usually take the following form:

• Standardization of attributes A1 . . . An of a set of lithic artifacts implies that the artifacts were made using a mental template. • The existence of a mental template implies the exis- tence of phenomenon P (symbolism, emic catego- ries, etc.). Conversely,

• Non-standardization of attributes A1 . . . An of a set of lithic artifacts implies that the artifacts were not DOI: 10.5876/9781607324942.c013 made using a mental template.

291 • The absence of a mental template implies the absence of phenomenon P. Thus, “mental template” seems to be thought of as an all-or-nothing phenome- non, whose existence or absence can be recognized through the standardization of whichever attribute or attributes a particular archaeologist chooses to analyze. However, there is no agreed-upon definition of a mental template. As a result, whenever the term is used, there is a real possibility that it will lead to misunder- standing. There is some danger of confusion about definition itself, but there is an even greater danger of misinterpretation at a deeper level. Because different authors conceive of mental templates differently, they also tend to think of the links between mental templates and phenomena such as intelligence or sym- bolism differently. Thus when one archaeologist argues that Middle Paleolithic technology involved or did not involve the use of mental templates, different readers will interpret the implications of that argument differently, and neither the writer nor the readers may be aware that this is the case. I argue here that the term “mental template” is completely unnecessary, that it is possible, without it, to express whatever it is one wants to explain. In fact, not using the term makes one’s point much clearer and unambiguous. This being the case, the term should be dropped from discussions of Paleolithic lithic technology.

History of the Term As far as I can determine, James Deetz was the first to define “mental tem- plate” in relation to archaeological artifacts. His definition is a very long one, but it is worth repeating part of it here. The idea of the proper form of an object exists in the mind of the maker, and when this idea is expressed in tangible form in raw material, an artifact results. The idea is the mental template from which the craftsman makes the object. The form of an artifact is a close approximation of this template, and variations in a group of similar objects reflect variation in the ideas which produce them. What gives form to the idea or mental template held by the maker of an arti- fact? Certainly tradition, since learning a craft entails the transmission of these templates from generation to generation, and many aspects of them have been present for so long that people simply feel that this shape for an axe or that color for a basket is inherently right, However, factors other than the purely tra- ditional can affect the form of the mental template, and there are other factors which affect the form of the finished product which are completely unrelated to the template involved.

292 Chase For example, consider the mental template used in the production of a basket made by the Chumash Indians of southern California. . . . This basket can be described in terms of its various attributes, discrete features which in their com- bination give the basket its distinctive form. Each of these attributes contributes to the form of the basket, and each is present in the basket and was an attribute of the mental template that produced it for a reason. These reasons are not all the same; they could be a matter of technology, function, innovation, or tradition. (Deetz 1967:45–47) Deetz was writing about Holocene archaeology, which enabled him to assume modern intelligence, symbolism, language, and so forth on the part of those who made an artifact. Nevertheless, looking at his definition of mental template from the perspective of Paleolithic archaeology would necessitate the following points: • A mental template is an idea in the mind of the maker of an artifact. • A mental template applies to an artifact as a whole. • A mental template includes or attends to a subset of all the attributes of the artifact. (Other attributes are omitted from the template.) • While “tradition” may determine all or part of the mental template, other fac- tors or considerations also contribute to the template. • Factors other than the maker’s mental template will contribute to the final form and appearance of the artifact. Since the publication of Deetz’s book, Paleolithic archaeologists have used the term in various ways. All of them include the first point listed above, but few of them keep all four of the other aspects of his definition, and there are other important differences among them. For example, consider the following range of opinions (in alphabetical order):

1. Gowlett (19 8 6 :2 51): Such operations require not just a mental template of “form” but a procedural template directing processes through time (Gowlett 1982). They especially require the ability to “see” in the mind’s eye what is required for the next flake, and how to make that compatible with the desired end product.

2 . Marks et al. (2 0 0 1:2 6 ): Are mental templates, in fact, manifested in the archaeological record by increasing standardization? Deetz . . . defined a mental template as “the idea of the proper form of an object [which] exists in the mind of the maker.” If,

F orm, Function, and Mental Templates in Paleolithic Archaeology 293 in fact, standardization results from this idea, it takes two major forms: stan- dardization of process and standardization of product.

3. Mellars (19 8 9 a:36 5; see also Mellars 19 9 6 a, 19 9 6 b): The forms of these distinctively Upper Palaeolithic tools appear to show not only a higher degree of “standardization” than those characteristic of the earlier Middle Palaeolithic industries (see Dibble 1987, 1989; Isaac 1972) but also a more obvious degree of “imposed form” in the various stages of their production and shaping. In other words, the shapes of the tools not only are more sharply defined but also appear to reflect more clearly conceived “men- tal templates” underlying the production.

4 . Monnier (2 0 0 6 :59 ): The notion of mental template was best described by Deetz [emphasis Monnier’s]: “The idea of the proper form of an object exists in the mind of the maker, and when this idea is expressed in tangible form in raw material, an artifact results. The idea is the mental template from which the craftsman makes the object.” In other words, when a specific form occurs repeatedly in an assemblage, it is assumed that it represents a desired end-product manu- factured according to certain socially defined parameters. These parameters result from mental categories similar to those which represent words, and are symbolic in nature.

5. (Monnier 2 0 0 6 :7 7 ): In sum, it is hypothesized that whereas overall tool morphology is unimportant in the majority of tasks to which stone tools are put, certain tasks, particularly those involving perforating or hafting, do require a very specific stone tool morphol- ogy. In other words, a mental template may well be required for hafted or perforating tools.

6 . Nowell et al. (2 0 0 3:19 3 [footnote]): The term “mental template” is being used here in the conventional sense of the word—namely that it is a preconceived idea in the mind of the knapper of the exact type and shape of tool that he or she desires to knap. In other words, it is suggested by some that handaxe morphology is the direct result of a specific idea in the mind of the knapper of what that tool should look like.

7. White and Dibble (19 8 6 :4 7 ): There are five main components to artifact variation: material, function, men- tal template, technology, and skill. . . . If these four variables [material, func-

294 Chase tion, technology, and skill] are controlled, then any variation which remains should result from variations in mental templates.

8 . Wynn (2 0 0 4 :6 7 2 ): There is no compelling reason to conclude that the knappers of large cutting tools relied on mental templates if by “mental template” we mean a preexist- ing image of a final product.

Even from this limited selection of quotes, it is clear that the concept of mental template means different things depending upon the author. For example, Mellars (quote 3), Monnier (quote 5), and Nowell et al. (quote 6) define mental templates in terms of overall artifact shape. Monnier (quote 5) and Wynn (quote 8) state that the manufacture of some artifacts may not require mental templates. White and Dibble (quote 7) define mental tem- plates (at least operationally) as a residual. Even more important is the fact that different authors draw different conclu- sions about cognition or culture from the perceived presence, absence, or nature of mental templates. This implies that they use different theoretical frameworks to think about mental templates, frameworks that are more often alluded to than explicitly described. For example, McNabb et al. (2004) link the concept to social tradition and to socially imposed norms of acceptability; Mellars (1996a) to symbolism and/or esthetic appreciation; Monnier (2006) to symbolic emic types or to special functions such as drilling or hafting; and Nowell et al. (2003) to intelligence, the use of symbols, and the appearance of language. By the same token, the variables that different archaeologists have observed or measured in order to demonstrate or refute the use of mental templates vary con- siderably. Marks et al. (2001), comparing Middle and Upper Paleolithic burins, looked at burin types, blank selection, shape of retouched edge, burin position, length, width, thickness, length/width ratio, length/thickness ratio, and width/ thickness ratio. McNabb et al. (2004), studying Acheulean large cutting tools, analyzed tip shape, extent and pattern of flaking, symmetry, and extent of edge working. McPherron (2000), in assessing whether regularities in biface form were simply a byproduct of resharpening rather than of intended form, mea- sured length/width ratio and the ratio of length to elongation (see also Rollefson, chapter 10, this volume). Monnier (2006) looked at size (length, width, or thick- ness), shape (length/width ratio, width/thickness ratio), location of retouch, and symmetry. Nowell et al. (2003) looked at the two-dimensional outline shape and location of retouch on bifaces. White and Dibble (1986), studying ethnographic artifacts, looked at length, width, length/width ratio, and edge angle.

F orm, Function, and Mental Templates in Paleolithic Archaeology 295 By pointing out the different ways in which “mental templates” are con- ceived by different authors, I am in no way criticizing their work. However, even if, taken individually, their thinking is impeccable, it is nevertheless clear that a real possibility for miscommunication exists.

Minimalist Model of Knapping One way to clarify the ambiguity that has developed with regard to “mental templates” is to analyze the very minimum that must exist in a knapper’s mind. In order to make a flaked-stone artifact, or any other artifact, the maker must have the following: • A purpose or goal—that is, there must be a reason to make the artifact, whether it be to have a tool for skinning an animal or a flag to mark one’s ethnic identity. • A mental template—that is, the maker must have an idea of what attributes an artifact must have to fulfill its purpose. (I will, for the present, use the term “mental template” in this way.) If, for example, a tool is to be used for cutting, it must at a minimum have a sharp edge and be large enough to hold or haft. Note that other attributes will not be part of the mental template, unless the tool is to fulfill other purposes as well. • A “procedural template”—that is, the maker must have an idea of how to go about making an artifact that has the attributes included in the mental tem- plate. (I have borrowed the term from Gowlett [1984, 1986], although he did not use it in exactly this way.) All three of these—purpose, mental template, and procedural template— must be present in the making of even the simplest artifact. This fact is illus- trated by experiments with a bonobo (Pan paniscus) named Kanzi (Schick et al. 1999; Toth et al. 1993). Kanzi was first shown how sharp-edged flakes struck from a core could be used to cut a cord or membrane to gain access to food visible but otherwise inaccessible in a box. He was then presented with flakes of differing sharpness and quickly learned to select the most suitable one. Finally, he was given a cobble suitable for a hammerstone and a nodule of stone suitable for flaking. With some difficulty, he learned to remove flakes, and invented his own method of flaking by throwing the nodule on the con- crete floor or throwing one stone against the other. Kanzi certainly had a goal in mind when he attempted to make flakes. He wanted something that would give him access to the food. It is also very clear from the published descriptions of his behavior that Kanzi knew the kind of artifact he wanted when he either selected or

296 Chase produced a flake: an edge sharp enough to cut the cord or membrane. His mental template may also have included other attributes of the edge (Schick et al. 1999:830), and probably a minimum flake size sufficient for prehension. Note that Kanzi’s mental template almost certainly did not involve the out- line shape of the flake he wanted to make, its symmetry, the location of the cutting edge relative to the bulb of percussion, or any similar attribute. Other attributes of the flake—and presumably all attributes of the core after the flake was removed—were probably of no interest to him whatsoever. Where Kanzi had difficulty was with the procedural template. He under- stood that percussion was required, but not exactly how to apply it. Nevertheless, he did develop his own procedural template, even if it was less efficient than that of a human knapper. In short, even Kanzi knew why he was trying to make a flake, and what the flake would have to be like to attain that end. He very clearly had in mind both a purpose and a mental template.

