Bioarchaeology and Identity Revisited 1683401530, 9781683401537

Choice Outstanding Academic Title This volume highlights new directions in the study of social identities in past popula

106 29

English Pages 244 [245] Year 2020

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Cover
BIOARCHAEOLOGY AND IDENTITY REVISITED
Title
Copyright
CONTENTS
List of Illustrations
Foreword
1. Identity Revisited: A Brief Introduction
2. Exploring Family, Ethnic, and Regional Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru
3. Bioarchaeology and the Narrative Construction of Tewa Identity
4. Negotiating Contact in the Periphery: Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia
5. Collective Bodies, Collective Identities: The Development of Identity in Bronze Age Cyprus
6. Death Ritual as a Social Strategy for Ancestral Affiliation: Constructing Identity and Persistent Place at Yoshigo Shell Mounds, Atsumi Peninsula, Japan
7. Identity and Health: Exploring Relationships among Health, Disease, and Identity in Past Populations
8. Intersectionality and the Multiplicity of Identities in the Andean Past
9. Exploring Identities in Forensic Biohistory
List of Contributors
Index
Recommend Papers

Bioarchaeology and Identity Revisited
 1683401530, 9781683401537

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Bioarchaeology and Identity Revisited Bioarchaeological Interpretations of the Human Past: Local, Regional, and Global Perspectives

Bioarchaeology i AND j

Identity Revisited EDITED BY

Kelly J. Knudson and Christopher M. Stojanowski Foreword by Clark Spencer Larsen

University of Florida Press Gainesville

Copyright 2020 by Kelly J. Knudson and Christopher M. Stojanowski All rights reserved Published in the United States of America. 25 24 23 22 21 20

6 5 4 3 2 1

Library of Congress Cataloging-in-Publication Data Names: Knudson, Kelly J., editor. | Stojanowski, Christopher M. (Christopher Michael), 1973– editor. Title: Bioarchaeology and identity revisited / edited by Kelly J. Knudson and Christopher M. Stojanowski. Other titles: Bioarchaeological interpretations of the human past. Description: Gainesville : University of Florida Press, 2020. | Series: Bioarchaeological interpretations of the human past: local, regional, and global perspectives | Includes bibliographical references and index. Identifiers: LCCN 2019054890 (print) | LCCN 2019054891 (ebook) | ISBN 9781683401537 (hardback) | ISBN 9781683401803 (pdf) Subjects: LCSH: Human remains (Archaeology) | Group identity. | Social archaeology. Classification: LCC CC79.5.H85 B495 2020 (print) | LCC CC79.5.H85 (ebook) | DDC 930.1—dc23 LC record available at https://lccn.loc.gov/2019054890 LC ebook record available at https://lccn.loc.gov/2019054891

University of Florida Press 2046 NE Waldo Road Suite 2100 Gainesville, FL 32609 http://upress.ufl.edu

Contents

List of Illustrations vii Foreword xi 1. Identity Revisited: A Brief Introduction 1 Christopher M. Stojanowski and Kelly J. Knudson

2. Exploring Family, Ethnic, and Regional Identities among TiwanakuAffiliated Communities in Moquegua, Peru 20 Kent M. Johnson

3. Bioarchaeology and the Narrative Construction of Tewa Identity 56 Scott G. Ortman

4. Negotiating Contact in the Periphery: Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia 85 Lesley A. Gregoricka

5. Collective Bodies, Collective Identities: The Development of Identity in Bronze Age Cyprus 107 Anna Osterholtz

6. Death Ritual as a Social Strategy for Ancestral Affiliation: Constructing Identity and Persistent Place at Yoshigo Shell Mounds, Atsumi Peninsula, Japan 136 Daniel H. Temple

7. Identity and Health: Exploring Relationships among Health, Disease, and Identity in Past Populations 163 Molly K. Zuckerman

8. Intersectionality and the Multiplicity of Identities in the Andean Past 185 Kelly J. Knudson, Christina Torres-Rouff, and Aliya R. Hoff

9. Exploring Identities in Forensic Biohistory 199 William N. Duncan and Christopher M. Stojanowski

List of Contributors 219 Index 221

Illustrations

Figures

2.1. Map of the middle Moquegua Valley, Peru 23 2.2. Unmodified cranium of specimen number M1–548 and frontooccipital cranial modification of specimen number M10 R-7 34 2.3. Genetic distance ordination for the study samples derived from principle components analysis of the scaled R-matrix 35 2.4. Scatter plot of the first and second canonical variates of cranial shape coded by ethnic affiliation and site 37 2.5. Graph theoretic layout of the inter-individual Mahalanobis matrix dichotomized at the 1st percentile 39 2.6. Combined ego network graphs of the three individuals with the highest degree centrality values 40 3.1. The study area in the U.S. Southwest and locations of regional samples 61 3.2. Principal Coordinates Analysis of the R matrix of the craniometric dataset 63 3.3. Residual Variance from Relethford-Blangero analysis of the craniometric dataset 64 3.4. Distribution of habitation sites, contemporary pueblo lands, and population strata for the Tewa Basin 69 3.5. Population history of the Tewa Basin 70 4.1. The Umm an-Nar–period tomb of Unar 2 (2300–2100 BCE) 90 4.2. Geologic map of southeastern Arabia 93

viii · Illustrations

4.3. Strontium and oxygen isotope ratios from Umm an-Nar–period human dental enamel across southeastern Arabia 95 4.4. Oxygen and carbon isotope values from individuals interred in six Umm an-Nar tombs from the United Arab Emirates 96 5.1. Map of Cyprus with sites marked 108 5.2. Migration/integration and colonization models of population movement 110 5.3. Boxplot of changes in dental wear through time 122 5.4. Examples of extramasticatory tooth use 123 5.5. Boxplot of overall area of cranial depression fractures (CDFs) through time 126 5.6. Boxplot of area of cranial depression fractures (CDFs) by sex 128 6.1. The location of Yoshigo Shell Mounds on the Atsumi peninsula 142 6.2. Spatial distribution of burials in the West, East, and South sections of the Yoshigo Shell Mounds cemetery 143 6.3. Examples of hip ornaments and animal implements (wild boar pendant) associated with Yoshigo Shell Mounds burials 145 6.4. Minimum genetic distances between cemetery sections based on eigenvectors of the R-matrix under the assumption of equal effective population sizes 149 8.1. Map of the San Pedro de Atacama oases, Chile 190 9.1. The Russian Imperial Family in 1913 200 9.2. The Fort King George calvaria 205 9.3. Memorial cairn built in tribute to victims of the Mountain Meadows Massacre, Utah; and newspaper depiction of the body of Mormon elder John D. Lee 207 9.4. Previous memorial for Richard III in Leicester cathedral; current tomb of Richard III; and a gallery from the Richard III visitor’s center in Leicester 209 Tables

2.1. Basicranial and temporal bone landmarks 27 2.2. Study samples by site 29 3.1. Summary of ancient mtDNA haplogroups for turkeys 67

Illustrations · ix

3.2. Results of Fisher’s exact tests (P-values) comparing haplogroup frequencies 67 4.1. Strontium, oxygen, and carbon isotope data 97 5.1. Sites examined and their associated time periods and calibrated radiocarbon dates 119 5.2. Linear enamel hypoplasia (LEH) frequency by site and time period 121 5.3. Cribra orbitalia (CO) and porotic hyperostosis (PH) frequencies 124 5.4. All cranial depression fractures (CDFs) recorded during analysis 127 6.1. Sample composition for the biodistance analysis 147 6.2. Residuals of observed relative to expected variation calculated for each sample 148 9.1. Examples of forensic biohistory 201

Foreword

I well recall reading the draft of Kelly Knudson and Christopher Stojanowski’s Bioarchaeology and Identity in the Americas (2009) submitted for publication in the Bioarchaeological Interpretations of the Human Past book series. Their book was just the second to appear in the series, and it clearly anticipated the expansion and maturation of the field of bioarchaeology to include new ideas and new ways of thinking about the human past. Their book was critical for encouraging the development of research in and contributions to a growing literature having a wider global reach. The following pages of the present volume present a further treatment of social identity building on the earlier volume. This new addition to the series makes very clear that biological outcomes are often socially mediated and lend themselves to addressing key questions about identity construction and social behavior in general. Bioarchaeology and Identity Revisited reveals considerable maturation in the study of social identity, underscoring the central role of human remains from archaeological settings for addressing fundamental issues and problems. Importantly, the contributors to the volume provide insights about the human past and the importance of individual and population identity using new tools for addressing newly emerging topics. Knudson and Stojanowski’s earlier volume focused on moving “beyond the sense of identity as biology, identity as origins and identity as positive identification.” This new work reveals the continued and growing interest in the topic. I am impressed by the implications of the research involving themes common to the social and behavioral sciences focusing exclusively on contemporary societies. Other disciplines are committed to the analysis of social behavior and outcomes but have remarkably little or no temporal context, certainly in comparison to the decades and centuries of context

xii · Foreword

in bioarchaeology. In this regard, the case studies in this book present new ways of documenting and interpreting social behavior in general and identity. For example, theoretical frameworks pertaining to structural violence, the context of social hierarchy, and reduced access to resources for some members of society were developed based on the study of living populations. The application of research frameworks in the general field of social contexts to the study of the remains of the dead offers new opportunities for developing an understanding of the historical depth and the long-term circumstances that portray health and well-being. One such opportunity is the record of individuals being exploited via colonization and the negative circumstances that pertain to the human past and continue with numerous descendant populations. The poor health outcomes deriving from labor exploitation, reduced access to quality nutrition, and the general deterioration of living conditions reveal a general pattern of increased physiological stress owing to the downstream effects of the exploitation, goals, and intentions of a colonial or otherwise dominant power. Clearly, investigation of the social determinants of health and its hierarchical organization is proving important for our growing appreciation of the role of social organization and hierarchical structures on health outcomes. The contributors to the book make clear that bioarchaeology offers opportunities to access the reservoir of life experiences presented in skeletal remains pertaining to stress and illness, deprivation and disease, and the circumstances that play out in growth and development of the skeleton and dentition owing to reduced quality of resources that are required for normal growth and development. For all stages of life, bioarchaeology provides clear confirmation that one’s biological and social identity are exhibited over the life history of a person. Largely trained as anthropologists, bioarchaeologists have developed an informed appreciation for the consequences and impacts of the lived experience (Larsen 2015). And, perhaps nowhere else in the archaeological record do we see the effects of negative circumstances and the dramatic effect these experiences have on health and well-being. We are reminded of the highly deleterious health outcomes of living in marginal circumstances via the investigation of both archaeological skeletal series and other composite samples, such as that so clearly illustrated in the reference collections amassed in the twentieth century (e.g., the Smithsonian Institution’s Robert J. Terry Anatomical Skeletal Collection

Foreword · xiii

and the Case Western Reserve University’s Hamann-Todd Osteological Collection). These important collections, fundamental to research on age estimation, sex identification, and numerous other topics in bioarchaeology, skeletal biology, and forensic science, exhibit abundant evidence of high levels of disease and stress largely derived from the persons in life having limited access to essential nutrition, health care, and conditions that promote good health. Importantly, the contributors to this volume provide a powerful composite picture derived from the study of the remains of past people contextualized by a wide range of knowledge derived from the social, behavioral, and natural sciences. The message that bioarchaeology brings to the table is a powerful one—namely, that social identity is a complex composite of lived experiences pertaining to access to food and other resources, age, sex, and perception of place in society by its members. Having read the research findings presented in the following pages, I further appreciate the growing understanding that identity—self, community, nation, and world—is central to understanding conditions that drive the human experience, past, present, and future. Clark Spencer Larsen Series Editor Works Cited Knudson, Kelly J., and Christopher M. Stojanowski, editors. 2009. Bioarchaeology and Identity in the Americas. Gainesville: University Press of Florida. Larsen, Clark S. 2015. Bioarchaeology: Interpreting Behavior from the Human Skeleton, 2nd ed. Cambridge: Cambridge University Press.

1 Identity Revisited A Brief Introduction Christopher M. Stojanowski and Kelly J. Knudson

The history of bioarchaeological inquiry and the evolving nature of the field is the subject of continued summarization and retrospection (Baker and Agarwal 2017; Larsen 2018; Rakita 2014), with a number of complementary research foci emerging over the last two decades. During the course of our graduate training, bioarchaeology was principally associated with the three D’s—diet, disease, and demography—as part of a biocultural focus on human adaptation, archaeological interest in major lifestyle transitions, and research surrounding the Columbian centennial in 1992. Although certainly not the solitary pursuit of bioarchaeology at the time, a diversification occurred in the field around the turn of the millennium. Initially reflected in studies of body modification and mortuary analysis, and later expanding to a more general interest in the “archaeological body”(Sofaer 2005), gender (Lucy 1997; Sofaer Deverenski 1997; Spielmann 1995), and age (Gilchrist 2000), the field of bioarchaeology experienced rapid growth with the emergence of more social theoretical framings and recognition of the more humanistic aspects of our work alongside continued interest in health, disease, and biocultural questions (Stojanowski and Duncan 2015). When we initially published Bioarchaeology and Identity in the Americas (2009), bioarchaeological studies that interrogated social identities or “identity” in the past were comparatively rare. However, a number of efforts were published within roughly the same five-year period (circa 2006–2010), reflective of the growing interest in social identity research in past populations (e.g., chapters in Gowland and Knüsel 2006; Agarwal and Glencross 2011; see also Geller 2006, 2008, 2009a, 2009b; Halcrow and Tayles 2008) and the embracing of broader

2 · Christopher M. Stojanowski and Kelly J. Knudson

areas of social theory as relevant to the nuanced interpretation of the biological signatures observed within the bodies of those that lived in the past (the “archaeological body”). In attempting to define what social identities were and are, we wrote, “we define identities research as not about who people were or where they or their ancestors came from, but who they thought they were, how they advertised this identity to others, how others perceived it, and the resulting repercussions of this matrix of inter-personal and inter-societal relationships”(Knudson and Stojanowski 2009: 5). In other words, our interest in identity studies emerged from the social sciences, with the goal of ascertaining more emic aspects of social experience in the past. That is, we specifically wanted to move beyond the sense of identity as biology, identity as origins, and identity as positive identification. The 2009 volume was divided into two sections: community identities and individual-level identities. Community identity referred to aspects of group affiliation with an emphasis on ethnogenesis and community building. Research on ethnogenesis has continued (Klaus 2008; Kurin 2012; Ortman 2012; Velasco 2018), though not without criticism (Voss 2015); however, interest in community identity has received considerably more attention (e.g., papers in Juengst and Becker 2017). Individual identity referred to aspects of ensoulment and personhood as well as personal aspects of self, such as age, gender, and ethnicity—topics with considerable current interest (Cerezo-Román 2015; Gowland 2017; Zuckerman and Crandall 2019). In retrospect, this dual structure was too simplistic and drew an artificial divide between categories of identity expression that are quite fluid, overlapping, and ephemeral. Biodistance, mortuary analysis, body modification, and isotopic signatures were the principal data sources, with many of the chapters focusing on Andean or Mayan contexts where preservation (Andes) and rich historical and archaeological records (Andes and Maya region) helped frame and embellish the bioarchaeological interpretations. The solitary North American chapter focused on recent historical events (colonial La Florida), where written documents also provided key information to inform the interpretation. The reader of the 2009 volume may have been left wondering if it is possible to study identity in more ancient contexts, something the current volume hopes to address. Since our 2009 publication, studies of social identities in bioarchaeological contexts have increased dramatically and expanded in scope. In fact, we would argue that the emergence of identity studies within

Identity Revisited: A Brief Introduction · 3

bioarchaeology helped open the field to a variety of different framings as scholars explored topics such as the bioarchaeology of marginalized peoples (Mant and Holland 2019), slavery, captivity and exploitation (Harrod and Martin 2015), elder abuse (Gowland 2006), postmortem agency (Crandall and Martin 2014), and quotidian experience (Schrader 2013, 2018). Debra Martin’s new series with Springer (2017–present) is entitled “Bioarchaeology and Social Theory” with volumes focusing on themes of social injustice and structural violence (Nystrom 2017), care (Tilley and Schrenk 2017), women and childhood (Martin and Tegtmeyer 2017), impairment and disability (Byrnes and Muller 2017), and social control (Harrod 2017). Topics that had previously received chapter-length discussion have now been the subject of entire book–length treatments, such as gender (Agarwal and Wesp 2017; Geller 2017) and body modification (Tiesler 2014). However, as the ten-year anniversary of the first volume approached, we wondered how the 2009 volume had helped define the bioarchaeology of identity as a novel and emerging research focus. We had both taught a graduate-level seminar on the Bioarchaeology of Identity a number of times and had witnessed the impact that class had on graduate student framing of their research projects. That said, the impact of the volume on the field as a whole was more difficult to gauge; citation data are modest. However, in 2017, Lynne Schepartz published a joint book review entitled, “The Body of Power, the Body of Memory” that includes a review of our 2009 volume. To our surprise (and delight), she described the initial volume as a “tour-de-force . . . [that] deserves to be recognized as a classic text in bioarchaeology and skeletal biology” (Schepartz 2017: 39, 41). With such a compliment, with the ten-year anniversary approaching, and with an additional decade of experience teaching and thinking about identity research, we decided to revisit the topic with an expanded and more comprehensive scope. Before introducing the chapters in this volume and their contributions to the field as a whole, we first briefly discuss some of the key trends in identities research in bioarchaeology. Recent and Future Directions in Identities Research

In the past ten years, bioarchaeological research on social identities has dramatically increased. The growing number of publications that grapple with specific social identities, such as gender identities, as well as

4 · Christopher M. Stojanowski and Kelly J. Knudson

the implications of different social identities in the past and present, are indicators of the general health and dynamism of bioarchaeology as a field. Here, we briefly review recent trends in research on identities in bioarchaeology. Concepts of ethnic identity in the past continue to derive from the work of Barth (1969) and other scholars of the Manchester School (Cohen 1969, 1974; Gluckman 1958). Following Barth (1969), most scholars today continue to emphasize the situational and mutable aspects of ethnicity, rather than a static and fixed “tribe.” However, some scholars in bioarchaeology and archaeology are arguing for new approaches to ethnicity and ethnic identity (e.g., Baitzel and Goldstein 2011; Mac Sweeney 2009; Stovel 2013), while others are focusing more broadly on group or community identities in the past (Jenks 2013; Lorentz 2017; Sheridan et al. 2014; Stojanowski, Carver, and Miller 2014; Tiesler and Lacadena 2018). In addition, a growing number of scholars are addressing specific social identities, such as gender identity or community identity. Of these, investigations into gender identities have the longest history of research and arguably the largest number of investigations. One of the most thoughtprovoking areas of research is that on bioarchaeology and sexuality, often using explicitly queer and/or feminist perspectives (Geller 2008, 2017; Voss 2007). Importantly, scholars are also interrogating constructions of masculinity and male gendered identities (Torres-Rouff 2011) and male activities (Knüsel 2011), in contrast to earlier archaeological and bioarchaeological research on gender that focused almost exclusively on women and females in the past. The implications of gender identities are also a growing area of research. For example, theoretically sophisticated research on the implications for health and disease (Mays et al. 2018; Western and Bekvalac 2017; Zuckerman 2017) and on the evidence of violence toward biological females and individuals gendered as female (Kurin 2016; Novak 2009; Redfern 2017; Tung 2012) is an important area for current and future research. In addition, researchers continue to investigate dietary variability that is sexbased (Coltrain and Janetski 2013; Dong et al. 2017; Eerkens and Bartelink 2013; Pearson et al. 2013; Reitsema and Vercellotti 2012; Somerville et al. 2015) and gendered differences in mobility (Bentley et al. 2007; Watson and Stoll 2013). While these areas of inquiry have a longer history than others discussed here, we see the increasing number and theoretical sophistication of these studies as a positive development for the field.

Identity Revisited: A Brief Introduction · 5

In contrast, investigations of age identities, including childhood, do not have the long history of research that gender identities have. However, there are a growing number of publications on childhood in the past (Alfonso-Durruty, Thompson, and Crandall 2014; Baitzel and Goldstein 2014; Barrett 2014; Bickle and Fibiger 2014; Blom and Knudson 2014; Cook, Thompson, and Alfonso-Durruty 2014; Crandall and Thompson 2014; Ellis 2014; Geber 2014; Halcrow and Tayles 2011; Holt et al. 2014; Martin, Thompson, and Crandall 2014; Palkovich 2014; Reynard and Tuross 2015; Schmidt, Kwok, and Keenleyside 2016; Wright 2013a, 2013b), as well as other age identities, such as older age (Baitzel and Goldstein 2016; Cave and Oxenham 2016; Gowland 2016). Though publications on this topic are rarer, violence against individuals with different age identities is also an important line of inquiry (Gowland 2016; Kamp 2009; Mays 2014; Timmins, Seréville-Niel, and Brickley 2017; Tung and Knudson 2010; Tung et al. 2015; Verlinden and Lewis 2015; Wheeler et al. 2013). In addition to investigating specific social identities, such as ethnic, gender, or age identities, one exciting area of growth for the field is the use of intersectional frameworks. Although less common than investigations into specific social identities, scholars are beginning to investigate concepts of intersectionality in the archaeological record. Following the work of Crenshaw (1991), scholars are examining intersecting identities (Torres-Rouff and Knudson 2017) and using multiple lines of evidence to examine identity (Goldstein 2013). Finally, one of the most fruitful directions for current and future work, we feel, is the recognition of the impacts of different social identities on how research is conducted, both in the past and in the present. Recent research on structural violence and particularly on bioarchaeological research on documented collections is an important example of this work (Geller and Stojanowski 2017; Klaus 2012; Nystrom 2014; Stone 2018; Zuckerman, Kamnikar, and Mathena 2014). We are also very excited about the important research on ethics in bioarchaeology, much of which involves considering how social identities such as gender, class, or ethnic identity impact both researchers and descendent groups (Boutin et al. 2017; Joyce 2017; Larsen 2018; Lewis and Tung 2013; Martin and Harrod 2014; Stojanowski and Duncan 2015; Turner and Andrushko 2011; Turner, Bernstein et al. 2018; Turner, Wagner, and Cabana 2018; Zuckerman et al. 2014).

6 · Christopher M. Stojanowski and Kelly J. Knudson

Contributions to Bioarchaeology and Identity Revisited

Contributions to Bioarchaeology and Identity Revisited were chosen to highlight new directions in bioarchaeological research in social identities. While Bioarchaeology and Identity in the Americas was confined to research in the New World (Knudson and Stojanowski 2009), the chapters here include case studies from both North America and South America, as well as Asia, Europe, and the Middle East; most contributors to this volume did not contribute chapters to Bioarchaeology and Identity in the Americas. Topically, these chapters address social identities at a variety of scales and include successful investigations of complex topics such as cultural memory, identity construction in a variety of geographic and social contexts, and intersectionality in past populations. Each chapter is briefly summarized below. In chapter 2, “Exploring Family, Ethnic, and Regional Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru,” Kent M. Johnson uses both biodistance analysis and cranial modification styles to better understand identity formation in the South Central Andes. Johnson’s treatment is admirable in that he uses a regional, multiscalar approach that incorporates mortuary populations not included in previous work in the region. He also very effectively includes a network analysis in conjunction with his phenotypic data to better understand how family networks were used to forge a collective “Moquegua Tiwanaku” identity among individuals from different ethnic backgrounds. Johnson finds support for the dual diaspora model of ethnic identity in that phenotypic data suggested that individuals did tend to marry within their own ethnic community. However, this was not always the case, and some analyses did suggest marriage occurred across these ethnic boundaries. Further, these practices tended to vary by ethnic affiliation (Omo-style vs. Chen Chen style), with some Omo-style samples demonstrating evidence for greater exogamy than others. These results are cast against a backdrop of relative uniformity in cranial modification practices, and the complex backdrop of variation in material culture leading Johnson to conclude that the inference of ethnic distinctions was correct, but also perhaps blurring in the context of longer-term changes associated with possible ethnogenesis. Johnson’s chapter demonstrates the power of biological distance analysis for asserting patterns of phenotypic variation into existing frameworks of

Identity Revisited: A Brief Introduction · 7

archaeological ethnic-based data sets. The use of network analysis is novel within biodistance and offers exciting possibilities for future research. In chapter 3, “Bioarchaeology and the Narrative Construction of Tewa Identity,” Scott G. Ortman addresses identity construction in the southwestern United States. Like Johnson, Ortman examines regional-level identity formation at multiple scales. Ortman integrates bioarchaeological data, particularly biodistance data that can be used to infer genetic relationships between humans, with ancient DNA data from domesticated animals to infer past migration events that led to Tewa ethnogenesis in New Mexico. Using these data, he reconstructs a likely population history of the Tewa Basin based on known habitation sites and settlement data and finds that Tewa ethnogenesis likely involved a large number of immigrants from the Mesa Verde region combining with a smaller group of existing residents in the region. Rather than simplistically using the bioarchaeological and archaeological data to “test” oral histories of Tewa ethnogenesis, Ortman instead compares the data sets to identify the ways in which they differ. Bioarchaeological and genetic approaches provide etic perspectives that suggest Tewa ethnogenesis resulted from a much larger immigrant population swamping a small, local population. Oral histories provide emic constructions of this same process, which suggest Tewa society results from a single population that had temporarily split and later rejoined in the Rio Grande region. The differences, Ortman argues, reveal important aspects of history building through the narrative construction of origin stories that have specific functions within the sociopolitics of past and present Tewa society. In chapter 4, “Negotiating Contact in the Periphery: Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia,” Lesley Gregoricka uses bioarchaeological and biogeochemical data to infer aspects of identity performance in Bronze Age Arabian populations. One important contribution of her chapter is the use of commingled and often fragmentary human remains. Her analyses demonstrate that one does not need intentional burials of single individuals to examine identity construction in the past. And, in fact, one of her primary insights is that the creation of the commingled deposit through agential mortuary behaviors is particularly telling of the dynamics of identity expression by the local population. For example, using isotopic analyses, she identifies nonlocal individuals (perhaps high-ranking merchants or diplomats) that

8 · Christopher M. Stojanowski and Kelly J. Knudson

are interpreted as being brought into the fold of the local community’s practices by virtue of their presence in the intentionally commingled deposits. Further, she argues that local communities were not subsumed by powerful neighbors in Mesopotamia and the Indus Valley, but rather continued their own mortuary traditions and maintained autonomy even as they interacted with other centers. The focus on collective funerary practices may be seen as a form of resistance against the growing inequalities manifest within the local community by virtue of the expanding trade and exchange dynamics within the broader Arabian region. That is, local power brokers may have accumulated status and prestige in life, but the assertion of communal burial and the separation of prestige burial goods from their owners brought about by the interment and commingling process served to resist social hierarchies in what was a tribal, kin-based society. In chapter 5, “Collective Bodies, Collective Identities: The Development of Identity in Bronze-Age Cyprus,” Anna Osterholtz uses bioarchaeological data, particularly health status and evidence for trauma, in mortuary populations dating to two different periods within the Bronze Age on the Mediterranean island of Cyprus. By presenting her bioarchaeological data in the context of the archaeological and historical evidence of major population size increases during these time periods, she evaluates two different models for the development of a Cypriot identity. Specifically, Osterholtz considers how health and trauma experience would change differentially based on two models of population size increase. In the first, called the colonization model, Cypriot ethnogenesis involves poor health outcomes and increasing rates of violence. In the second, called the hybridization model based on the third-space phenomena, Cypriot ethnogenesis would involve no changes in health experience and trauma rates. These models and their expectations are based on assumptions about how directional sociopolitical changes are during times of ethnic transformation. In the colonization model, more directed, unidirectional changes are seen as disruptive mechanisms with health outcomes. In the hybridization model, interactions are more bidirectional and integrative, thus leaving in place sociopolitical factors that buffer populations from periods of stress. Similar to Gregoricka, Osterholtz used commingled skeletal assemblages to effectively document changes in health and trauma experience. Her work concludes that little changes in osteological indicators of morbidity could be documented in the skeletal series, suggesting Cypriot

Identity Revisited: A Brief Introduction · 9

ethnogenesis occurred through a hybridization process with limited negative consequences on the indigenous populations of the island. In chapter 6, “Death Ritual as a Social Strategy for Ancestral Affiliation: Constructing Identity and Persistent Place at Yoshigo Shell Mounds, Atsumi Peninsula, Japan,” Daniel H. Temple examines community identities in hunter-gatherer populations during the Jomon period. Temple uses concepts of persistent place and social memory to examine the role of the ancestor in the definition and maintenance of Jomon community identity. He combines analyses of grave good allocation and spatial orientation with biodistance analysis of individuals buried within different sections of the Yoshigo shell mounds cemetery, representing two temporal phases of occupation. His work documents patterns of both continuity and change in mortuary practices. However, when combined with biodistance data on phenotypic variance and spatial organization of the graves, he is able to show how one specific sector of the cemetery demonstrates evidence for long-term use by a geographically diverse yet biologically homogenous population of hunter-gatherers. The physical proximity of these later graves to those of the earlier occupation and the intentional placement of similar grave goods with the interments suggests an emergent focus on the intentional construction of identity such that late-phase individuals affiliated themselves with earlier populations through both mortuary ritual and spatial distribution. Identity manifests through practices maintained in the face of changing natural and social environments as an intentional reflection of persistence and connection to the landscape. In chapter 7, “Identity and Health: Exploring Relationships among Health, Disease, and Identity in Past Populations,” Molly K. Zuckerman shifts the focus of the volume from the development of group identities to the complex relationships among identities, health, and disease. Using concepts of intersectionality, Zuckerman examines the roles of stress exposure and social identities in persistent infection with syphilis. Zuckerman first provides a thorough and thought-provoking overview of discrimination and health and disease in the social epidemiology and clinical medicine literature as well as on recent bioarchaeological research on frailty and structural and interpersonal discrimination. Her analysis of syphilis in two London cemeteries shows that individuals with skeletal evidence of chronic stress exposure during growth and development and during adulthood were associated with greater frailty relative to late stage syphilis. She points out the intersections between visible infection, declining

10 · Christopher M. Stojanowski and Kelly J. Knudson

socioeconomic status, expensive medical treatments, and chronic stress exposure, and effectively argues for the importance of contextualized bioarchaeological research for both past and present conceptions of health and diseases. In chapter 8, “Intersectionality and the Multiplicity of Identities in the Andean Past,” Kelly J. Knudson, Christina Torres-Rouff, and Aliya R. Hoff use a variety of lines of bioarchaeological and archaeological evidence to examine the complex ways in which different social identities manifested in the San Pedro de Atacama oases of northern Chile around AD 1000. Following the theoretical approaches of scholars and activists outside of anthropology, they use the concept of intersectionality to better understand the ways in which different social identities impacted and were impacted by each other. In doing this, they present detailed life histories of a small number of individuals buried in the oases, focusing on how social identities manifested and on how the power differentials that are an integral aspect of intersectional frameworks can be seen in past populations. Finally, in chapter 9, “Exploring Identities in Forensic Biohistory,” William N. Duncan and Christopher M. Stojanowski delve into the ethical complexities of research on social identities in the past. Using their work on the skeletal remains from the Fort King George Museum in Georgia, United States, which had long been thought to belong to a martyred sixteenth-century priest, Duncan and Stojanowski examine the many levels of “identity” in forensic biohistory. In forensic biohistory, which involves the analysis of famous dead in historical contexts, much research on “identity” involves positively identifying the dead, and, once there is a named body, identifying the circumstances surrounding the death. Given this focus on named individuals, forensic biohistory involves a series of ethical dilemmas regarding the roles and responsibilities of researchers, particularly given the often high-profile cases. Rather than a simplistic litany of “best practices,” Duncan and Stojanowski provide a series of difficult questions for individual researchers, and the field, to address as we move forward. Conclusion

Recent research on social identities in the past reflects the growing theoretical and methodological sophistication of bioarchaeology as a field. Exciting new work on the complex intersections among, and implications

Identity Revisited: A Brief Introduction · 11

of, multiple social identities provides new information about past lives as well as important directions for future scholarship and practices in our field. Bioarchaeology and Identity Revisited contributors have built on the growth of these important lines of inquiry and have, we hope, provided models for future research on these important topics. Acknowledgments

First, we would like to thank the participants in this edited volume for their excellent contributions, which move the field in exciting new directions and stimulate our own investigations of social identities in the past. We would also like to thank the students and faculty at Arizona State University, particularly in the Center for Bioarchaeological Research in the School of Human Evolution and Social Change, for their thoughtprovoking scholarship. Graduate students in our seminars on bioarchaeology and social identities have been a particularly welcome addition to our discussions on these important topics. Works Cited Agarwal, Sabrina C., and Bonnie A. Glencross. 2011. Social Bioarchaeology. Chichester: John Wiley & Sons. Agarwal, Sabrina C., and Julie K. Wesp. 2017. Exploring Sex and Gender in Bioarchaeology. Albuquerque: University of New Mexico Press. Alfonso-Durruty, Marta, Jennifer L. Thompson, and John J. Crandall. 2014. “Conclusion. Little Bodies, Big Voices: The Lives of Children in the Past.” In Tracing Childhood: Bioarchaeological Investigations of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta Alfonso-Durruty, and John J. Crandall, 246–258. Gainesville: University Press of Florida. Baitzel, Sarah I., and Paul S. Goldstein. 2011. “Manifesting Ethnic Identity in an Ancient Society: Evidence from a Tiwanaku Cemetery in Moquegua, Peru.” In Ethnicity from Various Angles and Through Varied Lenses: Yesterday’s Today in Latin America, edited by Christine Hunefeldt and Leon Zamosc, 30–44. Portland, OR: Sussex Academic Press. Baitzel, Sarah I., and Paul S. Goldstein. 2014. “More than the Sum of Its Parts: Dress and Social Identity in a Provincial Tiwanaku Child Burial.” Journal of Anthropological Archaeology 35: 51–62. Baitzel, Sarah I., and Paul S. Goldstein. 2016. “No Country for Old People: A Paleodemographic Analysis of Migration Dynamics in Early Andean States.” International Journal of Osteoarchaeology 26(6): 1001–1013.