G oals, Attributes, and Mental Templates A mental template exists because the flint-knapper wants an artifact that will fulfill a certain purpose or purposes. Such purposes may be functional, symbolic, cultural, or some combination of these. Kanzi’s purpose was entirely functional (in the practical sense). He wanted to cut his way into a box to obtain food. Therefore, the mental template he had of the flake he wanted to produce addressed only those attributes of the flake that related to achieving this functional purpose. By contrast, someone whose goal is to produce an artifact with symbolic meaning (one that will fulfill a symbolic purpose) will have a mental template that includes the relevant symbolic attributes. Although there may be a practi- cal function as well, the mental template will of necessity include ideas about attributes needed to achieve the artifact’s symbolic purpose. A company logo on the side of a vehicle is a good example of an artifact with an almost purely symbolic purpose. It serves to identify the ownership of the vehicle and to advertise the company. Mesoamerican “eccentrics” are an example of flaked-stone tools that were probably made to serve a largely symbolic purpose. The mental templates of both the painter of the logo and the maker of an eccentric would include only symbolic attributes (as well as technologically necessary attributes—hopefully the logo will not wash away the first time it rains). A military officer’s overcoat is a good example of an artifact intended to fulfill both practical and symbolic purposes. It must keep the wearer warm,

F orm, Function, and Mental Templates in Paleolithic Archaeology 297 but it must also indicate the officer’s military affiliation and rank. The uniform maker’s mental template would therefore be the product of functional and symbolic goals and would include functional and symbolic attributes. The point is that mental templates can include more than one kind of attri- bute because either (1) different artifacts have different purposes or (2) one artifact may have more than one purpose. It follows that mental templates may or may not be associated with symbolic goals and therefore may or may not include symbolic attributes. The absence of a symbolic goal or symbolic attributes does not imply the absence of a mental template, and the mere presence of a mental template does not automatically imply the presence of symbolic intent.

Standardization As mentioned above, “standardization” is a concept that has played a consid- erable role in discussions of Paleolithic industries, and has often been linked to mental templates (e.g., Klein 1995, 1999; Mellars 1989b, 1996b; Monnier 2006; Nowell 2000; Nowell et al. 2003). It is worth examining the link between standardization and mental templates. “Standardization” is roughly synonymous with “uniformity” in the sense that standardized artifacts are necessarily uniform. However, the term also implies that the uniformity is intended by the makers of the artifacts, although neither the fact nor the nature of that intent is always made explicit by those using the word. Recognizing standardization thus requires both (1) demonstrating uniformity and (2) determining the source of that uniformity. A number of authors have pointed out that uniformity may be produced by factors other than the intentions or mental templates of the makers. If raw material is available only in small nodules, then there is an upper limit to the size of any artifacts made, and this imposes a certain uniformity of size. It has been argued that the remarkable uniformity of length to width ratio in Acheulean bifaces is the result not of what the knappers wanted to make but of natural technological constraints (see especially Dibble 1989). Archaeological typology may impose a uniformity that was never a part of the makers’ emic concepts of their artifacts (e.g., Nowell et al. 2003; White et al. 1982). It is also possible that procedural templates may produce uniformity that is not part of either the goal or the mental template of the maker (see Olszewski, chapter 4, this volume). For example, Upper Paleolithic blade technology tended to produce blades that are remarkably uniform in size and shape. It does not necessarily follow that Upper Paleolithic knappers had as a goal the

298 Chase production of uniform blades. (This may indeed have been the case, but it cannot be assumed a priori.) McPherron (1999) has suggested that regular changes in the pointedness of bifaces are the result not of the makers’ mental templates but rather of the resharpening procedure. Given these caveats, how- ever, there are several additional factors involved in the relationship between mental templates and standardization.

Mental Templates Produce Standardized Artifacts Only If They Are Both Consistent and Precise When I say that mental templates must be consistent, I mean simply that artifact after artifact must be made to the specifications of mental templates that are similar. At least one attribute must be defined in the same way from template to template. This must be true of the templates of all the knappers responsible for the assemblage, industry, or collection for which standardiza- tion is being measured. Such consistency can be the product of a cultural norm, of the demands placed on an artifact by its intended purpose, or of other factors. When I say that mental templates must be precise, I mean that when an attribute of the artifact is defined, it is defined in a way that provides minimal leeway in how the artifact is made. For example, in modern industrial produc- tion, the size of the prongs of an electrical plug is very precisely defined. (This is due to a cultural but not symbolic convention. Prongs and sockets are made to the same specifications by an agreement that ensures their compatibility.) However, 60,000 years ago, when someone in France removed a flake from a core in order to dismember a reindeer carcass, it is entirely possible that he or she was concerned only that the size of the flake should fall between a minimum and maximum limit, large enough to hold and not so large as to be unwieldy. In this case, the mental template to which the flake was made was imprecise with regard to size. A large sample of flakes made to this exact same mental template would be variable in size—even though minimum and maxi- mum dimensions would have been included as attributes in mental templates that may have been consistent from one individual to another.

Only Those Attributes That Are Part of the Mental Template Will Be Standardized If an archaeologist finds variability in one or more attributes of a given set of artifacts, it does not follow that they were not made to a very consistent

F orm, Function, and Mental Templates in Paleolithic Archaeology 299 and precise mental template. It may be simply that the archaeologist has not looked at the right attributes. A class of tools that are extremely variable in terms of overall (outline) shape may have been made to very exacting specifi- cations with regard to edge angle or some other attribute of the cutting edge. It is difficult, in other words, to disprove standardization in toto, although it may be easy to do so with regard to any particular attribute.

Standardization Reflects the Goals That Underlie the Mental Template These goals need not be symbolic or cultural in nature. As Monnier (2006) has pointed out, purely functional considerations, such as hafting or perforat- ing, may impose specific forms on an artifact. In this case, (1) the mental tem- plates of the makers will include the relevant morphological attributes and (2) these attributes are likely to be defined precisely. The result is standardization that is entirely functional or technological in origin. On the other hand, the fact that the function of an artifact is symbolic or cultural rather than practical does not guarantee that its mental template must be precise. The variety of forms that the letters of the alphabet may take without being rendered unrecognizable is a case in point (figure 13.1). Thus, neither the presence nor the absence of standardization in an artifact assemblage demonstrates a priori the presence or absence of a symbolic role for those artifacts.

Mental Templates and the Frison Effect Although mental templates are inevitably involved in the manufacture of artifacts, it is not necessarily the case that the artifacts recovered from an archaeological excavation reflect those templates. As Jelinek (1976:22) pointed out, “the tool kit ultimately abandoned at the site is the result of the modi- fication of an original set of tools and may be quite different in form from the original set. This modification occurs through the use of these tools in a succession of tasks related to the processing of raw materials.” Therefore, “an understanding of most collections as largely composed of materials that were no longer wanted is essential to their interpretation” (Jelinek 1976:27). This concept has been taken up by Rolland, Dibble, McPherron, and others (Dibble 1984, 1987, 1988, 1995; Dibble and Rolland 1992; McPherron 1995, 1999, 2000; Rolland 1981; Rolland and Dibble 1990). For instance, Dibble (1987:110) observed that “in the study of the Bisitun material . . . it appeared that

300 Chase Figure 13.1. Letters of the alphabet may take many forms.

individual flake blanks were repeatedly retouched until a certain minimum width was obtained (presumably relating to either hafting or grasping require- ments), at which point the piece was discarded.” Similarly, Barton (1990:70) noted that “virtually all lithics found at sites entered the archeological context because they were no longer of value to the makers and users.” In other words, artifacts may be in the archaeological record precisely because they do not fit the mental template of their user. A tool may be kept and used until it is no longer perceived to be suitable for use, at which point it is discarded and enters the archaeological record. Thus it cannot be assumed that all the artifacts in an archaeological assemblage reflect the mental tem- plates of their makers.

Is the Term “Mental Template” Useful? Inferring phenomena such as emic or linguistic categories, style, or sym- bolism from standardization depends on (1) the attributes the knappers pay attention to and (2) the attributes measured by the archaeologist. With this understanding, the term “mental template” is superfluous because every arti- fact was made using a mental template. The knapper had to have an idea of what attributes an artifact should have to fulfill its purpose. However, an archaeologist intending to interpret an attribute, such as outline shape, by means of a “mental template” is likely to confuse the issue, because others may not have the same definition in mind. It is much simpler and more direct, and detracts nothing from the argument, to simply name the attribute or attributes. For example, consider the following: • The scraper outlines are highly standardized; • The knappers had mental templates; • Therefore, there were cultural norms. This argument can be reduced, with no loss of meaning, as follows: • The scraper outlines are highly standardized; • Therefore, there were cultural norms.

F orm, Function, and Mental Templates in Paleolithic Archaeology 301 Similarly, consider this argument: • The length/width ratios of the tools are not standardized; • The knappers had no mental templates; • Therefore, there were no emic categories. This, too, can be reduced, as follows: • The length/width ratios of the tools are not standardized; • Therefore, there were no emic categories. In fact, it is much clearer in the shorter arguments, to see exactly what logi- cal steps must be defended theoretically or empirically in order to make the argument work. If the use of “mental templates” adds nothing to the logic of an argument and in fact increases the chance of misunderstanding, then the inevitable conclusion is that the term should be dropped.

Acknowledgment I thank Gilliane Monnier for reading and commenting on an earlier version of this essay.

References Ambrose, Stanley H. 1998. “Chronology of the Later Stone Age Food Production in East Africa.” Journal of Archaeological Science 25(4):377–392. http://dx.doi.org/10 .1006/jasc.1997.0277. Ashton, Nick, and Mark J. White. 2003. “Bifaces and Raw Material: Flexible Flaking in the British Early Paleolithic.” In Multiple Approaches to the Study of Bifacial Technologies, ed. M. Soressi and H. L. Dibble, 109–124. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Barton, C. Michael. 1990. “Beyond Style and Function: A View from the Middle Paleolithic.” American Anthropologist 92(1): 7–72. http://dx.doi.org/10.1525/aa.1990 .92.1.02a00040. Bisson, Michael S. 2001. “Interview with a Neanderthal: An Experimental Approach for Reconstructing Scraper Production Rules, and Their Implications for Imposed Form in Middle Paleolithic Tools.” Cambridge Archaeological Journal 11(02):165–184. http://dx.doi.org/10.1017/S0959774301000099. Deetz, James. 1967. Invitation to Archaeology. American Museum Science Books. Garden City: Natural History Press.