12 · Christopher M. Stojanowski and Kelly J. Knudson

Baker, Brenda J., and Sabrina C. Agarwal. “Stronger Together: Advancing a Global Bioarchaeology.” Bioarchaeology International 1(1–2): 1–18. Barrett, Autumn R. 2014. “Childhood, Colonialism, and Nation-Building: Child Labor in Virginia and New York.” In Tracing Childhood: Bioarchaeological Investigations of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta Alfonso-Durruty, and John J. Crandall, 159–182. Gainesville: University Press of Florida. Barth, Fredrik. 1969. Ethnic Groups and Boundaries: The Social Organization of Culture Difference. Prospect Heights, IL: Waveland Press. Bentley, R. Alexander, Nancy Tayles, Charles Higham, Colin Macpherson, and Tim C. Atkinson. 2007. “Shifting Gender Relations at Khok Phanom Di, Thailand: Isotopic Evidence from the Skeletons.” Current Anthropology 48(2): 301–314. Bickle, Penny, and Linda Fibiger. 2014. “Ageing, Childhood and Social Identity in the Early Neolithic of Central Europe.” European Journal of Archaeology 17(2): 208–228. Blom, Deborah E., and Kelly J. Knudson. 2014. “Tracing Tiwanaku Childhoods: A Bioarchaeological Study of Age and Identity in Tiwanaku Society.” In Tracing Childhood: Bioarchaeological Investigations of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta Alfonso-Durruty, and John J. Crandall, 228–245. Gainesville: University Press of Florida. Boutin, Alexis T., Madison Long, Rudy A. Dinarte, and Erica R. Thompson. 2017. “Building a Better Bioarchaeology through Community Collaboration.” Bioarchaeology International 1(3–4): 191–204. Byrnes, Jennifer F., and Jennifer Muller. 2017. Bioarchaeology of Impairment and Disability. New York: Springer. Cave, C., and M. Oxenham. 2016. “Identification of the Archaeological ‘Invisible Elderly’: An Approach Illustrated with an Anglo-Saxon Example.” International Journal of Osteoarchaeology 26(1): 163–175. Cerezo-Román, Jessica I. 2015. “Unpacking Personhood and Funerary Customs in the Hohokam Area of Southern Arizona.” American Antiquity 80(2): 353–375. Cohen, Abner. 1969. Custom and Politics in Urban Africa: A Study of Hausa Migrants in Yoruba Towns. London: Routledge. Cohen, Abner, ed. 1974. Urban Ethnicity. London: Tavistock Publications. Coltrain, Joan Brenner, and Joel C. Janetski. 2013. “The Stable and Radio-Isotope Chemistry of Southeastern Utah Basketmaker II Burials: Dietary Analysis Using the Linear Mixing Model SISUS, Age and Sex Patterning, Geolocation and Temporal Patterning.” Journal of Archaeological Sciences 40(40): 4711–4730. Cook, Della C., Andrew R. Thompson, and Marta Alfonso-Durruty. 2014. “Death and the Special Child: Three Examples from the Ancient Midwest.” In Tracing Childhood: Bioarchaeological Investigations of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta Alfonso-Durruty, and John J. Crandall, 17–35. Gainesville: University Press of Florida. Crandall, John J., and Jennifer L. Thompson. 2014. “Beyond Victims: Exploring the Identity of Sacrificed Infants and Children at La Cueva de los Muertos Chiquitos, Durango, Mexico (AD 571–1168).” In Tracing Childhood: Bioarchaeological Investigations

Identity Revisited: A Brief Introduction · 13

of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta Alfonso-Durruty, and John J. Crandall, 36–57. Gainesville: University Press of Florida. Crandall, John J., and Debra L. Martin. 2014. “The Bioarchaeology of Postmortem Agency: Integrating Archaeological Theory with Human Skeletal Remains.” Cambridge Archaeological Journal 24(3): 429–435. Crenshaw, Kimberlé. 1991. “Mapping the Margins: Intersectionality, Identity Politics, and Violence against Women of Color.” Stanford Law Review 43(6): 1241–1299. Dong, Yu, Chelsea Morgan, Yurii Chinenov, Ligang Zhou, Wenquan Fan, Xiaolin Ma, and Kate Pechenkina. 2017. “Shifting Diets and the Rise of Male-Biased Inequality on the Central Plains of China during Eastern Zhou.” Proceedings of the National Academy of Sciences USA 114(5): 932–937. Eerkens, Jelmer W., and Eric J. Bartelink. 2013. “Sex-Based Weaning and Early Childhood Diet among Middle Holocene Hunter-Gatherers in Central California.” American Journal of Physical Anthropology 152(4): 471–483. Ellis, Meredith A.B. 2014. “A Disciplined Childhood in Nineteenth-Century New York City: A Social Bioarchaeology of the Subadults of the Spring Street Presbyterian Church.” In Tracing Childhood: Bioarchaeological Investigations of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta Alfonso-Durruty, and John J. Crandall, 139–158. Gainesville: University Press of Florida. Geber, Jonny. 2014. “Skeletal Manifestations of Stress in Child Victims of the Great Irish Famine (1845–1852): Prevalence of Enamel Hypoplasia, Harris Lines, and Growth Retardation.” American Journal of Physical Anthropology 155(1): 149–161. Geller, Pamela L. 2006. “Altering Identities: Body Modification and the Pre-Columbian Maya.” In Social Archaeology of Funeral Remains, edited by Rebecca Gowland and Christopher Knüsel, 279–291. Oxford: Oxbow Books. Geller, Pamela L. 2008. “Conceiving Sex: Fomenting a Feminist Bioarchaeology.” Journal of Social Archaeology 8(1): 113–138. Geller, Pamela L. 2009a. “Bodyscapes, Biology, and Heteronormativity.” American Anthropologist 111(4): 504–516. Geller, Pamela L. 2009b. “Identity and Difference: Complicating Gender in Archaeology.” Annual Review of Anthropology 38: 65–81. Geller, Pamela L. 2017. The Bioarchaeology of Socio-Sexual Lives: Queering Common Sense about Sex, Gender, and Sexuality. New York: Springer. Geller, Pamela L., and Christopher M. Stojanowski. 2017. “The Vanishing Black Indian: Revisiting Craniometry and Historic Collections.” American Journal of Physical Anthropology 162(2): 267–284. Gilchrist, Roberta. 2000. “Archaeological Biographies: Realizing Human Lifecycles, -courses and -histories.” World Archaeology 31(3): 325–328. Gluckman, Max. 1958. Analysis of a Social Situation in Modern Zululand. Manchester: Manchester University Press. Goldstein, Lynne. 2013. “Negotiating the Gateway: Working with Multiple Lines of Evidence to Determine Identity.” In The Dead Tell Tales: Essays in Honor of Jane E. Buikstra, edited by María Cecilia Lozada and Barra O’Donnabhain, 33–42. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press.

14 · Christopher M. Stojanowski and Kelly J. Knudson

Gowland, Rebecca L. 2016. “Elder Abuse: Evaluating the Potentials and Problems of Diagnosis in the Archaeological Record.” International Journal of Osteoarchaeology 26(3): 514–523. Gowland, Rebecca L. 2017. “Embodied Identities in Roman Britain: A Bioarchaeological Approach.” Britannia 48: 177–194. Gowland, Rebecca L., and Christopher Knüsel. 2006. Social Archaeology of Funerary Remains. Oxford: Oxbow Books. Halcrow, Siân E., and Nancy Tayles. 2008. “The Bioarchaeological Investigation of Childhood and Social Age: Problems and Prospects.” Journal of Archaeological Method and Theory 15: 190–215. Halcrow, Siân E., and Nancy Tayles. 2011. “The Bioarchaeological Investigation of Children and Childhood.” In Social Bioarchaeology, edited by Sabrina C. Agarwal and Bonnie A. Glencross, 333–360. Hoboken, NJ: Wiley-Blackwell. Harrod, Ryan P. 2017. The Bioarchaeology of Social Control. New York: Springer. Harrod, Ryan P., and Debra L. Martin. 2015. “Bioarchaeological Case Studies of Slavery, Captivity, and Other Forms of Exploitation.” In The Archaeology of Slavery, edited by Lydia Wilson Marshall, 41–63. Carbondale: Southern Illinois University. Holt, Sarah, Stacey Hallman, Mary Lucas Powell, and Maia M. Langley. 2014. “The ‘Other’ Burials at Torre de Palma: Childhood as Special Death in a Medieval Portuguese Site.” In Tracing Childhood: Bioarchaeological Investigations of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta Alfonso-Durruty, and John J. Crandall, 75–98. Gainesville: University Press of Florida. Jenks, Kelly L. 2013. “Building Community: Exploring Civic Identity in Hispanic New Mexico.” Journal of Social Archaeology 13(3): 371–393. Joyce, Rosemary A. 2017. “Sex, Gender, and Anthropology: Moving Bioarchaeology Outside the Subdiscipline.” In Exploring Sex and Gender in Bioarchaeology, edited by Sabrina C. Agarwal and Julie K. Wesp, 1–12. Albuquerque: University of New Mexico Press. Juengst Sara L., and Sara K. Becker. 2017. “The Bioarchaeology of Community.” Archeological Papers of the American Anthropological Association, special issue, volume 28. Kamp, Kathryn A. 2009. “Children in an Increasingly Violent Social Landscape: A Case Study from the American Southwest.” Childhood in the Past 2(1): 71–83. Klaus, Haagen D. 2008. “Bioarchaeology of Life and Death in Colonial South America: Systemic Stress, Adaptation, and Ethnogenesis in the Lambayeque Valley, Peru AD 900–1750.” Doctoral dissertation, The Ohio State University, Departent of Anthropology. Klaus, Haagen D. 2012. “The Bioarchaeology of Structural Violence.” In The Bioarchaeology of Violence, edited by Debra L. Martin, Ryan P. Harrod, and Ventura R. Pérez, 29–62. Gainesville: University Press of Florida. Knudson, Kelly J., and Christopher M. Stojanowski. 2009. “The Bioarchaeology of Identity.” In Bioarchaeology and Identity in the Americas, edited by Kelly J. Knudson and Christopher M. Stojanowski, 1–23. Gainesville: University Press of Florida. Knüsel, Christopher J. 2011. “Men Take Up Arms for War: Sex and Status Distinctions of Humeral Medial Epicondylar Avulsion Fractures in the Archaeological Record.”

Identity Revisited: A Brief Introduction · 15

In Breathing New Life into the Evidence of Death: Contemporary Approaches to Bioarchaeology, edited by Aubrey Baadsgaard, Alexis T. Boutin, and Jane E. Buikstra, 221–250. Santa Fe, NM: School for Advanced Research Press. Kurin, Danielle S. 2012. “The Bioarchaeology of Collapse: Ethnogenesis and Ethnocide in Post-Imperial Andahuaylas, Peru (AD 900–1250).” Doctoral dissertation, Vanderbilt University, Department of Anthropology. Kurin, Danielle S. 2016. “Trauma, Nutrition, and Malnutrition in the Andean Highlands During Peru’s Dark Age (1000–1250 C.E.).” In The Archaeology of Food and Warfare, edited by Amber M. VanDerwarker and Gregory D. Wilson, 229–257. New York: Springer International Publishing. Larsen, Clark S. 2018. “Bioarchaeology in Perspective: From Classifications of the Dead to Conditions of the Living.” American Journal of Physical Anthropology 165(4): 865– 878. Lewis, Cecil, Jr., and Tiffiny A. Tung. 2013. “Methodological and Ethical Considerations When Sampling Human Osteological Remains.” In The Dead Tell Tales: Essays in Honor of Jane E. Buikstra, edited by María Cecilia Lozada and Barra O’Donnabhain, 24–30. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press. Lorentz, K.O. 2017. “Marking Identity through Cultural Cranial Modification within the First Sedentary Communities (Ninth to Eighth Millennium BCE) in the Near East: Tepe Abdul Hosein, Iran.” International Journal of Osteoarchaeology 27(6): 973–983. Lucy, S.J. 1997. “Housewives, Warriors, and Slaves? Sex and Gender in Anglo-Saxon Burials.” In Invisible People and Processes, edited by Jenny Moore and Eleanor Scott, 150–168. London: Leicester University Press. Mac Sweeney, Naoise. 2009. “Beyond Ethnicity: The Overlooked Diversity of Group Identities.” Journal of Mediterranean Archaeology 22(1): 101–126. Mant, Madeleine, and Alyson Holland. 2019. Bioarchaeology of Marginalized People. New York: Academic Press. Martin, Debra L., and Ryan P. Harrod. 2014. Bioarchaeology of Climate Change and Violence: Ethical Considerations. New York: Springer. Martin, Debra L., and Caryn Tegtmeyer. 2017. Bioarchaeology of Women and Children in Times of War. New York: Springer. Martin, Debra L., Jennifer L. Thompson, and John J. Crandall. 2014. “Children of the Working Class: Environmental Marginality and Child Health at Black Mesa, Arizona (AD 900–1150).” In Tracing Childhood: Bioarchaeological Investigations of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta Alfonso-Durruty, and John J. Crandall, 198–218. Gainesville: University Press of Florida. Mays, S., T. Prowse, M. George, and M. Brickley. 2018. “Latitude, Urbanization, Age, and Sex as Risk Factors for Vitamin D Deficiency Disease in the Roman Empire.” American Journal of Physical Anthropology 167(3): 484–496. Mays, Simon. 2014. “The Bioarchaeology of the Homicide of Infants and Children.” In Tracing Childhood: Bioarchaeological Investigations of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta Alfonso-Durruty, and John J. Crandall, 99–122. Gainesville: University Press of Florida. Novak, Shannon A. 2009. “Beneath the Facade: A Skeletal Model of Domestic Violence.”

16 · Christopher M. Stojanowski and Kelly J. Knudson

In Social Archaeology of Funeral Remains, edited by Rebecca Gowland and Christopher Knüsel, 238–252. Oxford: Oxbow Books. Nystrom, Kenneth C. 2014. “The Bioarchaeology of Structural Violence and Dissection in the 19th‐Century United States.” American Anthropologist 116(4): 765–779. Nystrom, Kenneth C. 2017. The Bioarchaeology of Dissection and Autopsy in the United States. New York: Springer. Ortman, Scott G. 2012. “Winds from the North: Tewa Origins and Historical Anthropology.” Doctoral dissertation, Arizona State University, School of Human Evolution and Social Change. Palkovich, Ann M. 2014. “Surviving Childhood: Health, Identity, and Personhood in the Prehistoric American Southwest.” In Tracing Childhood: Bioarchaeological Investigations of Early Lives in Antiquity, edited by Jennifer L. Thompson, Marta AlfonsoDurruty, and John J. Crandall, 219–227. Gainesville: University Press of Florida. Pearson, Jessica, Matt Grove, Metin Özbek, and Hitomi Hongo. 2013. “Food and Social Complexity at Çayönü Tepesi, Southeastern Anatolia: Stable Isotope Evidence of Differentiation in Diet According to Burial Practice and Sex in the Early Neolithic.” Journal of Anthropological Archaeology 32(2): 180–189. Rakita, Gordon F. M. “Bioarchaeology as a Process: An Examination of Bioarchaeological Tribes in the USA.” In Archaeological Human Remains: Global Perspectives, edited by Barra O’Donnabhain and María Cecilia Lozada, 213–234. New York: Springer. Redfern, Rebecca C. 2017. “Identifying and Interpreting Domestic Violence in Archaeological Human Remains: A Critical Review of the Evidence.” International Journal of Osteoarchaeology 27(1): 13–34. Reitsema, Laurie J., and Giuseppe Vercellotti. 2012. “Stable Isotope Evidence for Sex-and Status-Based Variations in Diet and Life History at Medieval Trino Vercellese, Italy.” American Journal of Physical Anthropology 148(4): 589–600. Reynard, Linda M., and Noreen Tuross. 2015. “The Known, the Unknown and the Unknowable: Weaning Times from Archaeological Bones Using Nitrogen Isotope Ratios.” Journal of Archaeological Science 53: 618–625. Schepartz, Lynne A. 2017. “The Body of Power, the Body of Memory.” Reviews in Anthropology 46(1): 35–53. Schmidt, Jodi, Cynthia Kwok, and Anne Keenleyside. 2016. “Infant Feeding Practices and Childhood Diet at Apollonia Pontica: Isotopic and Dental Evidence.” American Journal of Physical Anthropology 159(2): 284–299. Schrader, Sarah A. 2013. “Bioarchaeology of the Everyday: Analysis of Activity Patterns and Diet in the Nile Valley.” Doctoral dissertation, Purdue University, Department of Anthropology. Schrader, Sarah A. 2018. “The Anthropology and Bioarchaeology of Quotidian Experiences.” In Activity, Diet and Social Practice, edited by Sara Schrader, 1–17. New York: Springer. Sheridan, Susan G., Jaime Ullinger, Lesley A. Gregoricka, and Meredith S. Chesson. 2014. “Bioarchaeological Reconstruction of Group Identity at Early Bronze Age Bab edgDhra’, Jordan.” In Remembering the Dead in the Ancient Near East: Recent Contribu-

Identity Revisited: A Brief Introduction · 17

tions from Bioarchaeology and Mortuary Archaeology, edited by Benjamin W. Porter and Alexis T. Boutin, 133–184. Boulder: University Press of Colorado. Sofaer, Joanna. 2005. The Body as Material Culture. Cambridge: Cambridge University Press. Sofaer Deverenski, Joanna. 1997. “Age and Gender at the Site of Tiszapolgár-Basatanya, Hungary.” Antiquity 71(274): 875–889. Somerville, Andrew D., Paul S. Goldstein, Sarah I. Baitzel, Karin L. Bruwelheide, Allisen C. Dahlstedt, Linda Yzurdiaga, Sarah Raubenheimer, Kelly J. Knudson, and Margaret J. Schoeninger. 2015. “Diet and Gender in the Tiwanaku Colonies: Stable Isotope Analysis of Human Bone Collagen and Apatite from Moquegua, Peru.” American Journal of Physical Anthropology 158(3): 408–422. doi:10.1002/ajpa.22795 Spielmann, Katherine A. 1995. “Glimpses of Gender in the Prehistoric Southwest.” Journal of Anthropological Research. 51(2): 91–102. Stojanowski, Christopher M., Charisse L. Carver, and Katherine A. Miller. 2014. “Incisor Avulsion, Social Identity and Saharan Population History: New Data from the Early Holocene Southern Sahara.” Journal of Anthropological Archaeology 35: 79–91. Stojanowski, Christopher M., and William N. Duncan. 2015. “Engaging Bodies in the Public Imagination: Bioarchaeology as Social Science, Science, and Humanities.” American Journal of Human Biology 27(1): 51–60. Stone, Pamela K., ed. 2018. Bioarchaeological Analyses and Bodies: New Ways of Knowing Anatomical and Archaeological Skeletal Collections. New York: Springer. Stovel, Emily. 2013. “Concepts of Ethnicity and Culture in Andean Archaeology.” Latin American Antiquity 24(1): 3–20. Tiesler, Vera. 2014. The Bioarchaeology of Artificial Cranial Modifications. New York: Springer. Tiesler, Vera, and Alfonso Lacadena. 2018. “Head Shapes and Group Identity on the Fringes of the Maya Lowlands.” In Social Skins of the Head: Body Beliefs and Ritual in Ancient Mesoamerica and the Andes, edited by Vera Tiesler and Maria Cecilia Lozada, 37–58. Albuquerque: University of New Mexico Press. Tilley, Lorna, and Alicia A. Schrenk. 2017. New Developments in the Bioarchaeology of Care. New York: Springer. Timmins, S., C. Seréville-Niel, and M. Brickley. 2017. “Childhood Cranial Trauma from a Late Roman and Merovingian Context from Michelet, Lisieux, France.” International Journal of Osteoarchaeology 27(4): 715–722. Torres-Rouff, Christina. 2011. “Piercing the Body: Labret Use, Identity, and Masculinity in Prehistoric Chile.” In Breathing New Life into the Evidence of Death: Contemporary Approaches to Bioarchaeology, edited by Aubrey Baadsgaard, Alexis T. Boutin, and Jane E. Buikstra, 153–178. Santa Fe, NM: School for Advanced Research Press. Torres-Rouff, Christina, and Kelly J. Knudson. 2017. “Integrating Identities: An Innovative Bioarchaeological and Biogeochemical Approach to Analyzing the Multiplicity of Identities in the Mortuary Record.” Current Anthropology 58(3): 381–409. Tung, Tiffiny. 2012. “Violence against Women: Differential Treatment of Local and Foreign Females in the Heartland of the Wari Empire.” In The Bioarchaeology of Violence,