302 Chase Dibble, Harold L. 1984. “Interpreting Typological Variation of Middle Paleolithic Scrapers: Function, Style, or Sequence of Reduction?” Journal of Field Archaeology 11(4):431–436. http://dx.doi.org/10.2307/529322. Dibble, Harold L. 1987. “The Interpretation of Middle Paleolithic Scraper Morphology.” American Antiquity 52(1):109–117. http://dx.doi.org/10.2307/281062. Dibble, Harold L. 1988. “Typological Aspects of Reduction and Intensity of Utilization of Lithic Resources in the French Mousterian.” In Upper Pleistocene Prehistory of Western Eurasia, University Museum Symposium Series, ed. H. L. Dibble and A. Montet-White, 181–197. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Dibble, Harold L. 1989. “The Implications of Stone Tool Types for the Presence of Language during the Middle Paleolithic.” In The Human Revolution: Behavioral and Biological Perspectives on the Origins of Modern Humans, ed. P. Mellars and C. B. Stringer, 415–432. Edinburgh: Edinburgh University Press. Dibble, Harold L. 1995. “Middle Paleolithic Scraper Reduction: Background, Clarification, and Review of Evidence to Date.” Journal of Archaeological Method and Theory 2(4):299–368. http://dx.doi.org/10.1007/BF02229003. Dibble, Harold L., and Nicholas Rolland. 1992. “On Assemblage Variability in the Middle Paleolithic of Western Europe: History, Perspectives and a New Synthesis.” In The Middle Paleolithic: Adaptation, Behavior and Variability, ed. H. L. Dibble and P. Mellars, 1–20. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Gowlett, John A. J. 1984. “Mental Abilities of Early Man: A Look at Some Hard Evidence.” In Hominid Evolution and Community Ecology: Prehistoric Human Adaptation in Biological Perspective, ed. R. A. Foley, 199–200. London: Academic Press. http://dx.doi.org/10.1111/j.1468-2273.1984.tb01387.x. Gowlett, John A. J. 1986. “Culture and Conceptualisation: The Oldowan-Acheulian Gradient.” In Stone Age Prehistory, ed. G. N. Bailey and P. Callow, 243–260. Cambridge: Cambridge University Press. Gowlett, John A. J. 1996. “Mental Abilities of Early Homo: Elements of Constraint and Choice in Rule Systems.” In Modelling the Early Human Mind, ed. P. Mellars and K. R. Gibson, 191–216. Cambridge: McDonald Institute Monographs, University of Cambridge. Gowlett, John A. J. 2006. “The Elements of Design Form in Acheulian Bifaces: Modes, Modalities, Rules and Language.” In Axe Age: Acheulian Tool-making from Quarry to Discard, ed. N. Goren-Inbar and G. Sharon, 203–222. London: Equinox.

F orm, Function, and Mental Templates in Paleolithic Archaeology 303 Jelinek, Arthur J. 1976. “Form, Function and Style in Lithic Analysis.” In Cultural Change and Continuity: Essays in Honor of James Bennett Griffin, ed. C. E. Cleland, 19–33. New York: Academic Press. Klein, Richard G. 1995. “Anatomy, Behavior, and Modern Human Origins.” Journal of World Prehistory 9(2):167–198. http://dx.doi.org/10.1007/BF02221838. Klein, Richard G. 1999. The Human Career: Human Biological and Cultural Origins. 2nd ed. Chicago: University of Chicago Press. Marks, Anthony E., Harold J. Hietala, and John K. Williams. 2001. “Tool Standardization in the Middle and Upper Palaeolithic: A Closer Look.” Cambridge Archaeological Journal 11(1):17–44. http://dx.doi.org/10.1017/S095977430 1000026. McNabb, John, Francesca Binyon, and Lee Hazelwood. 2004. “The Large Cutting Tools from the South African Acheulean and the Question of Social Traditions.” Current Anthropology 45(5):653–677. http://dx.doi.org/10.1086/423973. McPherron, Shannon P. 1995. “A Re-examination of the British Biface Data.” Lithics 16:47–63. McPherron, Shannon P. 1999. “Ovate and Pointed Handaxe Assemblages: Two Points Make a Line.” Préhistoire Européenne 14:9–32. McPherron, Shannon P. 2000. “Handaxes as a Measure of the Mental Capabilities of Early Hominids.” Journal of Archaeological Science 27(8):655–663. http://dx.doi.org /10.1006/jasc.1999.0467. Mellars, Paul. 1989a. “Major Issues in the Emergence of Modern Humans.” Current Anthropology 30(3):349–385. http://dx.doi.org/10.1086/203755. Mellars, Paul. 1989b. “Technological Changes at the Middle-Upper Palaeolithic Transition: Economic, Social and Cognitive Perspectives.” In The Human Revolution: Behavioral and Biological Perspectives on the Origins of Modern Humans, ed. P. Mellars and C. B. Stringer, 338–365. Edinburgh: Edinburgh University Press. Mellars, Paul. 1996a. The Neanderthal Legacy: An Archaeological Perspective from Western Europe. Princeton: Princeton University Press. Mellars, Paul. 1996b. “Symbolism, Language and the Neanderthal Mind.” In Modelling the Early Human Mind, ed. P. Mellars and K. R. Gibson, 15–32. McDonald Institute Monographs. Cambridge: University of Cambridge. Monnier, Gilliane F. 2006. “Testing Retouched Flake Tool Standardization dur- ing the Middle Paleolithic: Patterns and Implications.” In Transitions before the Transition: Evolution and Stability in the Middle Paleolithic and Middle Stone Age, ed. E. Hovers and S. L. Kuhn, 57–83. New York: Springer. http://dx.doi.org/10.1007/0 -387-24661-4_4.

304 Chase Monnier, Gilliane F. 2007. “Middle Palaeolithic Scraper Morphology, Flaking Mechanics, and Imposed Form: Revisiting Bisson’s ‘Interview with a Neanderthal.’ ” Cambridge Archaeological Journal 17(03):341–350. http://dx.doi .org/10.1017/S0959774307000406. Nowell, April. 2000. “The Archaeology of Mind: Standardization and Symmetry in Lithics and Their Implications for the Study of the Evolution of the Human Mind.” PhD dissertation, University of Pennsylvania, Philadelphia. Nowell, April, Kyoungju Park, Dimitris Metaxis, and Jinah Park. 2003. “Deformation Modeling: A Methodology for the Analysis of Handaxe Morphology and Variability.” In Multiple Approaches to the Study of Bifacial Technologies, ed. M. Soressi and H. L. Dibble, 193–209. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Rolland, Nicholas. 1981. “The Interpretation of Middle Paleolithic Variability.” Man 16(1):15–42. http://dx.doi.org/10.2307/2801973. Rolland, Nicholas, and Harold L. Dibble. 1990. “A New Synthesis of Middle Paleolithic Variability.” American Antiquity 55(3):480–499. http://dx.doi.org /10.2307/281279. Schick, Kathy D., Nicholas Toth, Gary Garufi, E. Sue Savage-Rumbaugh, Duane M. Rumbaugh, and Rose Sevcik. 1999. “Continuing Investigations into the Stone Tool-Making and Tool-Using Capabilities of a Bonobo (Pan paniscus).” Journal of Archaeological Science 26(7):821–832. http://dx.doi.org/10.1006/jasc.1998.0350. Toth, Nicholas, Kathy D. Schick, E. Sue Savage-Rumbaugh, Rose A. Sevcik, and Duane M. Rumbaugh. 1993. “Pan the Tool-Maker: Investigations into the Stone Tool-Making and Tool-Using Capabilities of a Bonobo (Pan paniscus).” Journal of Archaeological Science 20(1):81–91. http://dx.doi.org/10.1006/jasc.1993.1006. White, J. Peter, and Harold L. Dibble. 1986. “Stone Tools: Small-Scale Variability.” In Stone Age Prehistory: Studies in Memory of Charles McBurney, ed. G. N. Bailey and P. Callow, 47–53. Cambridge: Cambridge University Press. White, Randall K., Nico Arts, Paul G. Bahn, Lewis R. Binford, Michel Dewez, Harold L. Dibble, Paul R. Fish, Clive Gamble, Christopher Meiklejohn, Milla Y. Ohel, et al. 1982. “Rethinking the Middle-Upper Paleolithic Transition.” Current Anthropology 23(2):169–192. http://dx.doi.org/10.1086/202802. Wynn, ThomasG . 2004. “Comment on McNabb, Binyon and Hazelwood: The Large Cutting Tools from the South African Acheulean and the Question of Social Traditions.” Current Anthropology 45:672–673.

F orm, Function, and Mental Templates in Paleolithic Archaeology 305

14

Understanding chipped-stone technology essentially The Role of Controlled comes down to understanding how a single flake is Experiments in made. Even though a knapper may remove many flakes Understanding Variation to prepare a core or shape a piece, thin it, or modify its in Flake Production edges, each and every one of these removals requires a certain degree of control so that particular effects are achieved. While there is currently an emphasis in Zeljko Rezek, Sam Lin, lithic analysis on reconstructing the totality of vari- and Harold L. Dibble ous reduction sequences (e.g., see Olszewski, chapter 4, this volume; Rollefson, chapter 10, this volume), we still have much to learn about how individual flakes are formed. In order to quantify the various mechani- cal aspects of flake production, a number of research- ers have designed experiments under highly controlled conditions, often using shaped cores and mechanical strikers. These types of experiments allow control over several aspects of flaking, from the angle and force of the blow and type of hammer used to various core sur- face and platform morphologies. In this way, it is pos- sible to study in a very detailed fashion the effects of particular independent variables on flake morphology. Such experiments are called “controlled experiments” because their goal is to control as many as possible of the variables involved with knapping in order to isolate the effects of a single variable. Because of the design of these kinds of experiments, such controls are much tighter than is generally possible with replicative flint- knapping experiments, and the results are clearly more amenable to quantification. On the other hand, the DOI: 10.5876/9781607324942.c014

307 downside of many controlled experiments is that the processes and products do not always accurately simulate archaeological ones. This artificial nature inherent to many controlled experiments has often made it difficult to apply the results directly to archaeological materials. Controlled experiments in lithic technology began more than 40 years ago, but there has been little consistency in terms of the types of cores used or the ways in which force has been applied to them. One of the biggest problems is that most of the previous controlled experiments relied on core shapes that do not resemble those used in prehistoric assemblages, due primarily to the dif- ficulty of obtaining suitable experimental cores. In fact, until recently, virtually all of the experiments performed thus far have been part of the investigators’ doctoral research (Speth 1972, 1974; Faulkner 1992; Bonnichsen 1977; Dibble and Whittaker 1981; Pelcin 1996, 1997a, 1997b, 1998), and thus were designed to incorporate materials that were both inexpensive and readily available, such as plate glass. Unfortunately, such materials often only allowed for limited varia- tions in core morphology, and they never really approached the form of actual cores. The same is true of the devices constructed for the delivery of the force, which has varied from steel ball bearings dropped from an electromagnet (Speth 1972; Dibble and Whittaker 1981; Dibble and Pelcin 1995) to devices used to push or pull on plate glass (Cotterell and Kamminga 1987). In general, those devices suffered from the inability to treat some variables independently of each other, namely the speed or the mass of the hammer and the overall force the hammer can apply. In addition to difficulties that exist in applying the results to the archaeological record, the results from one experiment are not always directly comparable to the results of another. While these types of experiments have a great deal of potential for under- standing how prehistoric knappers approached their materials, there is a clear need to develop a set of standards so that future experiments can be designed to enhance comparability both among themselves and with actual archaeo- logical assemblages.

Ovv er iew of Previous Experimental Designs Speth Speth (1972, 1974, 1975, 1981) was one of the first to investigate the quanti- tative relationships among various independent variables and resulting flake attributes using controlled experiments for flake formation. His experimental design consisted of dropping steel balls onto core specimens from an electro- magnet attached to a drop tower of a certain height. The cores used in these

308 Rezek, Lin & Dibble experiments were glass optical prisms obtained from military tank periscopes (figure 14.1A). These cores were triangular in cross-section and by using tri- angular glass prisms of two different geometries in their cross-section (equi- lateral and right-angle) effects of two different exterior platform angles were tested. He also used balls of different diameters (and thus of different mass) and dropped them from varying heights (which varied their velocity) onto the cores’ platform surfaces at varying distances from the exterior edge (thus vary- ing platform depth). In addition, he was able to change the mount holding the core so that the angle of blow (the angle of the platform surface relative to ver- tical) could be varied. Thus, Speth isolated effects of five independent variables (hammer mass and velocity, exterior platform angle, and platform depth, and angle of blow), and examined their effects on flake length and width. The flakes produced in this set of experiments resembled archaeological flakes to some degree, but they were characteristically very short and wide, with what Speth called “wings” on their lateral margins. They always exhib- ited hinged terminations. Both of these characteristics are common to other experiments that removed flakes from flat surfaces, rather than the surfaces seen in most prehistoric cores, which tend to exhibit convexity both in their longitudinal (long axis) and lateral (side-to-side) aspects. The edges of the prisms were slightly beveled and the size of this bevel varied from prism to prism, which may have introduced uncontrolled variation in the exterior sur- face morphology of the cores. Another shortcoming of Speth’s experimental design had to do with the application of force. Although force is often thought of as simply being how hard an object (in this case, the core) is struck, in this case it is more useful to use the term “momentum,” defined as the product of mass and velocity. As is discussed below, many experiments followed the same protocol as Speth in varying momentum, either by using ball bearings of dif- ferent sizes (i.e., of different mass) and/or by changing the height from which they fell (i.e., changing velocity). This relationship can easily confound the relative effects of mass and velocity, and furthermore, when relying on gravity to increase the velocity of the falling ball bearing, it is also necessary to take into account the acceleration that will increase with drop height.

Faulkner In his dissertation, Faulkner (1992) presented the results of controlled experiments in which the primary objective was the description of the frac- ture mechanics and stress distribution that occur during the manufacture of Mesoamerican blades. He used two types of core specimens: either rectangles

U nderSTANDINg Variation in Flake Production 309 Figure 14.1. Schematic illustrations of core specimens and flake outcomes from the experimental work of (A) Speth (1974:24); (B) Faulkner (1992:99), Dibble and Whittaker (1981:285), Dibble and Pelcin (1995), and Pelcin (1996:88); (C) Pelcin (1996:89); (D) Cotterell et al. (1985:216); and (E) Dibble and Rezek (2009). In addition to these exam- ples, quasi-cylindrical obsidian cores were used (nonexten- sively) by Faulkner (1992), spherical glass cores were used by Pelcin (1996), and rectangular plate-glass cores (along with obsidian and quartzite cores of an unspecified form) were used by Bonnichsen (1977). The majority of flakes obtained from example (B) were overshots, while all flakes from example (A) had hinged terminations. of plate glass that were struck to remove flakes from their straight edges (fig- ure 14.1b) or cylindrical cores prepared from obsidian; the latter specimens exhibited a lateral convex surface that was supposed to imitate a dorsal ridge on cores used for blade production. His experimental setup consisted of a strain frame in which the glass specimen was placed firmly in the special vice. Pressure was exerted through a hydraulic ram, which was also able to pivot in order to adjust angle of blow; a pressure gauge indicated the amount of hydraulic pressure applied before a flake was detached. Thus, Faulkner was the first to use static loading (a gradual buildup of pressure) versus the dynamic loading (a quick strike to the platform surface), although it is not clear that hammer velocity was controlled. The hammer tip material used was copper, and it was possible for him to place the hammer at different platform depths. Based on his limited testing of cylindrical cores, he concluded that ridges and convexities on the core surface can be used to manufacture longer flakes. The majority of flakes that resulted from Faulkner’s experiments (those removed from the edges of plate glass) tended to resemble burin spalls, though a high percentage of them were overshot. Of course, the flakes from the plate glass cores all had identical widths (i.e., the thickness of the plate glass), and so this dependent variable could not be included in the analyses. This was the same issue that affected later experiments using these kinds of cores.

Bonnichsen Bonnichsen (1977) was the first to conduct controlled experiments that altered both the material of the hammer (moose antler and fine-grained quartzite) and of the cores. In addition to the rectangular plate glass cores, he used obsidian and quartzite cores of an unspecified shape. Again, the momen- tum of the hammer was achieved essentially by dropping it from varying heights (the trajectory of the hammer being controlled by attaching it loosely to a metal rod), though again hammer mass and velocity were confounded. Information on platform depth was not presented. Angle of blow was varied not by changing the core’s platform surface with respect to the vertical, but rather by cutting the platform surface at different angles from its exterior sur- face. While this did affect the angle at which the core’s platform surface was struck, it simultaneously changed the core’s exterior platform angle, which resulted in the confounding of the effects of those two independent variables as well. Bonnichsen was also the first and only experimenter to control the pressure with which the core was held in a vise.

U nderSTANDINg Variation in Flake Production 311 Dibble and Whittaker The controlled experiments of Dibble and Whittaker (1981) followed many of the experimental designs of Speth, in that they also dropped steel balls of different sizes (but not from different heights) onto glass cores. Like in the experiments done by Faulkner and Bonnichsen, plate glass was used, with flakes struck along the edges (figure 14.1b). Again, their flakes resembled burin spalls most closely, although in their case termination (feather, hinge, or over- shot) became a dependent variable that was analyzed in relation to the other independent variables. Angle of blow was varied by mounting the cores at different angles relative to the vertical, and a much higher range of exterior platform angles was formed by cutting the plate glass. Platform depth was also varied. Dependent variables included flake length and thickness, interior platform angle, and flake termination.

Dibble and Pelcin The earlier controlled experiments by Speth and by Dibble and Whittaker had suggested that there was a direct relation between the momentum and the size of a flake, expressed primarily in its length and thickness. Dibble and Pelcin (1995) tested these relationships in a different way and showed that the direct effects of momentum on flake attributes are minimal. Their experimen- tal design was identical to that of Dibble and Whittaker (figure 14.1b), but this time they isolated the independent effects of hammer velocity and mass in addition to controlling exterior platform angle and platform depth. When they varied velocity but held the overall momentum constant by using an appropriately sized steel ball, it was clear that velocity alone makes no differ- ence in flake mass. The same was true when they used balls of different sizes but adjusted their velocity (through changes in drop height) so that the same momentum was achieved. Flake shape was still in the form of burin spalls.

Pelcin Pelcin (1996, 1997a, 1997b; 1998) continued with controlled experiments as part of his doctoral research, adding several new controls. Although in some of his experiments he continued with the use of steel balls, he also developed a striking apparatus that fell in an arc instead of straight down. The primary reason for this was so that he could attach different kinds of materials to use as hammers—one was a steel ball and the other was antler ground down to the same size and shape. Weights were added to the striking apparatus so that

312 Rezek, Lin & Dibble the overall weight was the same for either type of hammer. Although plate glass was used for many of his experiments, in some cases the edge of the glass was removed (as in Faulkner’s and Dibble and Whittaker’s experiments) (figure 14.1b) and in other cases flakes were removed from the flat surface by striking the glass plate on the adjacent edge (figure 14.1c). Pelcin also used glass lenses for his beveling experiment, because those cores had spherical surface morphology that introduced an external bevel in the area just below and exterior to the platform. Altogether, Pelcin varied 8 independent variables (exterior platform angle, bevel angle, platform depth, momentum, angle of blow, hammer material, area of the hammer tip, and core surface morphology) and studied their effects on 11 dependent variables (length, width, thickness, weight, interior platform angle, flake termination, presence/absence of plat- form lipping, ring crack diameter, bulb length and thickness, and expansion angle). In some of Pelcin’s experiments cores were struck multiple times.

Cotterell, Kamminga, and Dickson Interest in fracture propagation led Cotterell, Kamminga, and Dickson (1985; Cotterell and Kamminga 1986, 1987, 1990) to perform a series of controlled experiments that, in many ways, are very dissimilar to the others reviewed here, although the cores were still plate glass with flakes removed from the edge. While in other experiments a compressive load was applied to the core (i.e., the cores were struck by a hammer), in this case the force was applied by tensioning special load arms attached to the cores—in other words, flakes were pulled off the edges of the plate glass (figure 14.1d). In addition, before applying this tensile load, a fracture was initiated on the cores by a saw cut and by application of heat to produce a crack (Cotterell et al. 1985). By varying the load arms, the angle at which the tensile load was applied could be varied. Other independent variables that had been the focus of earlier experiments, such as exterior platform angle and platform depth, were not experimen- tally varied, and of course there was no hammer used. Dependent variables included flake termination and the direction of the fracture propagation.

Dibble et al. Dibble and Rezek (2009) introduced a new experimental design to more closely resemble actual flint-knapping. There were two primary changes. First, glass cores were molded to have a flaking surface that was convex in both the longitudinal and lateral aspects (figure 14.1e). This allows for the production

U nderSTANDINg Variation in Flake Production 313 of flakes that are identical in all respects to those found in the archaeological record. Second, their flaking apparatus used a pneumatic cylinder to apply the compressive load to the core’s platform surface. The use of a pneumatic cylinder allows them to vary the hammer’s velocity, although the maximum potential pressure that is applied remains constant. In this way, the load can be either static or dynamic, and materials used for the hammer tip can be varied as well. A load cell attached to the hammer measures the actual load applied to the core up to the point where the flake is detached. Building on the experimental design of Dibble and Rezek, a series of studies were conducted by Dibble and other colleagues to investigate various aspects of flake formation (Rezek et al. 2011; Lin et al. 2013; Magnani et al. 2014). For example, Rezek et al. (2011) examined the effects of different core sur- face morphologies on various aspects of flake size and shape. To do this, they designed cores that had five different surface morphologies as defined by dif- ferent patterns of ridges—convergent, divergent, parallel, a single longitudinal ridge, and no ridges. In all other aspects, including degree of both longitudinal and lateral convexity, the cores were identical. The only other independent variables included in the experiment were exterior platform angle and plat- form depth. Using the same experimental setup, Magnani et al. (2014) investigated the interaction between flake variability and various variables of force application, including hammer material and shape, location of force application, angle of blow, and hammer displacement speed. The authors employed steel, copper, and synthetic bone as hammer material to simulate different hammer hard- ness and material properties. Two locations of force application were tested: either the hammer strikes directly on the platform surface some distance away from the core exterior, or the hammer strikes the exterior platform edge, simi- lar to techniques associated with biface thinning. The effects of hammer dis- placement speed were tested between static loads (where pressure increases slowly) at a speed of 0.05 ipm (0.002 cm/sec) and dynamic loads (i.e., a quick blow to the platform) at a speed of 65 ipm (27.516 cm/sec). The two types of displacement speed roughly simulate the general separation between pressure flaking and direct percussion, respectively.