18 · Christopher M. Stojanowski and Kelly J. Knudson

edited by Debra L. Martin, Ryan P. Harrod, and Ventura R. Pérez, 180–200. Gainesville, FL: University Press of Florida. Tung, Tiffiny A., and Kelly J. Knudson. 2010. “Childhood Lost: Abductions, Sacrifice, and Trophy Heads of Children in the Wari Empire.” Latin American Antiquity 21(1): 44–66. Tung, Tiffiny A., Melanie Miller, Larisa DeSantis, Emily A. Sharp, and Jasmine Kelly. 2015. “Patterns of Violence and Diet among Children during a Time of Imperial Decline and Climate Change in the Ancient Peruvian Andes.” In The Archaeology of Food and Warfare: Food Insecurity in Prehistory, edited by Amber M. VanDerwarker and Gregory D. Wilson, 193–228. New York: Springer International Publishing. Turner, Bethany L., and Valerie A. Andrushko. 2011. “Partnerships, Pitfalls, and Ethical Concerns in International Bioarchaeology.” In Social Bioarchaeology, edited by Sabrina C. Agarwal and Bonnie A. Glencross, 44–67. Hoboken, NJ: Wiley-Blackwell. Turner, Trudy R., Robin M. Bernstein, Andrea B. Taylor, Abigail Asangba, Traci Bekelman, Jennifer Danzy Cramer, Sarah Elton, Katarina Harvati, Erin Marie WilliamsHatala, Laurie Kauffman, Emily Middleton, Joan Richtsmeier, Emőke Szathmáry, Christina Torres-Rouff, Zaneta Thayer, Amelia Villaseñor, and Erin Vogel. 2018. “Participation, Representation, and Shared Experiences of Women Scholars in Biological Anthropology.” American Journal of Physical Anthropology 165(S65): 126–157. Turner, Trudy R., Jennifer K. Wagner, and Graciela S. Cabana. 2018. “Ethics in Biological Anthropology.” American Journal of Physical Anthropology 165(4): 939–951. Verlinden, Petra, and Mary E. Lewis. 2015. “Childhood Trauma: Methods for the Identification of Physeal Fractures in Nonadult Skeletal Remains.” American Journal of Physical Anthropology 157(3): 411–420. doi:10.1002/ajpa.22732 Velasco, Matthew C. 2018. “Ethnogenesis and Social Difference in the Andean Late Intermediate Period (AD 1100–1450).” Current Anthropology 59(1): 98–106. Voss, Barbara L. 2007. “Feminisms, Queer Theories, and the Archaeological Study of Past Sexualities.” In The Archaeology of Identities: A Reader, edited by Timothy Insoll, 124–136. New York: Routledge. Voss, Barbara L. 2015. “What’s New? Rethinking Ethnogenesis in the Archaeology of Colonialism.” American Antiquity 80(4): 655–670. Watson, James T., and Marijke Stoll. 2013. “Gendered Logistic Mobility among the Earliest Farmers in the Sonoran Desert.” Latin American Antiquity 24(4): 433–450. Western, A.G., and J.J. Bekvalac. 2017. “Hyperostosis Frontalis Interna in Female Historic Skeletal Populations: Age, Sex Hormones and the Impact of Industrialization.” American Journal of Physical Anthropology 162(3): 501–515. Wheeler, Sandra M., Lana Williams, Patrick Beauchesne, and Tosha L. Dupras. 2013. “Shattered Lives and Broken Childhoods: Evidence of Physical Child Abuse in Ancient Egypt.” International Journal of Paleopathology 3(2): 71–82. Wright, Lori E. 2013a. “Examining Childhood Diets at Kaminaljuyu, Guatemala, through Stable Isotopic Analysis of Sequential Enamel Microsamples.” Archaeometry 55(1): 113–133. Wright, Lori E. 2013b. “An Isotope Study of Childhood Diet and Mobility at Copan, Honduras.” In The Dead Tell Tales: Essays in Honor of Jane E. Buikstra, edited by María

Identity Revisited: A Brief Introduction · 19

Cecilia Lozada and Barra O’Donnabhain, 95–105. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press. Zuckerman, Molly K. 2017. “Mercury in the Midst of Mars and Venus: Reconstructing Gender, Sexuality, and Socioeconomic Status in Relation to Mercury Treatment for Syphilis in Seventeenth- to Nineteenth-Century London.” In Exploring Sex and Gender in Bioarchaeology, edited by Sabrina C. Agarwal and Julie K. Wesp, 223–262. Albuquerque: University of New Mexico Press. Zuckerman, Molly K., and John Crandall. 2019. “Reconsidering Sex and Gender in Relation to Health and Disease in Bioarchaeology.” Journal of Anthropological Archaeology 54: 161–171. Zuckerman, Molly K., Kelly R. Kamnikar, and Sarah A. Mathena. 2014. “Recovering the ‘Body Politic’: A Relational Ethics of Meaning for Bioarchaeology.” Cambridge Archaeological Journal 24(3): 513–522.

2 Exploring Family, Ethnic, and Regional Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru Kent M. Johnson

Ethnographic and ethnohistoric accounts attest to the importance of ethnicity and family in the social fabric of Andean communities past and present. Archaeological data indicate that ethnic identity was integral to social organization in pre-Hispanic Andean societies, but the significance of family identity has not been investigated to the same extent in archaeological contexts. For example, previous bioarchaeological research in the Moquegua Valley, Peru, has fruitfully explored ethnicity at Tiwanaku-affiliated sites, but efforts to investigate family and kinship at the same sites have yielded mixed results. Here, I expand on prior research in the Moquegua Valley using a regional, multiscalar approach to social identity that includes ethnicity and family. I define social identity as culturally salient aspects of affiliation and identification to which persons assign themselves and others (see Barth 1969; Epstein 1978; Jenkins 2008; Jones 1997). Social identities structure social organization through expectations for normative values and behaviors, as conceived and enacted according to community standards, concerning social interactions with others variably defined as similar to or different from oneself in regard to various intersecting dimensions of collective affiliation. Examples of salient aspects of social identity in contemporary U.S. society include age, sex and gender, sexuality, race, ethnicity, religion, and political party affiliation. Bioarchaeological research tends to focus on one of two different analytical scales: population-based approaches that emphasize large-scale

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 21

collectives and individual-based approaches such as biohistories (Komar and Buikstra 2008; Stojanowski and Duncan 2017) and osteobiographies (e.g., Boutin 2012; Hawkey 1998; Robb 2002; Saul 1972; Saul and Saul 1989; Stodder and Palkovich 2012; Zvelebil and Weber 2013). In social identity research, this manifests in studies that emphasize individual- or community-based identities (Knudson and Stojanowski 2009). However, multiscalar studies that consider how dimensions of social identity structure interaction patterns at multiple analytical scales (e.g., individual, community, population, and regional levels) are becoming more prevalent (e.g., Blom 1999; Cornelison, Lackey-Cornelison, and Goldstein 2017; Miller 2013; Pilloud 2009; Torres-Rouff and Knudson 2017; see Johnson and Paul 2016). Although recent studies have begun exploring the internal structure of larger communities (e.g., Alt and Vach 1998; Blom 1999; Cornelison, Lackey-Cornelison, and Goldstein 2017; Hoshower et al. 1995; Marsteller 2015; Meyer et al. 2012; Pilloud 2009; Stojanowski 2013), bioarchaeological analyses of midlevel social collectives such as peer groups, neighborhoods, and family groups (e.g., clans and households) remain underdeveloped. These “intervening levels” of affiliation are critical to reconstructing social organization in the past because they connect individuals within and across larger-scale social and political collectives and shape individual behavior (Read and van der Leeuw 2015: 32). In this chapter, I present a multiscalar approach to social organization within the Tiwanaku colonies of the Moquegua Valley, Peru. I analyze family-, ethnic-, and region-based affiliations to develop a holistic approach to social organization within and between pre-Hispanic Andean communities. Following Barth (1969), I take ethnicity to be a dynamic process of self–other differentiation wherein cultural groups create and maintain collective boundaries using culturally salient markers rather than an innate or primordial aspect of collective identity. Thus, ethnicity is not an inherent characteristic of a social group whose members share cultural traditions and perhaps biological ancestry; ethnicity is a process of social interaction in which individuals intentionally use cultural attributes to signal collective affiliations and differences within specific historic, political, and economic contexts (e.g., Janusek 2005; Jones 1997; Lucy 2005; Meskell 2007; Stovel 2013). Here, I analyze cranial modification and phenotypic data from human skeletal remains from five Tiwanaku-affiliated sites to explore whether biocultural indicators of ethnic identity are consistent with archaeological

22 · Kent M. Johnson

evidence of strong ethnic community boundaries. I use biodistance and exploratory data analysis of phenotypic data to evaluate two informal models of gene flow and interaction among skeletal samples from Tiwanaku-affiliated sites: isolation by distance and the dual diaspora model. I apply social network visualization and analytical techniques to cranial shape data to explore family organization as an intermediate level of social organization. I argue that multiple social identities, including family, ethnic, and regional affiliations, influenced social life within the Moquegua Tiwanaku colonies. Further, I suggest family networks were critical to forging a collective Moquegua Tiwanaku identity among individuals from different ethnic backgrounds. Tiwanaku Communities in the Moquegua Valley, Peru

Situated in the Andean altiplano (high plain) of present-day Bolivia, the settlement of Tiwanaku developed into a complex urban center whose influence spanned the south central Andes during the Middle Horizon period (c. AD 500–1100) (Anderson 2013; Bermann 1997; Browman 1997; Goldstein 1993, 2005; Janusek 2004, 2008; Kolata 1992, 1993a, 1993b; Oakland Rodman 1992; Stanish 2003). The Tiwanaku polity was a loosely confederated segmentary state whose affiliated communities were linked through noncoercive, reciprocal relationships cemented through ritual, shared ideology, and economic interaction (Albarracín-Jordán 1996, 2003; Browman 1978, 1997; Goldstein 2005; McAndrews, AlbarracínJordán, and Bermann 1997; cf. Kolata 1993a, 1993b; Stanish 2002, 2003). Tiwanaku influence spread, in part, through the establishment of agricultural colonies in lower altitude, warmer areas of the Andes conducive to growing maize, cotton, coca, and other warm-weather plants (Berryman 2010; Goldstein 2003; Kolata 1992). A substantial Tiwanaku agricultural colony was located in the Moquegua Valley of southern Peru, approximately three hundred kilometers southwest of the capital of Tiwanaku (Blom et al. 1998; Goldstein 2005). Between around AD 600 and AD 1100, two distinct Tiwanakuaffiliated populations established settlements at multiple sites in the valley. The earliest were camelid agropastoralists associated with blackware ceramics who are referred to archaeologically as “Omo-style” Tiwanaku communities. They were followed by maize agricultural specialists associated with redware ceramics who are archaeologically referred to as

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 23

Figure 2.1. Map of archaeological sites from the middle Moquegua Valley, Peru, included in the present study. Map data: Google, DigitalGlobe, CNES/Airbus. Inset map data: Google, SIO, NOAA, U.S. Navy, NGA, GEBCO, Landsat/Copernicus.

“Chen Chen–style” Tiwanaku communities (Goldstein 2005, 2015; Owen 2005).1 The producers of Chen Chen–style ceramics settled alongside the communities who produced Omo-style ceramics at several of the largest site groups in the middle Moquegua Valley (figure 2.1) but generally maintained distinct settlements within those site groups (Goldstein 2005). Despite living in close proximity, Chen Chen– and Omo-style communities maintained distinct cultural identities for over three hundred years (Goldstein 2005, 2015; Owen 2005), marked by differences in settlement patterns, residential architecture, and funerary practices in addition to the aforementioned differences in ceramic assemblages (Baitzel 2008,

24 · Kent M. Johnson

2016; Goldstein 1989, 1993, 2000, 2005; Hoshower et al. 1995; Owen 2005). This “intentional persistence of differentiation” despite cohabitation and interaction is consistent with active ethnic identity expression rather than passive expressions of cultural tradition (Stovel 2013: 11; see also Barth 1969; Jones 1997). Thus, archaeologists describe these populations as “Omo-style” and “Chen Chen–style” ethnic or ethnic-like communities (e.g., Goldstein 2000, 2005, 2015; Owen 2005; see also Becker and Goldstein 2018; Somerville et al. 2015). Goldstein (2005) suggests these Tiwanaku-affiliated communities maintained separate ethnic identities, in part through endogamous marriage practices, but this hypothesis has not been formally evaluated using biological data. Early bioarchaeological studies corroborated the importance of ethnic identity to social life in Moquegua Tiwanaku communities but found mixed evidence concerning the centrality of corporate kin groups. Blom (1999, 2005a, 2005b; Blom et al. 1998) found little evidence of spatial structuring of phenotypic traits and cranial modification practices among samples from the cemeteries at the Chen Chen–style site of Chen Chen M1, suggesting that a common ethnic identity was emphasized over smaller-scale corporate or kin group affiliations (see also Lewis 2005; Sharratt 2011). Additionally, Blom (1999, 2005a, 2005b) found that the culturally modified cranial shape favored by Moquegua colonists differed from the shape favored by inhabitants of Tiwanaku-affiliated sites in the Katari Valley in the altiplano, evoking strong region-based distinctions within Tiwanaku society, although both modification styles are present among Middle Horizon skeletal samples from the Tiwanaku Valley. In contrast, Hoshower and colleagues (1995) found that subtle differences in cranial modification practices at the Chen Chen–style site of Omo M10 were spatially structured. Skeletal samples from different cemeteries at Omo M10 exhibit different patterns of head shaping. This is interpreted as possible evidence that corporate group or lineage-based identities were important to the communities that used the M10 cemeteries (Hoshower et al. 1995; cf. Baitzel 2016). Together, these early bioarchaeological studies provided critical insights on social organization within Tiwanaku society. Blom’s work affirms that social life within the Moquegua Valley Tiwanaku colonies was structured by ethnic affiliations. Collectively, studies by Blom and Hoshower and colleagues demonstrate the need to consider multiple levels of affiliation (family, ethnic, and interregional) in investigations of

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 25

Tiwanaku social organization in the Moquegua Valley. However, skeletal samples representing Omo-style ethnic communities were unavailable to early investigators, and there has been no direct investigation of ethnic interaction using phenotypic and cranial modification data. Further, social organization has not been investigated from an intraregional perspective, comparing phenotypic data and cranial modification practices in samples from multiple Tiwanaku-affiliated sites in the Moquegua Valley. As a result, essential aspects of Moquegua Valley Tiwanaku social organization remain under-investigated. Recent bioarchaeological research has explored similarities and differences among samples from Chen Chen– and Omo-style contexts in terms of activity patterns (Becker and Goldstein 2018), diet (Somerville et al. 2015), and paleomobility (Knudson et al. 2014). Becker and Goldstein (2018) found no overall differences between Chen Chen– and Omo-style samples in activity patterns as evidenced by rates of osteoarthritis (OA), but they did identify subtle differences in skeletal evidence of OA when age and skeletal sex were considered along with ethnic affiliation. Younger adults from Chen Chen–style contexts had significantly higher rates of OA in the wrist, whereas middle‐aged adults from Omo‐style contexts had significantly higher rates of OA in the shoulder and sacroiliac joints (Becker and Goldstein 2018). Compared to their counterparts from Chen Chen–style contexts, females from Omo-style contexts had significantly higher rates in the right hip joint, and middle‐aged females from Omo-style contexts had higher rates in both sides of the sacroiliac joint (Becker and Goldstein 2018). These findings may suggest distinctions in habitual activities between Chen Chen– and Omo-style communities at different stages of the life course and by sex. Despite the aforementioned differences in subsistence practices, there is little isotopic evidence to suggest any dietary differences between skeletal samples from Tiwanaku-affiliated sites in the Moquegua Valley (Somerville et al. 2015). Specifically, individuals from the Omo-style site of Rio Muerto M70 and the Chen Chen–style sites of Chen Chen M1, Omo M10, and Rio Muerto M43 exhibit similar ranges of carbon and nitrogen stable isotope values. These findings suggest that communities at the different townsites had comparable access to C4 plants and marine resources and similar or perhaps reciprocal subsistence strategies (Somerville et al. 2015).