Discussion Over the course of their 40-year history, controlled experiments in flake production have produced several results that have significant implications for lithic research. Although some of the experiments (primarily the earlier ones)

314 Rezek, Lin & Dibble were focused on the fracture mechanics underlying conchoidal flaking, there has also been concern on isolating the variables that flint-knappers themselves control to get desired results. Over the years, some of these findings have been applied to actual archaeological assemblages, both from the point of view of trying to confirm or deny the experimental results, and in some cases, to make further inferences regarding prehistoric behavior (e.g., Lin et al. 2013). However, it is quite clear that controlled experiments have a long way to go in isolating major independent variables that underlie significant variation in flake production. In fact, at this point, the single biggest cause-effect relation- ship that has been repeatedly demonstrated is the role of exterior platform angle and platform depth on flake size. Numerous experiments (Dibble and Pelcin 1995; Dibble and Rezek 2009; Dibble and Whittaker 1981; Pelcin 1996, 1997b; Speth 1972, 1974) and certain analyses of lithic assemblages (Braun et al. 2008; Dibble 1997) have shown that increasing either or both exterior platform angle and platform depth will result in larger flakes. At the same time, the amount of force applied to the core as well as the speed at which the hammer strikes do not influence flake size, but rather, there is only a minimum amount of force necessary to detach a flake of a given size. In other words, larger flakes require more force in order to be detached from the core—any less force, and a flake will not propagate even though the fracture may have initiated, and the application of more force beyond the minimum will have no effect. There are two things that are especially interesting about these results. The first is that these findings isolate three independent strategies that prehistoric flint-knappers may have used to produce flakes of desired sizes—either by changing the exterior platform angle, changing platform depth, or changing both. It is already been shown that certain archaeological industries tend to emphasize one of these independent variables over the other (Braun et al. 2008; Dibble 1997; Lin et al. 2013). From a technological point of view, this means that the varying emphasis on one of these independent variables versus the other provides a means of characterizing the strategies employed by dif- ferent groups. Although exactly why one approach versus the other would be adopted is still unknown, studies have linked changes in relative flake size to flake utility and functional efficiency (Dibble 1997; Key and Lycett 2014; Lin et al. 2013; Prasciunas 2007). Flake shape has been more difficult to explain. For many flint-knappers and lithic analysts, flake shape is often thought to be largely, if not com- pletely, under the control of core surface morphology. This conclusion has also been reached on the basis of some controlled experiments (Faulkner, Pelcin) where the cores used in these experiments exhibited extreme differences.

U nderSTANDINg Variation in Flake Production 315 Pelcin (1996), for example, concluded that the increase in platform depth will increase the flake mass, but the core surface morphology determines how that mass is distributed in flake size dimensions, affecting, therefore, the shape of the flake. As demonstrated in the experiments by Rezek et al. (2011), and confirmed in their analysis of flakes from archaeological assemblages, more realistic (from an archaeological sense) variation in core surface morphologies (at least in terms of surface edge configurations) did not have as great an effect as did exterior platform angle and platform depth. However, there are still a tremendous number of factors that may underlie flake shape that have not yet been explored, and so it is still too early to say that we understand fully the shape variability. To a very large degree, the biggest problem in trying to explain flake shape has been with the designs of the cores used in most of the controlled experi- ments. Removing flakes from a totally flat surface (as in the experiments of Speth and Pelcin, for example), or removing them from the edges of plate glass (as in Dibble and Whittaker, Dibble and Pelcin, etc.), does not result in very realistic flakes. This has been one of the primary deficiencies of many of the experiments done so far, and it is one reason why the use of more realistic core designs is essential for future experiments. On the other hand, Magnani et al. (2014) reported that hammer material, location of force application, and the angle of blow all contribute in varying degrees to the variation in flake morphology. In particular, using softer hammer material and a lower angle of blow, and striking the platform edge, makes flakes more elongated and thinner and also increases the chance of obtaining flakes with lipped platforms that resemble the so-called soft-hammer flakes. An important implication of this finding is that the type of flake morphological characteristics (e.g., platform lipping) commonly associated with advanced percussion methods, such as the use of soft hammer and pressure flaking, can in fact also be achieved by other combinations of force application without the involvement of these innova- tive force application techniques.

Cuoncl sions Of the two major experimental approaches used in lithic research, replicative and controlled, the latter offer a much greater scientific value for quantifying and confirming causal relationships between the actions that a flint-knapper­ takes and the innumerable attributes expressed on the resulting flakes. In a well-designed controlled experiment individual variables can be isolated and controlled to a much higher degree than is possible with replicative experiments,

316 Rezek, Lin & Dibble in which there is almost always a significant level of confounding. This means that controlled experiments are able not only to recognize the effects that a single independent variable has on a flake, but also determine more precisely the extent of those effects. At the same time, a highly controlled design allows us to see if a particular flake attribute is governed by one independent variable or a combination of two or more independent variables. Likewise, controlled experiments are much more amenable to independent verification, since theo- retically all of the same conditions can be reproduced exactly. So far, the role of exterior platform angle and platform depth in affecting flake size has been confirmed in several controlled experiments, even though they have employed somewhat different experimental designs. This does not deny an important role for replicative experiments, which have been tremendously influential on lithic studies. At the same time, however, differences in skill and inherent variation in raw materials, hammers, and the like make it much more difficult for replicative experiments to achieve a significant level of control. Then again, it has been pointed out earlier that there exists a challenge where the effects of the numerous variables in knapping are all expressed in a limited number of flake attributes. It is therefore extremely difficult to dis- cern exactly which independent variable(s) were responsible for the resulting flake morphology. This is largely an issue of equifinality, by which different configurations of the independent variables can result in the same flake mor- phology. This, in turn, means that traditional flint-knapping experiments may demonstrate only particular ways out of many of achieving a specific out- come (Hayashi 1964). While replicative experiments arguably are of greater resemblance to real-life flint-knapping situations, these studies alone cannot justify an inductive interpretation about the past based on the limited range of modern observations. Rather, it is essential that the potential relationships identified by replicative experiments be broken down to independent variables and further tested under controlled settings. On the other hand, in spite of their potential, this review has shown that controlled lithic experiments have many limitations that need to be overcome. Perhaps the single biggest issue so far has often been that neither the pro- cesses involved in flake removals nor the products themselves resemble very closely their prehistoric counterparts. This in turn can lead to questions about the applicability of the results to analyses of archaeological lithic assemblages (e.g., Davis and Shea 1998; cf. Dibble 1997; Braun et al. 2008; Nonaka et al. 2010). However, and as discussed in Magnani et al. (2014), the applicability of the results of controlled experiments to archaeological assemblages is not in the high resemblance of experimental design to the empirical reality and

U nderSTANDINg Variation in Flake Production 317 “real-world” settings in flake production, but rather in its ability to isolate and determine the independent causal effects of individual variables (a quality known as “internal validity” [Kirk 2012]). The latest experiments by the pres- ent authors have made significant improvements in both of these regards, but we argue that the primary concern in conducting lithic experimental studies should be in their internal validity, while the resemblance of processes involved in flaking and flake removals to their archaeological counterparts is of second- ary importance. In other words, any improvement in the comparability between experimentally produced specimens and archaeological materials should not be made on the expense of the experimental design. This, rather than anything else related to controlled lithic experiments, makes possible the application of experimental results to the archaeological record (Dibble 1997; Lin et al. 2013; Rezek et al. 2011). Notwithstanding, it is clear that controlled experiments have already demonstrated their relevance to lithic studies, and it is equally certain that future experiments will continue to inform us of exactly how prehistoric flint-knappers went about making their stone tools, as well as of the fundamen- tal properties that gave rise to lithic variability among prehistoric assemblages.

References Bonnichsen, Robson. 1977. Models for Discovering Cultural Information from Stone Tools. Ottawa: National Museums of . Braun, David R., Michael J. Rogers, John W. K. Harris, and Steven J. Walker. 2008. “Landscape-Scale Variation in Hominin Tool Use: Evidence from the Developed Oldowan.” Journal of Human Evolution 55(6):1053–1063. http://dx.doi.org/10.1016/j .jhevol.2008.05.020. Cotterell, Brian, and Johan Kamminga. 1986. “Finials on Stone Flakes.” Journal of Archaeological Science 13(5):451–461. http://dx.doi.org/10.1016/0305-4403(86)90014-2. Cotterell, Brian, and Johan Kamminga. 1987. “The Formation of Flakes.” American Antiquity 52(4):675–708. http://dx.doi.org/10.2307/281378. Cotterell, Brian, and Johan Kamminga. 1990. Mechanics of Pre-Industrial Technology. Cambridge: Cambridge University Press. Cotterell, Brian, Johan Kamminga, and Frank P. Dickson. 1985. “The Essential Mechanics of Conchoidal Flaking.” International Journal of Fracture 29(4):205–221. http://dx.doi.org/10.1007/BF00125471. Davis, Zachary J., and John J. Shea. 1998. “Quantifying Lithic Curation: An Experimental Test of Dibble and Pelcin’s Original Flake-Tool Mass Predictor.” Journal of Archaeological Science 25(7):603–610. http://dx.doi.org/10.1006/jasc.1997.0255.