26 · Kent M. Johnson

In a study of paleomobility among samples from the Rio Muerto sites (M43 and M70), Knudson and colleagues (2014) identified four potential first-generation migrants to the Moquegua Valley, two of whom were buried in the Chen Chen–style M43B cemetery and two of whom were buried in the Omo-style M70B cemetery. In contrast, the only potential isotopic evidence for transhumance among the individuals sampled came from four adult males buried in the Omo-style M70B cemetery (Knudson et al. 2014). It is compelling that both ethnic communities at the Rio Muerto townsites were receiving new immigrants, but the only evidence of repeated movement between the Lake Titicaca Basin and the Moquegua Valley is from individuals from the Omo-style M70B cemetery, which is consistent with expectations for a highly mobile agropastoral lifestyle. What emerges from recent bioarchaeological research on life in the Moquegua Tiwanaku colonies is a more fine-grained perspective on similarities and differences in cultural practices among members of the two ethnic communities. These findings underscore the importance of comparative studies for producing nuanced reconstructions of social life in the Tiwanaku colonies and of investigating, rather than assuming, that specific (or all) cultural practices were used to signal ethnic affiliations (see Jones 1997; Sutter 2005). An intraregional, multiscalar analysis of cranial modification style and phenotypic data that includes samples from both Chen Chen– and Omo-style contexts can build on this literature and provide an effective way to (1) assess Goldstein’s hypothesis regarding ethnic group endogamy and (2) clarify the role of family affiliations within the overall scheme of Moquegua Tiwanaku social organization. Ethnohistoric accounts from the Andes describe cranial modification as a highly visible and permanent means of signaling group affiliations at a variety of sociocultural scales (e.g., Cieza de León 1984 [1553]; de la Vega 1966 [1609]). In Tiwanaku contexts, cranial modification is linked to regional, ethnic, and corporate level affiliations (Blom 1999, 2005a; Blom et al. 1998; Hoshower et al. 1995; Torres-Rouff 2002, 2008). As discussed above, previous research indicates that two different scales of social affiliation were emphasized at different Chen Chen–style sites in the Moquegua Valley: ethnic affiliation at Chen Chen M1 and corporate group affiliation at Omo M10 (Blom 1999; Hoshower et al. 1995). The current study analyzes cranial modification presence and style in skeletal samples from five Tiwanaku-affiliated sites to evaluate patterns of cranial modification in Chen Chen– and Omo-style communities and assess whether cranial

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 27

Table 2.1. Basicranial and temporal bone landmarks Landmark

Description

1. Basion

3. Inferior nuchal

Midline point on the anterior margin of the foramen magnum Midline point at the posterior margin of the foramen magnum Midline point on the inferior nuchal line

4. Condylar foramen

The posterior point on the margin of the condylar foramen

5. Condyle posterior

The most posterior point on the occipital condyle

6. Condyle anterior

The most anterior point on the occipital condyle

7. Jugular

Most lateral point of the jugular fossa

8. Mastoidale

The most inferior point on the mastoid process

9. Postglenoid

Most inferior point on the postglenoid process

10. Lateral eminence

12. Entoglenoid

Point on the center of the lateral margin of the articular surface of the articular eminence Most anterior point on the articular surface of the articular eminence Most inferior point on the entoglenoid process

13. Lateral ovale

Most lateral point on the margin of the foramen ovale

14. Petrous apex

Apex of petrous part of the temporal bone

15. Tympanic

Most inferolateral point on the tympanic element of the temporal Most superior point of the external auditory meatus

2. Opisthion

11. Anterior articular

16. Porion 17. Auriculare

Point of deepest incurvature on the lateral aspect of the root of the zygomatic process

modification styles covary with other cultural characteristics used to signal ethnic identity. To assess whether Omo- and Chen Chen–style communities represent endogamous ethnic groups, I use biodistance and exploratory data analyses of basicranial and temporal bone landmarks to investigate patterns of morphological variation and gene flow among samples of human skeletal remains from the Moquegua Valley (table 2.1). Biodistance analysis uses skeletal and dental metric and morphological traits as proxies for the degree of genetic relatedness among skeletal samples (Buikstra, Frankenberg, and Konigsberg 1990; Larsen 2015). Results of biodistance analysis can be used to make inferences regarding population structure

28 · Kent M. Johnson

and population history, including patterns of marital migration and gene flow (Cheverud 1988; Konigsberg and Ousley 1995; Spence 1974a, 1974b). It is critical to evaluate Goldstein’s hypothesis using biological data because cultural and biological boundaries are not inherently coterminous. Biological distance and population genetic studies have found that social boundaries can influence patterns of gene flow through socially prescribed parameters for mate selection (Cannings and Skolnick 1975; Chapman 1993; Fix 1979; Relethford 2010; Stojanowski 2010; Workman, Mielke, and Nevanlinna 1976). However, numerous biodistance studies have documented cases where marked social boundaries do not inhibit mate exchange and gene flow (e.g., Lozada Cerna 1998; Morton et al. 1971; Tomczak 2001). Despite diverse lines of evidence suggestive of pronounced social boundaries between Chen Chen–style and Omo-style Tiwanaku-affiliated communities, it is necessary to evaluate whether patterns of phenotypic variation and gene flow within the Moquegua Tiwanaku colonies are consistent with two endogamous ethnic communities. As Stovel (2013: 11) notes, phenotypic and genetic data can serve as “interesting counterpoints to material culture.” When biological data do not correspond with material culture indicators of identity, this pattern may represent strong evidence of intentional ethnic identity construction (Stovel 2013). Materials and Methods

I collected cranial modification and phenotypic data from human skeletal remains from five Tiwanaku-affiliated sites in the middle Moquegua Valley (table 2.2). Chen Chen M1 is a large mortuary site with multiple, distinct Chen Chen–style cemeteries (see Blom 1999; Sharratt 2011). Omo Alto M16 and Rio Muerto M70 are Omo-style sites with domestic and mortuary components (Baitzel 2008; Goldstein 2005). Rio Muerto M43 is a culturally complex multicomponent site. The domestic sector and cemetery M43B are Chen Chen–style contexts, while cemetery M43A appears to be a hybrid Chen Chen– and Omo-style mortuary context (see Baitzel 2016). Omo M10 has domestic, mortuary, and monumental sectors, and it is home to the only Tiwanaku-style temple outside of the altiplano (Goldstein 2005), which may explain the cultural diversity present in the cemetery sector. The majority of Middle Horizon period cemeteries are Chen

Table 2.2. Study samples by site Site

Calibrated Dates

Cultural Affiliation

Observed

Included in Studya

Chen Chen M1

AD 656–1155b

Chen Chen

446

45c

Omo M10

AD 705–1005d

Chen Chen

223

35e

AD

765–1025f

Omo Alto M16

AD 635–890g

Omo

22

3

Rio Muerto M43

AD 780–1017h

Chen Chen

65

7i

Rio Muerto M70

AD 705–1005j

Omo

78

12

834

102

AD 780–997j Total aThis

represents the number of intact and undamaged adult crania suitable for Notes: collection of basicranial and temporal bone landmarks. bThe dates listed represent the maximum range derived from twelve calibrated radiocarbon dates (2 sigma) reported by Sharratt (2011: 156, table 5). cThe sample from Chen Chen M1 is drawn from nine of the cemeteries at the site: A (n = 4), B (n = 2), C (n = 3), D (n = 4), E (n = 1), I (n = 7), K (n = 3), L (n = 6), and 30 (n = 1). Additionally, fourteen individuals of unclear provenience within the M1 mortuary sector are included in the study sample. dCalibrated radiocarbon date (2 sigma) reported by Goldstein (1993). eThe sample from Omo M10 is drawn from ten of the cemeteries at the site: B (n = 3), H (n = 2), I (n = 2), M (n = 5), R (n = 2), S (n = 4), T (n = 4), U (n = 2), V (n = 1), and W (n = 2). Each of these cemeteries is considered a Chen Chen–style context, with the exception of Cemetery R, which is a potentially hybrid Chen Chen– and Omostyle context (Baitzel 2016). Additionally, eight individuals of unknown provenience from within the M10 mortuary sector are included in the study sample. fCalibrated radiocarbon date (2 sigma) reported by Goldstein (1989a). gCalibrated radiocarbon date (1 sigma) reported by Goldstein (2005: 128–131, table 5.2). hCalibrated radiocarbon date (2 sigma) reported by Goldstein (2005: 128–131, table 5.2). iThe sample from Rio Muerto M43 is drawn from two cemeteries at the site. Four individuals are from the hybrid M43A cemetery, and two are from the Chen Chen–style M43B cemetery. One individual of unknown provenience from the M43 mortuary component is also included in the study sample (M43–3000). jCalibrated radiocarbon dates (2 sigma) reported by Magilligan and Goldstein (2000).

30 · Kent M. Johnson

Chen–style contexts, but there is also a coastal-style cemetery (cemetery X), a potentially Omo-style cemetery (cemetery P), a potentially hybrid Chen Chen– and Omo-style cemetery (cemetery R), and a post–Chen Chen Tumilaca-style Tiwanaku cemetery (see Goldstein 2005; Baitzel 2016). Of the 834 total burials examined, only 102 individuals have crania suitable for geometric morphometric data collection. As detailed in table 2.2, study samples from Chen Chen M1, Omo M10, and Rio Muerto M43 have been pooled from multiple cemeteries within each site.2 I assessed cranial modification according to a scoring protocol used in previous studies of Moquegua Valley skeletal samples (Blom et al. 1998; Hoshower et al. 1995) to evaluate whether Chen Chen– and Omo-style communities exhibit similar styles of modified cranial shape. I examined crania for evidence of modification, and, for modified crania, recorded modification style as either annular or fronto-occipital. To evaluate whether individuals from Omo-and Chen Chen–style contexts represent distinct biological subpopulations, I recorded seventeen landmarks from the basicranium and temporal bone following standard methodology (Adams, Rohlf, and Slice 2004; McKeown and Jantz 2005; Slice 2005). Landmarks were mechanically registered in three-dimensional space using a Microscribe digitizer MX. Data were assessed for variation attributable to measurement error, age, sex, and cranial modification using standard methodologies.3 There were no significant effects on basicranial and temporal bone shape variation due to age, sex, or cranial modification style. However, the average error for two traits (inferior nuchal crest and posterior condyle) exceeds acceptable levels, and these landmarks were removed from the dataset. One trait, condylar foramen, has high levels of missing data and was removed from the dataset. This resulted in a data matrix that is 97.8 percent complete. Missing landmark coordinates were estimated using the GPA mean substitution method in Morpheus (Slice 2013). A generalized Procrustes analysis was performed on the complete dataset of raw landmark coordinates in MorphoJ (Klingenberg 2011) to extract the shape variation (i.e., Procrustes coordinates) from the raw coordinates. Biodistance data were interpreted with respect to two informal models. The first is a simple isolation by distance model, which assumes that rates of gene flow are structured primarily by spatial proximity. Individuals and communities who live near one another are expected to be more

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 31

phenotypically similar than communities who live farther apart. Under isolation by distance, marriage networks are influenced primarily by spatial proximity rather than by ethnic affiliation. Therefore, under isolation by distance, skeletal samples from neighboring sites, such as Rio Muerto M43 and M70, are expected to cluster together in sample-based plots of phenotypic variation and biodistance, despite being associated with different ethnic communities. Additionally, family-based social networks are expected to include individuals from nearby sites, regardless of ethnic affiliation. In other words, spatial proximity, not ethnicity, structures patterns of mate exchange. The second model is the dual diaspora model and is based on Goldstein’s (2005) research. In this model, Chen Chen– and Omo-style communities are considered endogamous ethnic groups, and individuals and samples affiliated with the same ethnic community are expected to have smaller biodistances, signaling greater phenotypic similarity due to higher levels of intra-ethnic gene flow. Individuals and samples affiliated with different ethnic communities are expected to have greater biodistances due to limited gene flow between ethnic communities. Additionally, under a model of ethnic endogamy, family networks are expected to adhere to ethnic boundaries and include individuals from only a single ethnic community. I performed a combination of sample-based and individual-based multivariate analytical methods to explore patterns of phenotypic variation and evaluate the models. Sample-based analyses include R-matrix analysis and a Mantel test. R-matrix analysis (Harpending and Ward 1982) was performed in RMET 5.0 to generate estimates of phenotypic distances between samples based on craniometric data as well as estimates of extra-local gene flow (Relethford-Blangero residuals) (Relethford 2003; Relethford and Blangero 1990; Relethford, Crawford, and Blangero 1997; Williams-Blangero 1989; Williams-Blangero and Blangero 1989). Phenotypic distances provide estimates of the intensity of mate exchange under the assumption that phenotypic distances are proportional to genetic distances. Relethford-Blangero residuals estimate whether a specific population in a study design received significantly lesser (endogamy) or greater (exogamy) gene flow with a population or populations external to that particular mate exchange network (i.e., the samples included in the analysis). A Mantel test (Mantel 1967; Smouse, Long, and Sokal 1986) was used

32 · Kent M. Johnson

to compare a matrix of spatial distances between sites with a matrix of phenotypic distances among samples. I used individual-based analyses such as discriminant function analysis (DFA) and canonical variates analysis (CVA) to complement the samplebased analyses and provide a fine-grained perspective of data patterns. These ordination methods identify new variables (e.g., factors or variates) from the Procrustes shape coordinates that best differentiate individuals from two (DFA) or more (CVA) known groups. Thus, they provide a means of assessing whether individuals are more phenotypically similar to other individuals from the same ethnic community (DFA) or the same site (CVA). I applied social network visualization and analytical techniques to cranial shape data to explore patterns of morphological variation within the study sample at multiple analytical scales. The fourteen basicranial and temporal bone landmarks were used to generate an inter-individual Mahalanobis (D2) dissimilarity matrix. The matrix was dichotomized to simplify visual representation and to aid interpretation of results.4 The dichotomized D2 matrix was visualized as a network using a graph theoretic layout. I used two approaches to identify potential networks of biological relatives. First, components analysis was used as a top-down approach for identifying potential extended family networks. In graph theory, a component is a maximally connected subgraph (Borgatti, Everett, and Johnson 2013; Scott 2013), a section of a network within which every node (i.e., individual or actor) can reach every other node but there are no connections between individuals in different sections. Second, the ego networks of actors with the highest degree centrality scores were visualized and evaluated as potential biological kin networks.5 Ego network graphs depict a specific actor (ego) and all other actors directly connected to that actor (Hanneman and Riddle 2005). They also depict any direct ties among the various nodes included in the ego network. The ego networks of the individuals with the highest degree centrality scores were graphed together to simulate a bottom-up approach to identifying networks of biological relatives. All network analyses were conducted using the UCINET 6.610 software package (Borgatti, Everett, and Freeman 2002), while visualizations were completed using Netdraw version 2.158 (Borgatti 2002).

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 33

Results

Of the 102 individuals in the sample, ninety-six are observable for cranial modification style. Of those, eighty-six are modified and ten are unmodified (figure 2.2). All eighty-six modified crania exhibit fronto-occipital modification. Fisher’s exact test (p = 0.358) indicates no significant difference between individuals from Chen Chen– and Omo-style contexts in terms of cranial modification presence when individuals from hybrid contexts are excluded. When individuals from the hybrid Chen Chen– style and Omo-style cemeteries of Omo M10R and Rio Muerto M43A are included as a third hybrid sample, Fisher’s exact test (p = 0.426) again indicates no significant difference in cranial modification presence among individuals from Chen Chen–style, Omo-style, and hybrid mortuary contexts. Thus, Omo- and Chen Chen–style samples do not differ in terms of cranial modification presence or style. Patterns of craniometric variation are more complex compared to cranial modification presence and style. The Chen Chen M1 (–0.105, P = 0.0006) and Rio Muerto M43 (–0.007, P = 0.9252) samples have negative Relethford-Blangero residuals, suggesting these communities experienced slightly lower than expected levels of extra-local gene flow, but this is significant at alpha 0.05 only for the sample from Chen Chen M1. Omo M10 (0.038, P = 0.3692), Omo Alto M16 (0.719, P = 0.0031), and Rio Muerto M70 (0.292, P = 0.0249) have positive Relethford-Blangero residuals, indicating samples from these sites experienced greater than expected levels of extra-local gene flow, but these residuals are significant only for the two Omo-style samples from Omo Alto M16 and Rio Muerto M70. It seems members of Omo-style communities married outside of the Moquegua Tiwanaku colonies at a higher rate compared to their counterparts in Chen Chen–style contexts. This is consistent with previous research on pastoral communities that indicates they tend to have more expansive marriage networks and higher rates of exogamy compared to sedentary agricultural communities (e.g., Hausman 1984). Results of the Mantel test indicate a weak, positive correlation between the spatial and biodistance matrices, but the correlation is not statistically significant for either Pearson’s r (0.035, P value = 0.925) or Spearman’s rho (0.115, P value = 0.735). The positive correlations suggest gene flow among the Moquegua Tiwanaku communities is consistent with a model of

Figure 2.2. Unmodified cranium of specimen number M1-548 (top) and frontooccipital cranial modification of specimen number M10 R-7 (bottom) typical of Tiwanaku-affiliated populations in the Moquegua Valley (photos by Kent Johnson).

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 35

Figure 2.3. Genetic distance ordination for the study samples derived from principle components analysis of the scaled R-matrix.

isolation by distance. However, the fact that the correlations are weak and nonsignificant suggests that isolation by distance only partially explains patterns of mate exchange; other factors influenced gene flow within the Moquegua Valley Tiwanaku colonies in addition to spatial proximity. The pattern of genetic distances among samples also suggests spatial proximity was not the sole factor structuring gene flow and marital migration (figure 2.3). The smallest Mahalanobis (D2) distance is between the Chen Chen–style sample from Chen Chen M1 and the Omo-style sample from Rio Muerto M70 (0.081192), the two samples farthest apart spatially. However, this biodistance is not statistically significant at alpha of 0.05, perhaps due to the small sample size for Rio Muerto M70. Discriminant function analysis effectively differentiates individuals from Omo- and Chen Chen–style contexts. The discriminant function correctly allocated 93.1 percent of individuals from Chen Chen–style cemeteries and 86.7 percent of individuals from Omo-style cemeteries. Cross-validation suggests the discriminant function would perform moderately well with new data, correctly allocating 83.9 percent of individuals

36 · Kent M. Johnson

from Chen Chen–style contexts and 66.7 percent of individuals from Omo-style contexts. The poorer cross-validation performance for the Omo-style sample likely reflects the small sample size of individuals from Omo-style contexts and the large proportion of variation within this sample. Regardless, results of DFA indicate Chen Chen– and Omo-style samples do represent distinct biological subpopulations consistent with the expectations of the dual diaspora model. Canonical variates analysis produced three canonical variates with eigenvalues greater than one, and together they account for 87.5 percent of the cumulative variance. A scatter plot of the first two canonical variates indicates a partial differentiation among individuals from contexts associated with Chen Chen–style and Omo-style samples (figure 2.4), but members of these communities do not form distinct clusters. Contrary to expectations of the dual diaspora model of ethnic endogamy, some individuals from different ethnic contexts cluster together. An interesting pattern emerges when considering the same scatter plot with individuals coded by site (figure 2.4). Individuals from the Omostyle cemetery at Omo Alto M16 form a distinct cluster, whereas there is substantial overlap between the Chen Chen–style samples from Omo M10 and Chen Chen M1 and the Omo-style sample from Rio Muerto M70. This may indicate members of the two Omo-style communities (Omo Alto M16 and Rio Muerto M70) had different patterns of marital migration and gene flow. The four individuals from the hybrid Rio Muerto M43A cemetery form a cluster with two individuals from Rio Muerto M70B and a single individual from the Chen Chen–style cemetery of Rio Muerto M43B. In contrast, the two individuals from the potentially hybrid cemetery of Omo M10R do not appear to differ from the rest of the M10 sample, as they are situated near the center of the plot adjacent to individuals from Chen Chen M1, Omo M10, and Rio Muerto M70. In summary, results of ethnic-level and site-based analyses do not neatly support the dual diaspora model or the isolation by distance model. Although the normative marriage pattern likely was to marry someone from the same ethnic community, there were certainly exceptions. These exceptions probably were structured by a mosaic of factors including spatial proximity and individual and/or family group agency. Turning to the social network results, the graph theoretic layout of the inter-individual Mahalanobis distance matrix presents a network with a

Figure 2.4. Scatter plot of the first (43.5%) and second (25.0%) canonical variates of cranial shape coded by ethnic affiliation (top) and site (bottom). Symbols are as follows: (top) white squares = individuals from Chen Chen–style contexts, black circles = individuals from Omo-style contexts; (bottom) black circles = M1, white squares = M10, white circles = M16, white triangles = M43, and black diamonds = M70.

38 · Kent M. Johnson

densely interconnected core, a number of less well-connected nodes, and numerous isolates (figure 2.5). Overall, there is limited structure within the network. Components analysis identified two components that contain more than one individual: component 1 consists of thirty-six individuals from three different sites (Chen Chen M1, Omo M10, and Rio Muerto M70) and component 2 consists of three individuals from the two Rio Muerto sites: M43B and M70B (see figure 2.4). Contrary to the dual diaspora model of ethnic endogamy, both components include individuals from different ethnic communities. The sixty-three individuals in the study sample not present in either of the two components are isolates, meaning they have no ties with any other actor, and are not displayed in the graph. The six individuals from the possible hybrid Chen Chen–style and Omo-style mortuary contexts of M43A and M10R are isolates. This is interesting because one might suspect individuals from hybrid or shareduse contexts might have functioned as brokers (i.e., go-betweens) connecting sections of the network each comprised of only individuals from Omo- or Chen Chen–style contexts, but that does not appear to be the case. Three individuals with the highest degree centrality scores were selected for ego network visualization. Individual M-2 is an adult of undetermined sex from the Chen Chen–style site of Omo M10 with a degree centrality value of 16. Individuals from Omo-style (M70) and Chen Chen–style contexts (M1 and M10) are included in M-2’s ego network. Individual M-5 is an adult female also from the Chen Chen–style site of Omo M10 with a degree centrality value of 13. M-5’s ego network is comprised only of individuals from Chen Chen–style contexts at Chen Chen M1 and Omo M10. Individual 2868 is an adult of undetermined sex from the Omo-style cemetery at Rio Muerto M70 with a degree centrality value of 12. The majority of individuals in individual 2868’s ego network are from Chen Chen–style contexts at Chen Chen M1 and Omo M10; only a single actor from an Omo-style site (Rio Muerto M70) is part of 2868’s ego network. The graph of the combined ego networks (figure 2.6) approximates the shape of component 1, suggesting the top-down and bottom-up social network results are robust.

Figure 2.5. Graph theoretic layout of the interindividual Mahalanobis matrix dichotomized at the 1st percentile (isolates not depicted). Component 1 is on the left and component 2 is on the right. Node color represents ethnic affiliation: white is Chen Chen–style and black is Omo-style. Node shape represents skeletal sex (circle = female, square = male, triangle = undetermined).

Figure 2.6. Combined ego network graphs of the three individuals with the highest degree centrality values: M10 M-2, M10 M-5, and M70 2868. Node color represents ethnic affiliation: white is Chen Chen–style and black is Omo-style. Node shape represents skeletal sex (circle–female, square–male, triangle–undetermined).

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 41

Discussion

Within Tiwanaku society, region-, ethnic-, and family-level social affiliations were signaled through habitual practices and material culture styles (Blom 1999, 2005a, 2005b; Buikstra 1995; Hoshower et al. 1995; Janusek 2004; Goldstein 2005; Owen 2005). Archaeological evidence from the Moquegua Valley indicates that members of Tiwanaku-affiliated colonial communities expressed their shared Tiwanaku cultural heritage yet communicated distinct ethnic identities for several hundred years. Similarly, ethnic boundaries within Moquegua Tiwanaku communities were thought to covary with social and biological divisions (Goldstein 2005, 2015; Hoshower et al. 1995; Owen 2005). Biodistance results presented here suggest that, overall, individuals within the Moquegua Tiwanaku colonies tended to marry within their own ethnic community. This pattern is consistent with Goldstein’s (2005) prediction regarding endogamous marriage practices reinforcing ethnic community boundaries. However, these communities were not entirely endogamous. Although the normative marriage pattern is consistent with ethnic group endogamy, marital migration and gene flow were not prohibited among individuals of different ethnic affiliations. Perhaps marriage practices (endogamy/exogamy) were not part of the suite of cultural practices used to demarcate ethnic boundaries within the Moquegua Valley Tiwanaku colonies. However, given the overall pattern of social endogamy, evidence for marriages across a culturally salient ethnic boundary could represent, along with material culture evidence of cultural hybridization (see below) and shared use mortuary contexts at Omo M10 cemetery R and Rio Muerto M43A, the active alteration (i.e., erasure) of an existing ethnic boundary and the initial emergence of a new ethnic identity. Although speculative, this scenario of ethnogenesis among at least some members of Chen Chen– and Omo-style ethnic communities could explain the eventual disappearance of Chen Chen– and Omostyle contexts in the Moquegua Valley coincident with the emergence of Tumilaca-style archaeological contexts (Goldstein 1985) and the decline of Tiwanaku state influence in the region during the late Middle Horizon (Goldstein 2005; Owen 2005; Sutter and Sharratt 2010; Williams 2002). The results also suggest a degree of variation in marriage practices among Moquegua Tiwanaku-affiliated communities. As reported, Omostyle agropastoralists at Omo Alto M16 and Rio Muerto M70 had higher

42 · Kent M. Johnson

rates of extra-local exogamy compared to their Chen Chen–style counterparts. The higher rates of transhumance among individuals from the Omo-style context of Rio Muerto M70 identified by Knudson and colleagues (2014) may have contributed to geographically expansive marriage networks in Omo-style communities. Further, there appear to be different marriage practices within Omo-style sites, as the results of canonical variates analysis and social network analysis suggest individuals from the Omo-style site of Rio Muerto M70 married individuals from Chen Chen–style sites in the Moquegua Valley, but the limited evidence from Omo Alto M16 indicates members of this Omo-style community married outside of the Moquegua Valley. Cranial modification results indicate members of both Moquegua Tiwanaku ethnic communities shaped their children’s heads to produce the same modified cranial style (fronto-occipital). The uniformity in culturally derived cranial shape among Chen Chen– and Omo-style communities in Moquegua stands in stark contrast to the heterogeneous cranial shapes produced by culturally diverse communities in other Tiwanakuaffiliated peripheral regions. O’Brien (2003) and Torres-Rouff (2002, 2008) documented both annular and fronto-occipital modified crania in Middle Horizon period Tiwanaku-affiliated skeletal samples from highland sites in Cochabamba, Bolivia, and San Pedro de Atacama, Chile, respectively. Why did Moquegua Valley Tiwanaku ethnic communities produce a single culturally modified cranial shape when Tiwanaku-affiliated communities in other peripheral regions produced multiple cranial shapes? It may be that cranial modification style was not part of the suite of traits used to signal ethnic affiliation among Moquegua Valley Tiwanaku communities. As discussed above, cranial modification does not have any inherent social or cultural meaning (Sutter 2005). Its significance to social organization and cultural identity vary in different contexts, and although it may covary with material culture indicators of ethnic identity in some contexts (e.g., Lozada Cerna 1998), we should not assume that it does in all contexts (Stovel 2013). Alternatively, Chen Chen– and Omo-style communities may have embodied a shared regional cultural identity (i.e., Moquegua Tiwanaku) vis-à-vis some “other” via their culturally determined cranial shape. If this is the case, from whom did they seek to differentiate themselves? Moquegua Tiwanaku colonists did not use cultural cranial shape to signal difference from indigenous Huaracane communities who preceded

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 43

them in the middle Moquegua Valley; Huaracane samples also exhibit fronto-occipital cranial modification (Blom 1999). It is possible that Tiwanaku colonists used culturally modified cranial shape to signal their difference from Wari colonists in the upper Moquegua Valley (Moseley et al. 1991; Williams 2001), but there are no comparative Wari skeletal samples from Moquegua Valley sites available to evaluate this prediction.6 Future research emphasizing a diachronic approach to cranial modification practices in Moquegua Valley Tiwanaku skeletal samples is necessary to evaluate these alternative explanations. The evidence for gene flow among Chen Chen– and Omo-style samples is compelling in light of recent interpretations of hybrid Chen Chen–style and Omo-style cemeteries from two sites in the Moquegua Valley. Early archaeological surveys and excavation yielded little evidence of shared use of residential and mortuary sites (Goldstein 1989, 2005; cf. Lewis 2005), but recent excavations at Rio Muerto M43A and Omo M10R have identified potential shared-use cemeteries based on cemetery location and the type and quantity of grave goods recovered (Baitzel 2016). Future research will explore phenotypic variation within these hybrid cemeteries in greater detail. There is additional evidence of commonality and shared belonging among Chen Chen– and Omo-style communities that acknowledges difference but also marks collective affiliation and identity. Although in this chapter I have emphasized stylistic differences in ceramics as a reflection of underlying expression of ethnic identity, makers of both Omoand Chen Chen–style ceramics expressed their shared Tiwanaku cultural heritage through common decorative motifs and vessel forms, notably the kero used for drinking maize beer (Goldstein 1989, 2005; Owen 2005). A hybrid vessel from Chen Chen M1 with a polished black-ware, Omo-style interior and a red-slipped, Chen Chen–style exterior presents compelling, albeit limited, evidence of the integration of their disparate material culture styles (Sharratt 2011). In combination, analysis of phenotypic and cranial modification data present compelling evidence for similarity and interaction among members of distinct Tiwanaku ethnic communities. Together with recent studies of paleomobility, activity patterns, and dietary practices, bioarchaeological studies are generating a more nuanced understanding of social relationships and cultural practices among Tiwanaku colonial communities in Moquegua. Despite their outward cultural differences, individuals

44 · Kent M. Johnson

from Chen Chen–style and Omo-style communities forged a common Moquegua Tiwanaku identity. They did so by intermarrying, modifying their children’s heads to have a similar shape, eating similar foods and/or sharing foods, and burying their dead in some of the same cemeteries.7 Overall, this pattern of communities with different material culture styles and distinct subsistence practices who otherwise maintain close biosocial interactions is consistent with Rostworowski’s (1977a, 1977b, 1978) classic horizontal model of Andean interaction. I have suggested that interethnic marriage within families was crucial to integrating Chen Chen– and Omo-style ethnic communities. That these family networks are identified based solely on phenotypic similarity as a proxy for biological relatedness is problematic for two reasons. First, studies of skeletal samples with known genealogies attest to the difficulty of differentiating close biological relatives and nonrelatives using phenotypic data (e.g., Paul and Stojanowski 2015, 2017; Stojanowski and Hubbard 2017). Second, biological relatedness may be a poor and biased proxy for reconstructing the social dimensions of family organization (Johnson 2019; Johnson and Paul 2016). Andean ethnographic and ethnohistoric research suggests the criteria for kin group membership are flexible and not limited to biological relatedness (e.g., Abercrombie 1998; Bastien 1978; Harris 1978; Isbell 1978; Platt 1982; Rasnake 1988). To explore social dimensions of relatedness among potential kin groups identified using social network analysis, I evaluated whether cranial modification presence and patterns of paleomobility vary between components 1 and 2. There are thirty-six individuals in component 1, and six of these are unobservable for cranial modification. Of the remaining twenty-seven individuals, three are nonmodified (M1 3677, M10 85–8, and M10 85–25[B]) and twenty-seven have fronto-occipital modification. The three members of component 2 all have fronto-occipital modification. Fisher’s exact test (p = 0.999) indicates no significant difference between individuals from components 1 and 2 in terms of cranial modification presence. Published radiogenic strontium values are available for a sample of individuals from Chen Chen M1 (Knudson and Price 2007) and Rio Muerto M43 and M70 (Knudson et al. 2014), including nine individuals present in components 1 and 2. For component 1, five individuals (M1 0016, M1 2068, M1 3768, M70 2787, and M70 2985), all of whom are adult females, have local signatures, and one adult male (M70 2840) has an intermediate signature indicating he may have moved regularly between different geologic

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 45

zones as an adult. Based on the specific values derived, it is likely he traveled repeatedly between the Tiwanaku heartland and the Moquegua Valley (Knudson et al. 2014). Isotopic data are available for all three individuals in component 2. Two of these individuals (M43B 3054 and M70 4443) are adult males with local signatures, whereas the third member of component 2 is an adult female (M70 4468) with nonlocal signatures in her dental tissues. She likely represents a first-generation immigrant to the Moquegua Valley (Knudson et al. 2014). There is no clear distinction between potential family groups in terms of mobility patterns. Each family network is comprised of a majority of individuals with local isotopic signatures and includes at least one individual who moved between geologic zones at some point in their life. More broadly, if analysis of phenotypic data has identified close biological relatives, then family networks not only crossed site and ethnic community boundaries within the Moquegua Valley, they spanned different ecological zones. A pattern of dispersed kin and community groups is consistent with the vertical archipelago model of Andean socio-, economic-, and political organization described by John Murra and others (e.g., Goldstein 2000, 2005; Harris 1978; Murra 1968, 1972, 1985). Thus, social life within the Tiwanaku-affiliated communities in the Moquegua Valley exhibits elements of both horizontal and vertical organization. Conclusion

Social organization within Middle Horizon period Tiwanaku communities in the Moquegua Valley was influenced by diverse, cross-cutting affiliations including family-, ethnic-, and region-level affiliations. The findings presented here highlight the need to investigate the ways in which midlevel scales of social affiliation structured the social fabric of past communities. Analysis of phenotypic variation and cranial modification at multiple scales, including intermediate levels of affiliation, allowed for detection of inter-and intraethnic community variability in marriage practices and family composition. There is no question that ethnic affiliations substantially influenced sociality within Tiwanaku-affiliated communities in the Moquegua Valley. For several hundred years, members of these two seemingly distinct ethnic communities lived alongside one another peacefully. Despite outward

46 · Kent M. Johnson

expressions of cultural difference, they created a collective Moquegua Tiwanaku identity and sense of belonging as Tiwanaku colonists living far from their ancestral homelands through intermarriage, shared use of sites, and culturally determined head shape. Acknowledgments

My sincere thanks to the editors, Chris and Kelly, for inviting me to contribute to this volume. Their first edited volume on this subject was published during my time in graduate school, and it had a formative influence on my scholarship. Their comments on my chapter were insightful and helped strengthen it immeasurably. Chris, Kelly, Jane Buikstra, and Paul Goldstein provided valuable feedback on an earlier version of this project. I also thank the anonymous reviewers, whose feedback greatly improved the quality of the finished chapter. I would like to acknowledge the many people and institutions who facilitated the research presented here. My sincerest thanks to the current and former staff at Museo Contisuyo in Moquegua, Peru, for providing access and support during data collection, especially Director Patricia Palacios Filinich, former Director Antonio Oquiche, and Yamilex Tejada Alvarez. Luis (Lucho) Gonzales Peñaranda of la Dirección Desconcentrada de Cultura de Moquegua was immensely helpful in providing requested information and approving paperwork. I would like to acknowledge the following individuals for granting permission to access the skeletal collections used in this research: Sara Baitzel, Jane Buikstra, Paul Goldstein, Mike Moseley, Bruce Owen, Antonio Oquiche, Paty Palacios, Romulo Pari Flores, Nicola Sharratt, Bertha Vargas, and Ryan Williams. Ryan Williams and the Programa Collesuyo provided much-needed institutional affiliation and offered an expedited route to securing permission to conduct the research presented here. I thank Sofia Chacaltana for her diligent work as co-director, along with Jane Buikstra, in overseeing the paperwork that helped secure permission. Sara Marsteller and Martha Palma Malaga provided invaluable counsel on procedural matters related to conducting research in Peru, for which I will be forever grateful. Financial support was provided by the National Science Foundation Doctoral Dissertation Improvement Grant No. 1441894 and a Dissertation Completion Fellowship by the School of Human Evolution and Social Change at Arizona State University.

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 47

Notes 1. In her dissertation, Sarah Baitzel (2016: 121) clarifies the multiple archaeological usages of the terms “Omo” and “Chen Chen.” These include (1) location, (2) style, and (3) chronology. In terms of location, Omo and Chen Chen refer to specific archaeological site groups (i.e., clusters of sites) within the Moquegua Valley, Peru. “Omo-style” and “Chen Chen–style” are used to designate stylistic and formal differences in Tiwanaku ceramic assemblages first identified in the Moquegua Valley at the type sites of Omo M12 and Chen Chen M1, respectively (Goldstein 1985). Subsequently, these terms are also extended to contexts and sites in which those artifacts and assemblages are found (e.g., the Omo-style site of Rio Muerto M70) as well as the people who occupied and used those contexts and sites (e.g., the Omo-style community or population of Rio Muerto M70). Thus, “Omo-style” and “Chen Chen–style” are used as ethnonyms in the archaeological literature. In terms of chronology, the two styles initially were thought to represent discrete occupation phases (e.g., the “Omo phase” and the “Chen Chen phase”), but they are now recognized as overlapping but nonsynchronous occupations. The chronological use of the terms has waned. 2. Although it is problematic to include individuals from possible hybrid Chen Chen–style and Omo-style mortuary contexts with individuals from Chen Chen–style contexts in the sample-based analyses, this is necessary due to the small sample sizes of individuals from the hybrid cemetery sector at Omo M10R (n = 2) and the hybrid (M43A, n = 4) and Chen Chen–style (M43B, n = 2) sectors at Rio Muerto M43. 3. See Johnson (2016) for a detailed description of analytical procedures. 4. The smallest 1 percent of pairwise D2 distances were coded as relationship present (1) and the remaining 99 percent were coded as relationship absent (0). This conservative breakpoint establishes potential kinship ties between only the most phenotypically similar individuals in the study sample. 5. Centrality refers to a node’s structural position within a graph in terms of the number of its connections (Scott 2013). Degree centrality is a count of a node’s connections (Hanneman and Riddle 2005); here, it indicates the number of close biological relatives a node has within the study sample. 6. Wari-affiliated samples from different regions exhibit different modification styles. Fronto-occipital modification is documented at the coastal site of Beringa, whereas annular style is reported for the highland site of Conchopata (Tung 2003). 7. To be clear, this statement is not generalizable to the Omo-style sample from Omo Alto M16. Individuals from M16 were included in Becker and Goldstein’s analysis of osteoarthritis and activity patterns, but, to date, samples from this site have not been included in published isotopic studies of paleodiet and paleomobility.