318 Rezek, Lin & Dibble Dibble, Harold L. 1997. “Platform Variability and Flake Morphology: A Comparison of Experimental and Archaeological Data and Implications for Interpreting Prehistoric Lithic Technological Strategies.” Lithic Technology 22:150–170. Dibble, Harold L., and Andrew W. Pelcin. 1995. “TheE ffect of Hammer Mass and Velocity on Flake Mass.” Journal of Archaeological Science 22(3):429–439. http://dx .doi.org/10.1006/jasc.1995.0042. Dibble, Harold L., and Zeljko Rezek. 2009. “Introducing a New Experimental Design for Controlled Studies of Flake Formation: Results for Exterior Platform Angle, Platform Depth, Angle of Blow, Velocity, and Force.” Journal of Archaeological Science 36(9):1945–1954. http://dx.doi.org/10.1016/j.jas.2009.05.004. Dibble, Harold L., and John C. Whittaker. 1981. “New Experimental Evidence on the Relation between Percussion Flaking and Flake Variation.” Journal of Archaeological Science 8(3):283–296. http://dx.doi.org/10.1016/0305-4403(81)90004-2. Faulkner, Alaric. 1992. “Mechanical Principles of Flintworking.” PhD dissertation, Washington State University, Pullman. Hayashi, Kensaku. 1964. “The Fukui Microblade Technology and its Relationships in Northeast Asia and North America.” Arctic Anthropology 5:128–190. Key, Alastair J. M., and Stephen J. Lycett. 2014. “Are Bigger Flakes Always Better? An Experimental Assessment of Flake Size Variation on Cutting Efficiency and Loading.” Journal of Archaeological Science 41:140–146. http://dx.doi.org/10.1016/j .jas.2013.07.033. Kirk, Roger E. 2012. Experimental Design: Procedures for the Behavioral Sciences. 4th ed. Thousand Oaks, CA: Sage Publications. http://dx.doi.org/10.1002/9781118133880 .hop202001. Lin, Sam C., Zeljko Rezek, David Braun, and Harold L. Dibble. 2013. “On the Utility and Economization of Unretouched Flakes: The Effects of Exterior Platform Angle and Platform Depth.” American Antiquity 78(4):724–745. http://dx .doi.org/10.7183/0002-7316.78.4.724. Magnani, Matthew, Zeljko Rezek, Sam C. Lin, Annie Chan, and Harold L. Dibble. 2014. “Flake Variation in Relation to the Application of Force.” Journal of Archaeological Science 46:37–49. http://dx.doi.org/10.1016/j.jas.2014.02.029. Nonaka, Tetsushi, Bril Blandine, and Robert Rein. 2010. “How Do Stone Knappers Predict and Control the Outcome of Flaking? Implications for Understanding Early Stone Tool Technology.” Journal of Human Evolution 59(2):155–167. http://dx .doi.org/10.1016/j.jhevol.2010.04.006. Pelcin, Andrew W. 1996. “Controlled Experiments in the Production of Flake Attributes.” PhD dissertation, University of Pennsylvania, Philadelphia.

U nderSTANDINg Variation in Flake Production 319 Pelcin, Andrew W. 1997a. “The Effect of Core Surface Morphology on Flake Attributes: Evidence from a Controlled Experiment.” Journal of Archaeological Science 24(8):749–756. http://dx.doi.org/10.1006/jasc.1996.0156. Pelcin, Andrew W. 1997b. “The Formation of Flakes: The Role of Platform Thickness and Exterior Platform Angle in the Production of Flake Initiations and Terminations.” Journal of Archaeological Science 24(12):1107–1113. http://dx.doi.org /10.1006/jasc.1996.0190. Pelcin, Andrew W. 1998. “The Threshold Effect of Platform Width: A Reply to Davis and Shea.” Journal of Archaeological Science 25(7):615–620. http://dx.doi.org/10.1006 /jasc.1997.0253. Prasciunas, Mary M. 2007. “Bifacial Cores and Flake Production Efficiency: An Experimental Test of Technological Assumptions.” American Antiquity 72(2):334– 348. http://dx.doi.org/10.2307/40035817. Rezek, Zeljko, Sam Lin, Radu Iovita, and Harold L. Dibble. 2011. “The Relative Effects of Core Surface Morphology on Flake Shape and Other Attributes.” Journal of Archaeological Science 38(6):1346–1359. http://dx.doi.org/10.1016/j .jas.2011.01.014. Speth, John D. 1972. “The Mechanical Basis of Percussion Flaking.” American Antiquity 37(1):34–60. http://dx.doi.org/10.2307/278884. Speth, John D. 1974. “Experimental Investigations of Hard-Hammer Percussion Flaking.” Tebiwa 17:7–36. Speth, John D. 1975. “Miscellaneous Studies in Hard-Hammer Percussion Flaking: TheE ffects of Oblique Impact.” American Antiquity 40(2):203–207. http://dx.doi .org/10.2307/279616. Speth, John D. 1981. “The Role of Platform Angle and Core Size in Hard-Hammer Percussion Flaking.” Lithic Technology 10:16–21.

320 Rezek, Lin & Dibble Contributors

C. Michael Barton Professor School of Human Evolution and Social Change Arizona State University Tempe, AZ 85281 [email protected]

Brian F. Byrd Principal Investigator Far Western Anthropological Research Group Davis, CA 95618 [email protected]

Gerald Cadogan 3 The Old Rickyard Moreton Pinkney Daventry, NN 11 3TL, [email protected]

Philip G. Chase Consulting Scholar University of Pennsylvania Museum of Archaeology and Anthropology Philadelphia, PA 19104-6398 [email protected] Harold L. Dibble Kathryn A. Kamp Professor Professor Department of Anthropology Department of Anthropology University of Pennsylvania Grinnell College Philadelphia, PA 19104-6398 Grinnell, IA 50112 [email protected] [email protected]

Matthew J. Douglass Sam Lin Lecturer Centre for Archaeological Science Department of Anthropology University of Wollongong University of Nebraska Wollongong NSW 2522, Australia Lincoln, NE 68588 [email protected] [email protected] Emilia Oddo Patricia C. Fanning Assistant Professor Associate Professor Department of Classical Studies Graduate School of the Environment Tulane University Macquarie University New Orleans, LA 70118 NSW 2109, Australia [email protected] [email protected] Deborah I. Olszewski Lynne Goldstein Adjunct Professor Professor Department of Anthropology Department of Anthropology University of Pennsylvania Michigan State University Philadelphia, PA 19104-6398 East Lansing, MI 48824 [email protected] [email protected] Zeljko Rezek Simon J. Holdaway Department of Human Evolution Professor Max Planck Institute for Department of Social Sciences Evolutionary Anthropology University of Auckland D-04103 Auckland 1010, New Zealand Leipzig, Germany [email protected] [email protected]

322 C ontRIBUTORS Julien Riel-Salvatore Justin I. Shiner Associate Professor Operations Manager Department of Anthropology Australian Cultural Heritage University of Montreal Management Montreal, QC, Canada Sydney, Australia [email protected] [email protected]

Gary O. Rollefson Alan P. Sullivan III Emeritus Professor Professor Department of Anthropology Department of Anthropology Whitman College University of Cincinnati Walla Walla, WA 99362 Cincinnati, OH 45221 [email protected] [email protected]

Jeffrey Rosenthal John C. Whittaker Principal Investigator Professor Far Western Anthropological Department of Anthropology Research Group Grinnell College Davis, CA 95618 Grinnell, IA 50112 [email protected] [email protected]

Barbara J. Roth David R. Wilcox Professor Research Associate Department of Anthropology Arizona State Museum University of Nevada University of Arizona Las Vegas, NV 89154 Tucson, AZ 85721 [email protected] [email protected]

Sissel Schroeder Professor Department of Anthropology University of Wisconsin, Madison Madison, WI 53716 [email protected]

C ontRIBUTORS 323

Index

Aborigines, 8; population mod- ceremonial, 268; Myrtos- Page numbers in italics els of, 114–15 Pyrgos, 177, 178; ritual, 59; of indicate illustrations. Acheulean bifaces, 298 1000+ room pueblos, 204–5 Acoma, Flint Bird, 282 Arch Lake site, 127–28, 139 Adaptive Diversity Hypothesis, Arizona, Upper Basin, 9, 126 9, 125, 128, 139 Arroyo Hondo, 201, 204 adornment, 234; Natufian, artifacts, 5, 8, 11; intentionally 239–41; role of, 250–51; San damaged, 268–69; meaning Francisco Bay area hunter- and symbolism of, 267–68; gatherers, 245–46, 247 spatial distribution of, 6–7 adze, tula, 102, 104 artifact scatters: lithic, 125, agriculturalists, Neolithic 127–28; variability in, 136–39 Levant, 242–43 Australia, 8, 109; environment, Ain Keda points, 77, 78 101–2; population models, ‘Ain Soda: bifaces from, 223–24, 113–15; stone artifact assem- 226(n4); cleavers from, 215, blages, 99, 102–4, 116 216, 222 axe heads, from New South AMH. See anatomically mod- Wales, 102 ern humans Azraq area, cleavers from, 214, amphorae, from Myrtos- 222 Pyrgos cistern, 180 Aztalan (Wisc.), 157; anti- analogy, ethnographic, 279 quarian maps of, 155–56; anatomically modern humans archaeological traditions at, (AMH), 7, 36, 39; land use, 153–54; Barrett’s interpreta- 34, 41 tion of, 160–61; as coalescent ancestors, 57, 252, 284; associa- community, 166–67; forma- tion with stone tools, 282–83 tion history of, 163–64; Apaches, and ancient places, Mississippian construc- 284 tion at, 165–66; structure archaeological record, 3, 4; dis- of, 9–10; Wisconsin carded artifacts in, 300–301; Archaeological Survey work lithic production and use at, 161–63 in, 28–29 Aztecs, and Toltecs, 284 archaeological sites, ritual associations of, 283–84 Bandelier, Adolph, 56; and architecture: of Aztalan, 160– Casas Grandes, 192, 199–200, 63; Casa Grande, 10, 193–97; 206(n1), 207(n6) Barrett, Samuel, and Aztalan, 156, 160–61 Casas Grandes (Chih.), 9, 10, 201, 208(n12); Bartlett, John Russel, on Casas Grandes, 192, dates of, 207–8(n11); historical accounts of, 198–99, 206(n2), 207(n7) 197–200, 206(n1, n2), 207(n6, n7); size of, base camps, 29 191–92, 207(n8, n10); spatial organization Beauregard site, 57–58, 60, 61, 64 of, 204–6 biface:core ratios, 133 caves, Neanderthal, 26 bifaces: Acheulean, 298; from Azraq area, cemeteries, Early Natufian, 239 214–15, 216, 226(n4); Bordes typology, ceramics: from Aztalan, 156; from Casas 213–14, 217; from Grasshopper Pueblo, Grandes, 207–8(n11); from Myrtos-Pyrgos, 268–69, 274–76; life history of, 221–22; 175, 179–80, 181–83, 184–85 measurements of, 223–24; standardized ceremonialism, San Francisco Bay area forms, 224–25, 299; from Tabun Cave, 10–11, hunter-gatherers, 244 219–20, 225–26(n1); in US Southwestern ceremonial sites, in New South Wales, 111–13, ritual contexts, 272–73; from Upper Basin, 115 132–33, 136 chaine opératoire, 87 Binford, Lewis R., 4, 268 chert: ritual bifacial knives of, 272–73 biogeographic change, and species interac- Chihuahua. See Casas Grandes tions, 39–40 children, mortuary assemblages for, 235 Blackwater end-scraper retouch flakes, 127, chipped-stone assemblages, Old World 128 typologies, 71 bladelets: Epipaleolithic, 74–75, 76, 77; manu- choppers, Oldowan, 72 facture of, 80–81, 82–83, 87 Cistern 1 (Myrtos-Pyrgos), 177, 187 blades, 102; uniformity of, 298–99 Cistern 2 (Myrtos-Pyrgos), 10, 177, 178, 182; Bonnichsen, Robson, flake production experi- ceramics from, 175, 179–80, 187; deposi- ments, 311 tional history of, 181, 183, 184–85 Bordes, François, 71; biface typologies, 213–14, Classic period Mimbres, land use, 53, 54 217, 219 cleavers, 215, 222; Bordes definition of, 213–14; Bordes-Binford debate, 72, 221 from Tabun Cave, 11, 214, 217, 218, 220, Braidwood, Robert, 74 226(n3) Broken K Pueblo, 127 Cliff Valley, 56 Bronze Age, Minoan, 10, 175, 176, 178 climate, Australian, 101 burials: age-graded goods in, 251–52; grave climate change: and land-use strategies, 34, goods in, 6, 271–72, 273, 278; in Levant, 41–42; and persistent places, 64–65; Upper 236–43; Livermore-Amador Valley, 245–50; Paleolithic, 31, 33 in New South Wales sites, 112–13 coalescent communities, 10; Aztalan as, burins, 72, 295 166–67 Burkes Cave, 100, 112 Coalescent Communities Database, 201, 203, Burro Mountains, 56 207(n8) butchering sites, tools at, 215, 222 cognitive complexity, 12 community identity, Mimbres Mogollon, 55, caches, with stone tools, 276 64 Cahokia, 157; occupational development of, compound tools, Upper Pleistocene, 37 159–60 controlled experiments, in flake production, California, hunter-gatherer societies, 234, 235 307–18 Cameron Creek pueblo, 56 cores: burins as, 72; cleavers as, 224; in flake campsites, camps, 129; Neanderthal, 25–26 production experiments, 308–14, 315–16; cartography, archaeological, 153 in New South Wales assemblages, 102; in Casa Grande (Ariz.), 9, 10, 192, 193, 194, 195, Upper Basin assemblages, 133 196, 197, 206(n1) cortex ratios, and raw material sources, 106–7

326 Index Cotterrell, Brian, flake production experi- Faulkner, Alaric, flake production experi- ments, 313 ments, 309, 311 Country House (Myrtos-Pyrgos), 177, 186–87 feasting: mortuary practices and, 236; at Crete, Minoan, 175, 176. See also Myrtos-Pyrgos Myrtos-Pyrgos, 10, 184–87 cross-joins, in Myrtos-Pyrgos ceramics, 180, features, synchronicity of features on, 159–60 182–83 fetishes, Zuni, 280, 281 cultural elaboration, and mobility, 113 Fewkes, Jesse Walter, 56 curation, of artifacts, 284 fire-cracked-rock-pile sites, lithic assem- blages at, 135 damage, intentional, 268–69 Fitting, James, 56 debitage analysis, 141(n3); Southwest archae- flake cores, from Tabun, 224 ology, 9, 125–27, 133–35 flakes, 12, 38, 102, 299; manufacture of, 296–97, Deetz, James, on mental template, 292–93 307–18; shapes of, 315–16; transport of, Dentalium shell, in Early Natufian graves, 239, 107–9; US Southwest studies of, 127–28, 133, 240, 241 134, 135, 141(n4) depositional history, of Cistern 2 (Myrtos- Flint Bird, 282 Pyrgos), 181, 183, 184–85 Flower World, 279 Dibble, Harold: flake production experiments, food sources, Australia, 109 312, 313–14; on mental template, 295 foragers, foraging, 106; risk management of, Dickson, Frank, 313 109–10 Di Peso, Charles, at Casas Grandes, 191, 201 formation history, at Azatlan, 163–64 drains, at Casa Grande, 195, 197 formation processes, and site size, 112 Dreaming Track sites, 112, 113 Fortress Hills site, 270 Dufour points, 78 Fowlers Gap, 100, 112, 115 Fritz, John M., 127 Early Natufian: burials, 240, 241, 242, 251; socioeconomic strategies, 238–39 Galaz site, 56, 57 Early Period (San Francisco Bay area), 244, Gatton’s Park site, 57, 63 245 gene flow, and biogeographic change, 39–40 Early Pithouse period (Mimbres Mogollon), genomes, modern European, 42 7, 59–60, 62, 64 Georgetown phase, 60, 62, 64 eccentrics, Maya, 272 Ghar Cahar, 77 ecology, Upper Paleolithic, 31 Gila National Forest, 56 economic interaction, Mimbres Mogollon, 7 Gowlett, John A. J., on mental templates, 293 El Niño-Southern Oscillation (ENSO), and Grasshopper Pueblo: bifaces from, 268–69, Australian weather, 101 274–76; ritualized stone-tool use at, 271, environment, 5, 7; Australian, 101–2, 108, 272, 273 110–11 grave goods: age-graded, 251–52; Levant Epipaleolithic: burials, 236; stone-tool assem- hunter-gatherers, 239–41; role of, 250–51; blages, 8, 72–73, 74–75, 76, 77, 78–86 San Francisco Bay hunter-gatherers, 245–46, ethnohistory, ethnography, 9, 114; ritual use of 247–50; symbolic importance of, 271–72 stone at, 278–82 Graybill, Don, 56 Eurasia, 7, 41 ground stone, in New South Wales assem- Europe: climate change in, 31, 33; Middle and blages, 102, 103 Upper Paleolithic, 30–31 expediency, of New South Wales lithic Haliotis shell, in Livermore-Amador hunter- assemblages, 107–8 gatherer burials, 246, 247, 248 Expedient Core Hypothesis, 9, 125, 128–29, 139 hammerstones, New South Wales assem- extinction, Neanderthal, 40, 41 blages, 102

Index 327 Harris site, 56, 58, 62 knowledge transfer, and Aboriginal ceremo- Haury, Emil W., 56 nial life, 115 hearths, heat-retainer, 8, 101–2, 108, 109, 112 heirlooming, of artifacts, 284 Lake Roberts Vista site, 62–63 hero twins, Zuni, 281–82 La Mouillah points, 76, 80; manufacture of, Holmes, William Henry, 156 81, 82, 83, 87 Holocene, 101; Australian Aborigines, 8, 114 landscape, 115; of archaeological sites, 283–84; hominins, 39. See also anatomically modern engineering, 163; and persistent places, 8, humans; Neanderthals 53–55 Hopi, ritually used stone, 272, 280–81 land tenure, 54, 57; Mimbres Mogollon, 60–61 household-debris clearing, at Myrtos-Pyrgos, land-use, 6, 7; lithic assemblages, 29–30; per- 10 sistent places, 53–55, 64; Upper Pleistocene, Howe, Bruce, 74 26, 34, 36–39, 41–42 hunter-gatherers, 9, 25, 41; complex, 11, 233; Lapham, Increase, Aztalan map by, 155, 157 in Levant, 235, 236–43; mobility strategies, LAS. See Limited Activity Sites 33–34; population estimates, 114–15; in San Late Glacial Maximum (LGM), 31, 34, 41 Francisco Bay area, 243–50 Late Natufian, 252; complex hunter-gatherers, hybridization, and Neanderthal extinction, 239–41, 242 40–41, 42, 43 Late period (San Francisco Bay area), burials, hybrids, hominin, 40–41 244, 247, 252 Hyer, Nathanial, 155; map of Aztalan, 156 Late Pithouse period (Mimbres Mogollon), 59, 60–61, 64 Iberian Peninsula, Neanderthals on, 42 Late Woodland Tradition, at Aztalan, 153, 154, Iberomaurusian industry, 8, 77, 86 164 identity, community, 55, 64 Leakey, Mary, on Oldowan tools, 72 Iran, Epipaleolithic in, 73–75 Levant, 76, 233; bifaces, 213–15; mortuary practices, 236–43 Jabrud Rockshelter, 3, 76 LGM. See Late Glacial Maximum Jafr Basin, 215, 222, 225 lightning, and arrows, 280 Jebel al-Hittiya ( Jordan), 218 Limited Activity Sites (LAS), 128, 129 Jebel Issa ( Jordan), 218 Lion Spring ( Jordan), bifacial cleavers from, Jelinek, Arthur J., 56 214, 215 Jordan, 22; cleavers in, 214–15, 216, 218; lithic assemblages/industries, 6, 42, 87, 141(n3); Epipaleolithic in, 75–76 in archaeological record, 28–29; in artifact scatters, 125, 127–28; controlled experimen- Kaibab chert, 129 tation with, 307–18; cortex ratios, 106–7; Kamminga, Johan, flake production experi- debitage studies, 126–27; Epipaleolithic, ments, 313 72–73, 78–86; mental template concept and, Kanzi, flake production by, 296–97 293–96; Middle and Upper Paleolithic, katsinas, knives with, 281 32–33(table), 36–38, 42; mobility and, Kehf al Hammar, 77 26–27, 29–30; New South Wales, 99, 101–4, kill-and-butcher sites, Azraq region, 222, 225 106–7, 114, 116; standardized production of, Kin Tiel, 201, 202, 207(n9) 293–94; in survey data, 136–38; transport knapping: controlled experiments in, 307–18; issues, 107–9; typologies, 71–72, 74–75, 76, mental and procedural templates of, 77; Upper Basin (Arizona), 130, 132–36; use 296–97 modification, 300–301 knife-wing symbols, 282 lithic scatters, artifact variability in, 135–38 knives: bifacial, 272–73; in Hopi symbolism, Livermore-Amador Valley, hunter-gatherer 280–81; Hopi and Zuni ritual use of, 281–82 mortuary practices, 244–50, 252