Works Cited Abercrombie, Thomas A. 1998. Pathways of Memory and Power: Ethnography and History among an Andean People. Madison: University of Wisconsin Press.

48 · Kent M. Johnson

Adams, Dean C., F. James Rohlf, and Dennis E. Slice. 2004. “Geometric Morphometrics: Ten Years of Progress after the ‘Revolution.’” Italian Journal of Zoology 71(1): 5–16. Albarracín-Jordán, Juan. 1996. “Tiwanaku Settlement System: The Integration of Nested Hierarchies in the Lower Tiwanaku Valley.” Latin American Antiquity 7(3): 183–210. Albarracín-Jordán, Juan. 2003. “Tiwanaku: A pre-Inca, Segmentary State in the Andes.” In Tiwanaku and Its Hinterland: Archaeology and Paleoecology of an Andean Civilization, Vol. 2, edited by Alan L. Kolata, 95–111. Washington, DC: Smithsonian Institution Press. Alt, Kurt W., and Werner Vach. 1998. “Kinship Studies in Skeletal Remains: Concepts and Examples.” In Dental Anthropology: Fundamentals, Limits, and Prospects, edited by Kurt W. Alt, Freidrich W. Rösing, and Maria Teschler-Nicola, 537–554. New York: Springer. Anderson, Karen. 2013. “Tiwanaku Influence in the Central Valley of Cochabamba.” In Visions of Tiwanaku, Monograph 78, edited by Alexei Vranich and Charles Stanish, 87–112. Los Angeles: Cotsen Institute of Archaeology Press. Baitzel, Sarah I. 2008. “No Country for Old People: A Paleodemographic Study of Tiwanaku Return Migration in Moquegua, Peru.” Master’s thesis, University of California, San Diego, Department of Anthropology. Baitzel, Sarah I. 2016. “The Politics of Death and Identity in Provincial Tiwanaku Society (AD 600–1100).” Doctoral dissertation, University of California, San Diego, Department of Anthropology. Barth, Fredrick. 1969. “Introduction.” In Ethnic Groups and Boundaries: The Social Organization of Difference, edited by Fredrick Barth, 9–38. Long Grove, IL: Waveland Press. Bastien, Joseph W. 1978. Mountain of the Condor: Metaphor and Ritual in an Andean Ayllu. New York: West Publishing. Becker, S.K., and P.S. Goldstein. 2018. “Evidence of Osteoarthritis in the Tiwanaku Colony, Moquegua, Peru (AD 500–100).” International Journal of Osteoarchaeology 28(1): 54–64. Bermann, Marc. 1997. “Domestic Life and Vertical Integration in the Tiwanaku Heartland.” Latin American Antiquity 8(2): 93–112. Berryman, Carrie A. 2010. “Food, Feasts, and the Construction of Identity and Power in Ancient Tiwanaku: A Bioarchaeological Perspective.” Doctoral dissertation, Vanderbilt University, Department of Anthropology. Blom, Deborah E. 1999. “Tiwanaku Regional Interaction and Social Identity: A Bioarchaeological Approach.” Doctoral dissertation, University of Chicago, Department of Anthropology. Blom, Deborah E. 2005a. “Embodying Borders: Human Body Modification and Diversity in Tiwanaku Society.” Journal of Anthropological Archaeology 24(1): 1–24. Blom, Deborah E. 2005b. “A Bioarchaeological Approach to Tiwanaku Group Dynamics.” In Us and Them: Archaeology and Ethnicity in the Andes, Monograph 53, edited by Richard M. Reycraft, 153–182. Los Angeles: Cotsen Institute of Archaeology at UCLA. Blom, Deborah E., Benedikt Hallgrímsson, Linda Keng, Maria C. Lozada C., and Jane E. Buikstra. 1998. “Tiwanaku ‘Colonization’: Bioarchaeological Implications for Migration in the Moquegua Valley, Peru.” World Archaeology 30(2): 238–261. Borgatti, Stephen P. 2002. NetDraw Software for Network Visualization. Lexington, KY: Analytic Technologies.

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 49

Borgatti, Stephen P., Martin G. Everett, and Lin C. Freeman. 2002. Ucinet for Windows: Software for Social Network Analysis. Cambridge, MA: Harvard Analytic Technologies. Borgatti, Stephen P., Martin G. Everett, and Jeffrey C. Johnson. 2013. Analyzing Social Networks. Los Angeles: Sage Publications. Boutin, Alexis T. 2012. “Written in Stone, Written in Bone: The Osteobiography of a Bronze Age Craftsman from Alalakh.” In The Bioarchaeology of Individuals, edited by Ann L.W. Stodder and Ann M. Palkovich, 193–214. Gainesville: University Press of Florida. Browman, David L. 1978. “Toward the Development of the Tiwanaku (Tiahuanaco) State.” In Advances in Andean Archaeology, edited by David L. Browman, 327–349. The Hague: Mouton. Browman, David L. 1997. “Political Institutional Factors Contributing to the Integration of the Tiwanaku State.” In Emergence and Change in Early Urban Societies, edited by Linda Manzanilla, 229–243. New York: Plenum. Buikstra, Jane E. 1995. “Tombs for the Living . . . or . . . for the Dead: The Osmore Ancestors.” In Tombs for the Living: Andean Mortuary Practices, edited by Tom D. Dillehay, 229–279. Washington, DC: Dumbarton Oaks. Buikstra, Jane E., Susan R. Frankenberg, and Lyle W. Konigsberg. 1990. “Skeletal Biological Distance Studies in American Physical Anthropology: Recent Trends.” American Journal of Physical Anthropology 82(1): 1–7. Cannings, C., and M. H. Skolnick. 1975. “Genetic Drift in Exogamous Marriage Systems.” Theoretical Population Biology 7(1): 39–54. Chapman, Malcolm. 1993. “Social and Biological Aspects of Ethnicity.” In Social and Biological Aspects of Ethnicity, edited by Malcolm Chapman, 1–46. Oxford: Oxford University Press. Cheverud, James M. 1988. “A Comparison of Genetic and Phenotypic Correlations.” Evolution 42(7): 958–968. Cieza de León, Pedro de. 1984 [1553]. La Crónica del Perú, Obras Completos, 3 partes. “Monumenta Hispano-Indiana V Centenario del Descubriemiento de America II.” Madrid: Consejo Superior de Investigaciones Cientificas, Instituto “Gonzalo Fernández de Ovidedo.” Cornelison, Jered B., Wendy Lackey-Cornelison, and Lynne Goldstein. 2017. “Contextual and Biological Markers of Community Identity in the Effigy Mound Manifestation of Southern Wisconsin.” In The Bioarchaeology of Community, edited by Sara L. Juengst and Sara K. Becker. Archeological Papers of the American Anthropological Association 28(1): 66–81. de la Vega, el Inca Garcilaso. 1966 [1609]. Royal Commentaries of the Incas and General History of Peru. Austin: University of Texas Press. Epstein, A.L. 1978. Ethos and Identity: Three Studies in Ethnicity. London: Tavistock. Fix, Alan G. 1979. “Anthropological Genetics of Small Populations.” Annual Review of Anthropology 8(1): 207–230. Goldstein, Paul S. 1985. “Tiwanaku Ceramics of the Moquegua Valley, Peru.” Master’s thesis, University of Chicago. Goldstein, Paul S. 1989. “Omo: A Tiwanaku Provincial Center in Moquegua, Peru.” Doctoral dissertation, University of Chicago, Department of Anthropology.

50 · Kent M. Johnson

Goldstein, Paul S. 1993. “Tiwanaku Temples and State Expansion: A Tiwanaku SunkenCourt Temple in Moquegua, Peru.” Latin American Antiquity 4(1): 22–47. Goldstein, Paul S. 2000. “Communities without Borders: The Vertical Archipelago and Diaspora Communities in the Southern American Andes.” In The Archaeology of Communities: A New World Perspective, edited by Marcello A. Canuto and Jason Yaeger, 182–209. New York: Routledge. Goldstein, Paul S. 2003. “From Stew-Eaters to Maize-Drinkers: The Chicha Economy and the Tiwanaku Expansion.” In The Archaeology and Politics of Food and Feasting in Early States and Empires, edited by Tamara L. Bray, 143–172. New York: Springer. Goldstein, Paul S. 2005. Andean Diaspora: The Tiwanaku Colonies and the Origins of South American Empire. Gainesville: University Press of Florida. Goldstein, Paul S. 2015. “Multiethnicity, Pluralism, and Migration in the South Central Andes: An Alternate Path to State Expansion.” Proceedings of the National Academy of Sciences USA 112(30):9202–9209. Hanneman, Robert A., and Mark Riddle. 2005. Introduction to Social Network Methods. Riverside: University of California. http://faculty.ucr.edu/~hanneman Harpending, Henry C., and R. Ward. 1982. “Chemical Systematics and Human Evolution. In Biochemical Aspects of Evolutionary Biology, edited by Matthew H. Nitecki, 213–256. Chicago: University of Chicago Press. Harris, Olivia. 1978. “El Parentesco y la Economía Vertical en el Ayllu Laymi (Norte de Potosi).” Avances: Revista Boliviana de Estudios Históricos y Sociales 1: 51–64. Hausman, Alice J. 1984. “Holocene Human Evolution in Southern Africa.” In From Hunters to Farmers: The Causes and Consequences of Food Production in Africa, edited by J. Desmond Clark and Steven A. Brandt, 261–271. Berkeley: University of California Press. Hawkey, Diane E. 1998. “Disability, Compassion and the Skeletal Record: Using Musculoskeletal Stress Markers (MSM) to Construct an Osteobiography from Early New Mexico.” International Journal of Osteoarchaeology 8(5): 326–340. Hoshower, Lisa M., Jane E. Buikstra, Paul S. Goldstein, and Ann D. Webster. 1995. “Artificial Cranial Deformation at the Omo M10 Site: A Tiwanaku Complex from the Moquegua Valley, Peru.” Latin American Antiquity 6(2): 145–164. Isbell, Billie Jean. 1978. To Defend Ourselves: Ecology and Ritual in an Andean Village. Prospect Heights, IL: Waveland Press. Janusek, John W. 2004. Identity and Power in the Ancient Andes: Tiwanaku Cities through Time. New York: Routledge. Janusek, John W. 2005. “Of Pots and People: Ceramic Style and Social Identity in the Tiwanaku State.” In Us and Them: Archaeology and Ethnicity in the Andes, edited by Richard M. Reycraft, 34–51. Los Angeles: Cotsen Institute of Archaeology at UCLA. Janusek, John W. 2008. Ancient Tiwanaku. Cambridge: Cambridge University Press. Jenkins, Richard. 2008. Social Identity, 3rd ed. New York: Routledge. Johnson, Kent M. 2016. “Ethnicity, Family, and Social Networks: A Multiscalar Bioarchaeological Investigation of Tiwanaku Colonial Organization in the Moquegua Valley, Peru.” Doctoral dissertation, Arizona State University, School of Human Evolution and Social Change. Johnson, Kent M. 2019. “Opening up the Family Tree: Promoting More Diverse and

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 51

Inclusive Studies of Family, Kinship, and Relatedness in Bioarchaeology.” In Bioarchaeologists Speak Out: Deep Time Perspectives on Contemporary Issues, edited by Jane E. Buikstra, 201–230. New York: Springer. Johnson, Kent M., and Kathleen S. Paul. 2016. “Bioarchaeology and Kinship: Integrating Theory, Social Relatedness, and Biology in Ancient Family Research.” Journal of Archaeological Research 24(1): 75–123. Jones, Siân. 1997. The Archaeology of Ethnicity: Constructing Identities in the Past and Present. London: Routledge. Klingenberg, Christian P. 2011. “MorphoJ: An Integrated Software Package for Geometric Morphometrics.” Molecular Ecology Resources 11(2): 353–357. Knudson, Kelly J., Paul S. Goldstein, Allisen Dahlstedt, Andrew Somerville, and Margaret J. Schoeninger. 2014. “Paleomobility in the Tiwanaku Diaspora: Biogeochemical Analyses at Rio Muerto, Moquegua, Peru.” American Journal of Physical Anthropology 155(3): 405–421. Knudson, Kelly J., and T. Douglas Price. 2007. “Utility of Multiple Chemical Techniques in Archaeological Residential Mobility Studies: Case Studies from Tiwanaku-and Chiribaya-Affiliated Sites in the Andes.” American Journal of Physical Anthropology 132(1): 25–39. Knudson, Kelly J., and Christopher M. Stojanowski. 2009. “The Bioarchaeology of Identity.” In Bioarchaeology and Identity in the Americas, edited by Kelly J. Knudson and Christopher M. Stojanowski, 1–23. Gainesville: University Press of Florida. Kolata, Alan L. 1992. “Economy, Ideology, and Imperialism in the South-central Andes.” In Ideology and Pre-Columbian Civilizations, edited by Arthur A. Demarest and Geoffrey W. Conrad, 65–85. Santa Fe, NM: School of American Research Press. Kolata, Alan L. 1993a. The Tiwanaku: Portrait of an Andean Civilization. Cambridge, MA: Blackwell. Kolata, Alan L. 1993b. “Understanding Tiwanaku: Conquest, Colonization, and Clientage in the South Central Andes.” In Latin American Horizons, edited by Don S. Rice, 193–224. Washington, DC: Dumbarton Oaks. Komar, Debra A., and Jane E. Buikstra. 2008. Forensic Anthropology: Contemporary Theory and Practice. Oxford: Oxford University Press. Konigsberg, Lyle W., and Stephen D. Ousley. 1995. “Multivariate Quantitative Genetics of Anthropometric Traits from the Boas Data.” Human Biology 67(1): 481–498. Larsen, Clark S. 2015. Bioarchaeology: Interpreting Behavior from the Human Skeleton, 2nd ed. Cambridge: Cambridge University Press. Lewis, Cecil M., Jr. 2005. “Intercontinental to Intrasite Genetic Analyses of Ancient and Contemporary Native American Communities.” Doctoral dissertation, University of New Mexico, Department of Anthropology. Lozada Cerna, María Cecilia. 1998. “The Señorío of Chiribaya: A Bioarchaeological Study in the Osmore Drainage of Southern Perú.” Doctoral dissertation, University of Chicago, Department of Anthropology. Lucy, Sam. 2005. “Ethnic and Cultural Identities.” In The Archaeology of Identity: Approaches to Gender, Age, Status, Ethnicity and Religion, edited by Margarita DíazAndreu, Sam Lucy, Staša Babić, and David N. Edwards, 86–109. London: Routledge.

52 · Kent M. Johnson

Magilligan, Francis J., and Paul S. Goldstein. 2001. “El Niño Floods and Culture Change: A Late Holocene Flood History for the Rio Moquegua, Southern Peru.” Geology 29(5): 431–434. Mantel, Nathan. 1967. “The Detection of Disease Clustering and a Generalized Regression Approach.” Cancer Research 27(2): 209–220. Marsteller, Sara J. 2015. “Community Identity and Social Diversity on the Central Peruvian Coast: A Bioarchaeological Investigation of Ychsma Diet, Mobility, and Mortuary Practices (c. AD 900–1470).” Doctoral dissertation, Arizona State University, School of Human Evolution and Social Change. McAndrews, Timothy L., Juan Albarracín-Jordán, and Marc Bermann. 1997. “Regional Settlement Patterns in the Tiwanaku Valley of Bolivia.” Journal of Field Archaeology 24(1): 67–83. McKeown, Ashley H., and Richard L. Jantz. 2005. “Comparison of Coordinate and Craniometric Data for Biological Distance Studies.” In Modern Morphometrics in Physical Anthropology, edited by Dennis E. Slice, 215–230. New York: Kluwer Academic/ Plenum Publishers. Meskell, Lynn. 2007. “Archaeologies of Identity.” In The Archaeology of Identities: A Reader, edited by Timothy Insoll, 23–43. London: Routledge. Meyer, Christian, Robert Ganslmeier, Veit Dresely, and Kurt W. Alt. 2012. “New Approaches to the Reconstruction of Kinship and Social Structure Based on Bioarchaeological Analysis of Neolithic Multiple and Collective Graves.” In Theoretical and Methodological Considerations in Central European Neolithic Archaeology. BAR International Series, 2325, edited by Jan Kolár and Frantisek Trampota, 11–23. Oxford: Archaeopress. Miller, Alicia V. 2013. “Social Organization and Interaction in Bronze Age Eurasia: A Bioarchaeological and Statistical Approach to the Study of Communities.” Doctoral dissertation, University of Pittsburgh, Department of Anthropology. Morton, N.E., Shirley Yee, D.E. Harris, and Ruth Lew. 1971. “Bioassay of Kinship.” Theoretical Population Biology 2(4): 507–524. Moseley, Michael E., Robert A. Feldman, Paul S. Goldstein, and Luis Watanabe. 1991. “Colonies and Conquest: Tiahuanaco and Huari in Moquegua.” In Huari Administrative Structure: Prehistoric Monumental Architecture and State Government, edited by William H. Isbell and Gordon F. McEwan, 121–140. Washington, DC: Dumbarton Oaks. Murra, John V. 1968. “An Aymara Kingdom in 1567.” Ethnohistory 15(2): 115–151. Murra, John V. 1972. “El ‘Control Vertical’ de un Máximo de Pisos Ecológicos en la Economía de las Sociedades Andinas.” In Visita de la Provincia de León de Huánuco en 1562 por Iñigo Ortiz de Zuñiga. Documentos para la Historia y Etnología de Huánuco y la Selva Central, Vol. 2, edited by John V. Murra, 427–476. Huánuco: Universidad Nacional Hermilio Valdizan. Murra, John V. 1985. “‘El Archipélago Vertical’ Revisited.” In Andean Ecology and Civilization: An Interdisciplinary Perspective on Andean Ecological Complementarity, edited by Shodo Masuda, Izumi Shimada, and Craig Morris, 3–13. Tokyo: University of Tokyo. Oakland Rodman, Amy 1992. “Textiles and Ethnicity: Tiwanaku in San Pedro de Atacama, North Chile.” Latin American Antiquity 3(4): 316–340.