328 Index Lizard Man Village, 269, 276, 278; unusual Mindeleff, Cosmos, 193, 201 artifacts from, 270 Minoans, on Crete, 175, 176 Longacre, William A., 127 Mississippian Tradition, 9–10; and Aztalan, long-term occupation, at Mimbres sites, 59 165–66; migration and site structure, Lower Paleolithic, stone tools, 291 163–65; site distribution and structure, 156–58; town-and-mound communities Magician Burial, 272 in, 158–59 Magnani, Matthew, 314 Mitchell site, 57–58, 61 maps: of Aztalan, 155–56, 157, 160; of Casas mobility, mobility strategies, 7, 8–9, 41, 113; Grandes, 199–200; nineteenth-century, 9, lithics and, 26–27, 29–30, 37–39, 139–40; 153 logistical and residential, 33–34, 37–38, 42; Marks, Anthony, mental template, 293–94 provisioning locations and, 104–6; variable masonry ruins, artifact assemblages at, 135 and logistical, 30–31 Mattocks site, 56, 58, 62 Mogollon, 56. See also Mimbres Mogollon Maya, eccentrics, 272 Monks Mound (Cahokia), 159 McAnally site, 58, 62, 64 Monnier, Gilliane, on mental templates, 294 McPherron’s reduction model, 221–22 Montezuma site, 57–58, 60–61 Mellars, Paul, on mental templates, 294 Morocco, Epipaleolithic in, 76–77 mental template, 291–92; artifact modification mortuary practices, 11, 233; age-graded, 251–52; and, 300–301; history of term use, 292–96; Levant, 236–43; Livermore-Amador Valley, in knapping, 296–97; standardization and, 244–50; political complexity, 235–36; and 298–300; and symbolic artifacts, 297–98; socioeconomic changes, 234, 250–53 usefulness of, 301–2 mounds, at Aztalan, 165–66 Mesa Verde National Park, debitage and mound sites, eastern North American, 9–10, lithic artifact studies, 126–27 153, 155, 158, 161 Mesa Verde region, persistent places, 54 Myrtos-Pyrgos, 9, 176, 177; Cistern 2 at, microburins: Epipaleolithic, 72, 73, 75, 78; in 178–79, 181–83; feasting remains at, 184–87; microlith production, 80–84, 87 pottery assemblages at, 10, 175, 179–80 microliths: Epipaleolithic, 72–77, 79–86; from New South Wales, 102, 104, 110 Nadaouiyeh, 218 Middle-Late Transition period, grave goods, NAN Ranch Ruin, 54, 57; as persistent place, 247–50, 251 59–60, 64 Middle Mississippian period: at Aztalan, 153– Natufian, burials, 238–41 54; site distribution and structure, 156–58 Navajo, reuse of artifacts, 280, 284 Middle Paleolithic, lithic assemblages, 30, 36, Neanderthals, 7, 34, 36, 39; camps, 25–26; 37–38, 291, 295 hybridization, 40–41, 42, 43 Middle period (San Francisco Bay), 244, 251; Near East, 234; complex hunter-gatherer burials, 245–46 societies, 11, 235 Middle Pleistocene, Neanderthals, 25 Nebekian industry, 8, 76 middle-range theory, lithic diversity and, Nelson, Nels, 56 26–30 Neolithic period, Near East burials, 241–43, migration, and Mississippian site structure, 252 163–65 Neopalatial period, 187; Cistern 2 (Myrtos- Mimbres Foundation, survey, 56–57 Pyrgos) ceramics, 181, 184–85, 186 Mimbres Mogollon, 33; archaeological work New South Wales (NSW), 8, 100; artifact in, 56–57; climate and resource changes, assemblages, 99, 102–4, 106–8; ceremonial 64–65; economic and social interaction, 7–8; sites, 112–13; raw material sources, 105–6; persistent places, 57–62; upland sites, 62–63 risk management, 109–11; site types in, Mimbres Valley, 7, 57–59 111–12

Index 329 niche-construction theory, 5–6 platform mounds, at Mississippian sites, 165 Nowell, April, on mental template, 294 plaza, at Aztalan, 163 Nundooka (ND) study area, 100, 101 political complexity, and mortuary practices, 233, 235–36 obsidian, 129 Ponderosa Ranch site, 57, 63 Obregón, Father, 191, 192; on Paquimé, population models, Australian, 113–15 197–98 pottery, at Myrtos-Pyrgos, 10 occupation, 8; Mimbres Valley, 64–65 PPNA. See Pre-Pottery Neolithic A offerings, 276, 278 PPNB. See Pre-Pottery Neolithic B OIS. See Oxygen Isotope Stages Pre-Natufian, burials, 236 Old Town site, 57, 58 Pre-Pottery Neolithic A (PPNA), mortuary Old World typologies, 71 practices, 241–42 Olivella shell, in hunter-gatherer burials, 246, Pre-Pottery Neolithic B (PNB), 242 247, 248, 249 procedural templates, 297, 298 Opata, and Casas Grandes, 197–98, 207(n6) projectile points, 76, 110, 102, 136; ancestral optically stimulated luminescence (OSL), associations with, 282–83; Epipaleolithic, 101, 108 76, 77, 78, 80; ritually placed, 269, 270, 272, ornamentation. See adornment 273; from Sinagua sites, 276–78; on Zuni Osborne, Douglas, 126–27, 141(n2) fetishes, 280, 281 OSL. See optically stimulated luminescence Protopalatial period, 180, 187 Ouchtata points, 78 provisioning, locations, 104–6 Oxygen Isotope Stages (OIS), 23, 30, 31, 33; Pueblo Bonito, 274, 276 and lithic assemblages, 36, 37, 38 Pueblos (historic), religious forms in, 279 pueblos: lithic assemblages from, 126–27; Palegawra, 75 Mimbres, 56, 57, 61; 1000+ room, 201, 204 Paleolithic, 5. See also Lower Paleolithic; Middle Paleolithic; Upper Paleolithic Qalkhan points, 76, 78 palisaded mound sites, 9–10, 158; Aztalan as, quartz, in New South Wales assemblages, 102, 153, 155, 161 103–4 Paquimé. See Casas Grandes Pecos Pueblo, ritual stone artifacts in, 274, 278 radiocarbon dates: Australian heat-retainer Pecos Valley, 127 hearths, 8, 101–2, 108; Cahokia, 159; and Peery Lake (Rutherfords Creek), 100 population estimates, 114; Upper Basin Pelcin, Andrew, flake production experiments, (Ariz.) sites, 132 312–13 rainfall, in Australia, 101 persistent places, 7, 55; maintenance of, 64–65; raw material, 8, 127, 298; sources of, 105–6 Mimbres Mogollon, 53–54, 57–59; NAN religion, ethnographic analogy and, 279 Ranch as, 59–60 Reorganization period, 64 , as signs of the ancestors, 284 reproductive fitness, 110 pièces esquillées, 88; from Smugglers’ Cave, resharpening, of bifaces, 221–22 85–86 residences, lithic assemblages among, 128–29 piercers, from Tabun, 219 residential mobility, 29, 37 Pine Point-Langwell (Conservation, Karz), resources: Australian environments, 108, 100 110–11; Mimbres Valley, 64–65 Pirri points, 102, 104 retouched artifacts, 5, 34, 38; and residential Pithouse-period Mimbres, 53, 54, 57, 59–60 mobility, 29–30; temporal trends in, 35, 36; place, 7; power of, 283–84; reuse of, 111. See also US Southwest, 127–28, 135 persistent places retouched flakes, 102; from Tabun, 219–20; place making, 55–56 US Southwest, 127–28

330 Index Rezek, Zelko, flake production experiments, Sonora, coalescent sites in, 203 313–14 Southwest archaeology: debitage studies, 9, ring species, Neanderthal and AMH as, 39 125–27; ritual stone use in, 268–69, 271–78 risk: management of, 109–11; venture and Soyok ogres, knives used by, 280–81 contingency, 108 spatial distribution, of artifacts, 6–7 ritual objects, stone tools as, 11–12, 268–69, 270, Speth, John, flake formation experiments, 271–82 308–9 rituals, mortuary, 233 standardization, 298–99; and mental template, rock art, New South Wales, 112, 113 293–94 rockshelters, Neanderthal, 26 Stone Knife Society (Zuni), 282 Rudd Creek Pueblo, 278 stone tools, 7, 37; ancestral association of, 282–83; mobility and, 26–27; modification Sahl Umm Turifa, 218 of, 300–31; as ritual objects, 11–12, 268–69, salt erosion, at Casa Grande, 193, 195 270, 271–82; symbolic importance of, Sandy Creek site, 112 279–81; use-life of, 27–28 San Francisco Bay, hunter-gatherer societies, stone-tool typologies, 8; Epipaleolithic, 86–88 11, 233, 243–44 Stud Creek study area, 100, 101, 115 San Francisco phase, 60, 62 surface scatters, 125; variability in, 135–38 San Simon region, 54 survey, lithic assemblage data, 136–38 Sapillo Valley, 63 Swarts site, 56, 58 scrapers, 102, 110, 221 symbolism, 268; of grave goods, 271–72; and seed grinders, grinding dishes, 102, 103, 109 mental template of producing, 297–98 settlement pattern archaeology, 111 Syria, bifaces from, 218 settlement systems, 128, 197 sherd-and-lithic scatters, artifact variability Tabun Cave, 9; bifaces from, 10–11, 213, 215, 217, on, 137–39 219, 223–24, 225–26(n1); cleavers from, 214, shrines, stone tools at, 278 218, 220, 222 silcrete, in New South Wales assemblages, Taforalt, 77 102, 103–4 technology, 7, 8–9; mobility and, 139–40 Sinagua sites: projectile points from, 276–78; thermal-feature processing areas, 136–37 ritualized stone tool use at, 271, 272; Three Circle/Elk Ridge site, 58 unusual lithics from, 269, 270 Three Circle phase, 62; at Montezuma site, sites: artifact classes and, 135; feature syn- 60–61 chronicity, 159–60; type definitions, 111–12 Toltecs, 284 small tool tradition, Australian, 102, 104 tool types, Epipaleolithic, 8 Smith, Watson, “When is a Kiva?,” 268 Tor at-Tareeq, 75–76, 78, 82 Smugglers’ Cave, lithic assemblages, 76–77, 78, Tor Sageer, 75–76, 78, 82; microliths, 84 82, 85–86 trade and exchange systems, and site loca- social identity, and mortuary data, 254 tions, 113 social relationships, and mortuary rituals, transportation, of Australian Aboriginal 253–54 lithics, 107–9 social status, 12, 234, 272 tree-ring dates, Casas Grandes, 207–8(n11) social structure, and mortuary practices, 234 Tula, 284 social tradition, 295 typologies, 5; Bordes’ biface, 213–14, 217, 219; socioeconomic strategies, 11, 234; Early stone tool, 71–72, 74–75, 76, 77 Natufian, 238–39; mortuary practices and, 250–53; San Francisco Bay areas, 243–44, unifacial artifacts, in Upper Basin sites, 134, 247 135 sociopolitics, 113, 185, 208(n12) University of Michigan Mimbres Survey, 56

Index 331 Upper Basin (Ariz.), 129, 131, 139; lithic Western New South Wales Archaeology assemblages, 9, 126, 130, 132–36 Program (WNSWAP), 99 Upper Paleolithic: blade uniformity, 298–99; Wetherill Mesa, lithic studies, 126–27, 141(n2) lithic assemblages, 30, 36, 291; standardiza- Wheeler, Richard P., 127, 141(n2) tion of production, 294, 295 White, Peter, on mental templates, 294–95 Upper Pleistocene, 7, 25, 26; hominin interac- Whittaker, John, flake production experi- tion during, 39–40; mobility strategies, ments, 312 37–39 whole assemblage behavioral indicator use-life, of stone tools, 27–28 (WABI), 29 use-wear, biface, 221 Wisconsin, Late Woodland Tradition at, 164 Wisconsin Archaeological Survey, at Aztalan, Ventana Cave, projectile points in burials 161–63 at, 278 Wisconsin Historical Society, at Aztalan, von Humboldt, Alexander, on Aztlan, 155 161–63 WNSWAP. See Western New South Wales WABI. See whole assemblage behavioral Archaeology Program indicator Wynn, Thomas, on mental template, 295 Wadi al-Harad ( Jordan), bifaces for, 218 Wadi al-Hasa region, Epipaleolithic in, Y Bar site, 59–60, 64 75–76 warfare, ritual stone tool use and, 279 Zarzian, 8; lithic industry, 74–75, 82 Warwasi Rockshelter (Iran), 73–75; microbu- Zarzi Cave, 74 rins, 78 Zuni, 201, 284; fetishes, 280, 281; hero twins, Webster, Clement, 56 281–82

332 Index