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 53

O’Brien, Tyler G. 2003. “Cranial Microvariation in Prehistoric South Central Andean Populations: An Assessment of Morphology in the Cochabamba Collection, Bolivia.” Doctoral dissertation, Binghamton University, State University of New York, Department of Anthropology. Owen, Bruce D. 2005. “Distant Colonies and Explosive Collapse: The Two Stages of the Tiwanaku Diaspora in the Osmore Drainage.” Latin American Antiquity 16(1): 45–80. Paul, Kathleen S., and Christopher M. Stojanowski. 2015. “Performance Analysis of Deciduous Morphology for Detecting Biological Siblings.” American Journal of Physical Anthropology 157(4): 615–629. Paul, Kathleen S., and Christopher M. Stojanowski. 2017. “Comparative Performance of Deciduous and Permanent Dental Morphology in Detecting Biological Relatives.” American Journal of Physical Anthropology 164(1): 97–116. Pilloud, Marin A. 2009. “Community Structure at Neolithic Çatalhöyük: Biological Distance Analysis of Household, Neighborhood, and Settlement.” Doctoral dissertation, The Ohio State University, Department of Anthropology. Platt, Tristan. 1982. Estado Boliviano y Ayllu Andino: Tierra y Tributo en el Norte de Potosí. Lima: Instituto de Estudios Peruanos. Rasnake, Roger N. 1988. Domination and Cultural Resistance: Authority and Power among Andean People. Durham, NC: Duke University Press. Read, Dwight, and Sander van der Leeuw. 2015. “The Extension of Social Relations in Time and Space during the Palaeolithic.” In Settlement, Society and Cognition in Human Evolution, edited by Fiona Coward, Robert Hosfield, Matt Pope, and Francis Wenban-Smith, 31–53. Cambridge: Cambridge University Press. Relethford, John H. 2003. “Anthropometric Data and Population History.” In Human Biologists in the Archives: Demography, Health, Nutrition, and Genetics in Historical Populations, edited by D. Ann Herring and Alan C. Swedlund, 32–52. Cambridge: Cambridge University Press. Relethford, John H. 2010. “Population-specific Deviations of Global Human Craniometric Variation from a Neutral Model.” American Journal of Physical Anthropology 142(1): 105–111. Relethford, John H., and John Blangero. 1990. “Detection of Differential Gene Flow from Patterns of Quantitative Variation.” Human Biology 62(1): 5–25. Relethford, John H., Michael H. Crawford, and John Blangero. 1997. “Genetic Drift and Gene Flow in Post-Famine Ireland.” Human Biology 69(4): 443–465. Robb, John. 2002. “Time and Biography: Osteobiography of the Italian Neolithic Lifespan.” In Thinking through the Body: Archaeologies of Corporeality, edited by Yannis Hamilakis, Mark Pluciennik, and Sarah Tarlow, 153–171. London: Kluwer Academic/Plenum. Rostworowski, María. 1977a. “Coastal Fishermen, Merchants, and Artisans in Pre-Hispanic Peru.” In The Sea in the pre-Columbian World, edited by Elizabeth P. Benson, 167–186. Washington, DC: Dumbarton Oaks. Rostworowski, María. 1977b. Etnía y Sociedad: Costa Peruana Prehispánica. Lima: Instituto de Estudios Peruanos. Rostworowski, María. 1978. Señoríos Indígenas de Lima y Canta. Lima: Instituto de Estudios Peruanos.

54 · Kent M. Johnson

Saul, Frank P. 1972. “The Human Skeletal Remains from Altar de Sacrificios, Guatemala: An Osteobiographic Analysis.” Papers of the Peabody Museum 63(2): 1–123. Saul, Frank P., and Julie M. Saul. 1989. “Osteobiography: A Maya Example.” In Reconstruction of Life from the Skeleton, edited by Mehmet Yaşar İşcan and Kenneth A.R. Kennedy, 287–302. New York: Wiley-Liss. Scott, John. 2013. Social Network Analysis, 3rd ed. Los Angeles: Sage Publications. Sharratt, Nicola. 2011. “Social Identities and State Collapse: A Diachronic Study of Tiwanaku Burials in the Moquegua Valley, Peru.” Doctoral dissertation, University of Illinois at Chicago, Department of Anthropology. Slice, Dennis E. 2005. “Modern Morphometrics.” In Modern Morphometrics in Physical Anthropology, edited by Dennis E. Slice, 1–45. New York: Kluwer Academic / Plenum Publishers. Slice, Dennis E. 2013. Morpheus et al., Java Edition. Computer software. Smouse, Peter E., Jeffrey C. Long, and Robert R. Sokal. 1986. “Multiple Regression and Correlation Extensions of the Mantel Test of Matrix Correspondence.” Systemic Zoology 35(4): 627–632. Somerville, Andrew D., Paul S. Goldstein, Sarah I. Baitzel, Karin L. Bruwelheide, Allisen C. Dahlstedt, Linda Yzurdiaga, Sarah Raubenheimer, Kelly J. Knudson, and Margaret J. Schoeninger. 2015. “Diet and Gender in the Tiwanaku Colonies: Stable Isotope Analysis of Human Bone Collagen and Apatite from Moquegua, Peru.” American Journal of Physical Anthropology 158(3): 408–422. Spence, Michael W. 1974a. “Residential Practices and the Distribution of Skeletal Traits in Teotihuacan, Mexico.” Man 9(2): 262–273. Spence, Michael W. 1974b. “The Study of Residential Practices among Prehistoric Hunters and Gatherers.” World Archaeology 5(3): 346–357. Stanish, Charles. 2002. “Tiwanaku Political Economy.” In Andean Archaeology I: Variations in Sociopolitical Organization, edited by William H. Isbell and Helaine Silverman, 169–198. New York: Kluwer Academic. Stanish, Charles. 2003. Ancient Titicaca: The Evolution of Complex Society in Southern Peru and Northern Bolivia. Berkeley: University of California Press. Stodder, Ann L.W., and Ann M. Palkovich, eds. 2012. The Bioarchaeology of Individuals. Gainesville: University Press of Florida. Stojanowski, Christopher M. 2010. Bioarchaeology of Ethnogenesis in the Colonial Southeast. Gainesville: University Press of Florida. Stojanowski, Christopher M. 2013. Mission Cemeteries, Mission Peoples: Historical and Evolutionary Dimensions of Intracemetery Bioarchaeology in Spanish Florida. Gainesville: University Press of Florida. Stojanowski, Christopher M., and William N. Duncan, eds. 2017. Studies in Forensic Biohistory: Anthropological Perspectives. Cambridge: Cambridge University Press. Stojanowski, C.M., and A.R. Hubbard. 2017. “Sensitivity of Dental Phenotypic Data for the Identification of Biological Relatives.” International Journal of Osteoarchaeology 27(5): 813–827. Stovel, Emily M. 2013. “Concepts of Ethnicity and Culture in Andean Archaeology.” Latin American Antiquity 24(1): 3–20.

Exploring Identities among Tiwanaku-Affiliated Communities in Moquegua, Peru · 55

Sutter, Richard C. 2005. “A Bioarchaeological Assessment of Prehistoric Ethnicity among Early Late Intermediate Period Populations of the Azapa Valley, Chile.” In Us and Them: Archaeology and Ethnicity in the Andes, Monograph 53, edited by Richard M. Reycraft, 183–205. Los Angeles: Cotsen Institute of Archaeology at UCLA. Sutter, Richard C., and Nicola Sharratt. 2010. “Continuity and Transformation during the Terminal Middle Horizon (AD 950–1150): A Bioarchaeological Assessment of Tumilaca Origins within the Middle Moquegua Valley, Peru.” Latin American Antiquity 21(1): 67–86. Tomczak, Paula D. 2001. “Prehistoric Socio-economic Relations and Population Organization in the Lower Osmore Valley of Southern Peru.” Doctoral dissertation, University of New Mexico, Department of Anthropology. Torres-Rouff, Christina. 2002. “Cranial Vault Modification and Ethnicity in Middle Horizon San Pedro de Atacama, Chile.” Current Anthropology 43(1): 163–171. Torres-Rouff, Christina. 2008. “The Influence of Tiwanaku on Life in the Chilean Atacama: Mortuary and Bodily Perspectives.” American Anthropologist 110(3): 325–337. Torres-Rouff, Christina, and Kelly J. Knudson. 2017. “Integrating Identities: An Innovative Bioarchaeological and Biogeochemical Approach to Analyzing the Multiplicity of Identities in the Mortuary Record.” Current Anthropology 58(3): 381–409. Tung, Tiffiny A. 2003. “A Bioarchaeological Perspective on Wari Imperialism in the Andes of Peru: A View from Heartland and Hinterland Skeletal Populations.” Doctoral dissertation, University of North Carolina, Department of Anthropology. Williams, Patrick R. 2001. “Cerro Baúl: A Wari Center on the Tiwanaku Frontier.” Latin American Antiquity 12(1): 67–83. Williams, Patrick R. 2002. “Rethinking Disaster-induced Collapse in the Demise of the Andean Highland States: Wari and Tiwanaku.” World Archaeology 33(3): 361–374. Williams-Blangero, S. 1989. “Clan-structured Migration and Phenotypic Differentiation in the Jirels of Nepal.” Human Biology 61(2): 143–157. Williams-Blangero, S., and J. Blangero. 1989. “Anthropometric Variation and the Genetic Structure of the Jirels of Nepal.” Human Biology 61(1): 1–12. Workman, P.L., J.H. Mielke, and H.R. Nevanlinna. 1976. “The Genetic Structure of Finland.” American Journal of Physical Anthropology 44(2): 341–368. Zvelebil, Marek, and Andrzej W. Weber. 2013. “Human Bioarchaeology: Group Identity and Individual Life Histories—Introduction.” Journal of Anthropological Archaeology 32(3): 275–279.

3 Bioarchaeology and the Narrative Construction of Tewa Identity Scott G. Ortman

In recent years we have seen a decisive move in North American archaeology, stimulated by the Native American Graves Protection and Repatriation Act, to consider relationships between archaeology, bioarchaeology, language, and oral tradition in studies of native North American history, identity, and landscapes (Bernardini 2005; Eiselt 2012; Fowles 2010; Liebmann 2012; Mitchell 2013; Moore 1996; Ortman 2012; Stojanowski 2005b). This renewed commitment to historical anthropology has forced researchers to grapple with a basic issue: that the various lines of evidence available for writing native North American histories reflect different dimensions of social practice, and thus relate to past behavior in different ways. Nowhere is this complexity more apparent than in the relationship between bioarchaeology and oral tradition with respect to migration and ethnogenesis. Individuals can choose how to behave and what to remember, but they cannot alter their genes. The collection and interpretation of bioarchaeological data today does have a political dimension; and past relationships among reproduction, demography, and politics have also left their mark on the bioarchaeological record. Nevertheless, bioarchaeological evidence has the potential to provide a neutral perspective on specific demographic processes. Oral traditions, in contrast, result from peoples’ attempts to make sense of events and deploy the resulting narratives in ways that continue to serve peoples’ interests over time. As a result, oral traditions are

Bioarchaeology and the Narrative Construction of Tewa Identity · 57

political in a more fundamental way in that the raw data, the historical narratives themselves, originate as situated and motivated speech and post-hoc interpretations of events. Oral traditions thus provide a situated perspective on these events, from a vantage that reflects the evolution of the dominant fraction of society as the narrative has been passed down to the present. Given this situation, some have suggested, and continue to suggest, that oral traditions and archaeological (including bioarchaeological) narratives are simply different and cannot be unified (Lekson 2018; Lowie 1915; Mason 2000). In this chapter I argue that, although the relationship between oral tradition and bioarchaeology is complex, and the work of integrating them is hard, the effort is still worthwhile. Bioarchaeology provides reconstructions of past population processes from the etic perspective of analysts today. Oral traditions, on the other hand, derive from the same historical context, but incorporate the emic perspectives and interests of the participants into the narrative. The key point, then, is that bioarchaeology helps to clarify how people of the past made sense of events by providing a basis for comparing the canonical construals of events enshrined in oral traditions with reconstructions of demographic processes reflected in bioarchaeology. When this is done, it is actually the differences between bioarchaeology and oral tradition that are most revealing (see Schmidt 2006). Indeed, the comparison helps one to see how past peoples enshrined and represented the events they experienced through the process Jerome Bruner (1991) has labeled “the narrative construction of reality.” Through this process, we potentially learn things about both the privileged and the submerged discourses of the past, and the goals and interests of past leaders, that we could not learn in any other way. In making this argument I do not mean to suggest that bioarchaeological interpretation is inherently objective, or that oral traditions are necessarily inaccurate. What I do suggest is that the raw data of bioarchaeology result from real population processes, and the raw data of oral traditions enshrine canonical construals of real happenings. The question is not whether either source is more or less accurate with respect to social history. Indeed, historical narratives are always subjective post-hoc interpretations, regardless of who creates them, when, and on what basis. My point instead is that each line of evidence provides a separate vantage

58 · Scott G. Ortman

on the actual happenings that led to the bioarchaeological record and present-day oral traditions. So even if no narrative can be totally objective, I believe it is possible to get a little closer to the actual happenings of the past by triangulating between these radically different points of view. To illustrate this point, I compare the biodistance, genetic, settlement, and narrative evidence related to the origins of the Tewa Pueblo people of northern New Mexico. Due to the long-term persistence of Tewa communities, extensive documentation through ethnography and archaeology, and the confluence of several large-scale research projects, the evidence for reconstructing Tewa history is unusually rich. I have also had the opportunity to partner with contemporary Tewa people from many villages over many years and to incorporate their knowledge and perspectives into my research. Here, I will summarize the population processes involved in Tewa ethnogenesis through a review of current morphological, genetic, and settlement evidence. Then, I will compare the resultant picture with Tewa traditions concerning their origins to make inferences on the discursive practices of Tewa ancestors as they forged a new society and identity in the late thirteenth century. I will show that these traditions were constructed as a means of organizing chaotic experiences of social collapse, mass migration and interethnic interaction into a coherent story that gave reason and purpose to these events, and which supported the integration of numerically dominant immigrants with a smaller group of locals in the emergence of a new Tewa identity. Importantly, it is precisely in the ways that oral tradition and bioarchaeology differ that these discursive strategies become apparent. The Bioarchaeology of Tewa Origins

There are six Tewa-speaking pueblos in northern New Mexico today. In the centuries prior to Spanish contact, Tewa communities were spread across a broader area of the northern Rio Grande Drainage, from the Galisteo Basin in the southeast to the Chama River Valley in the northwest. The correspondence of traditional Tewa names with distinctive varieties of pottery suggests that many of these communities were established in the late thirteenth century (Anschuetz 2005; Harrington 1916; Mera 1934). In addition, it is apparent that the demographic center of gravity of the Pueblo world shifted from the San Juan Drainage to the Rio Grande Drainage during this same period (Cordell 1979; Hill, Clark, and Doelle

Bioarchaeology and the Narrative Construction of Tewa Identity · 59

2010; Kidder 1924; Wendorf and Reed 1955). Finally, statements by Tewa people regarding their origins consistently invoke migration from a distant homeland in the northwest (Harrington 1916; Jeançon 1925; Ortiz 1969; Parsons 1994 [1926]). These various lines of evidence suggest migration was somehow involved in Tewa origins. In the following paragraphs, I summarize biodistance, genetic, and settlement pattern evidence related to this process. The only aspect of migration that is true in all cases is that it involves the movement of people from one place to another. And whereas individuals can change their behavior, they cannot alter their genes. This means genetic data can at least potentially provide a reliable means of inferring past migrations. However, due to repatriation and the wishes of descendant communities, including Tewa people, it is not currently appropriate or feasible to address the sources of the ancestral Tewa population using human DNA. Because of this, my collaborators and I have pursued two alternative data sources as proxies: phenotypic traits of skeletal remains, and ancient DNA from domesticated animals. In the following paragraphs, I discuss these two proxies in turn. Phenotypic Traits A number of recent studies have confirmed that phenotypic traits, and especially metric traits, preserve a signal of regional genetic structure and mating networks (Carson 2006; Cheverud 1988; Konigsberg and Ousley 1995; Relethford 2004; Relethford and Lees 1982; Sparks and Jantz 2002). The advantage of using such traits to investigate migration is that a large amount of such data has been collected from skeletal remains over the past century using standardized skeletal landmarks, including from remains that have since been repatriated to culturally affiliated tribes. Due to sampling concerns, previous studies of regional biological variation in the U.S. Southwest (Akins 1986; Corruccini 1972; El-Najjar 1978, 1981, 1986; Mackey 1977, 1980; Peeples 2014; Schillaci 2003; Schillaci, Ozolins, and Windes 2001; Schillaci and Stojanowski 2005; Turner 1993) have limited applicability to the question of Tewa origins. To better address this question, I compiled a large database of craniometric data from published sources, archival documents, and personal files (Ortman 2009, 2010, 2012). As I compiled this dataset, I reviewed and checked the data from each source, deleted clearly erroneous data, and distinguished

60 · Scott G. Ortman

related measurements taken from different landmarks. I also compared cases where the same remains were measured by two analysts and found that their measurements are highly correlated (r2 >.99), even when collected decades apart. The resulting dataset contains twelve craniofacial measurements that are not influenced by cradle-boarding for approximately twelve hundred adults for which sex was also estimated (raw data available in Ortman [2009]). These remains were recovered from more than 120 archaeological sites across the San Juan and Rio Grande Drainages and date to the centuries immediately before and after the period of Tewa ethnogenesis. I grouped these sites into regional samples according to the archaeological district in which each site occurs, following the definitions of Adler and Johnson (1996) for the pre–AD 1275 period and Adams and Duff (2004) for the post–AD 1275 period. Figure 3.1 presents the centroid of the site coordinates for each regional sample. Note that no post–AD 1275 samples are from the San Juan Drainage. This is because ancestral Pueblo people vacated most of this area, including the Chaco Basin and Mesa Verde region, around AD 1275, coincident with the build-up of population in the Tewa Basin and other areas of the Rio Grande. To analyze these data, I followed procedures developed by John Relethford that estimate the genetic relationship or R matrix directly from metric phenotypic traits. These methods use a model of metric traits as polygenic traits governed by equal and additive effects of multiple genes (Relethford and Lees 1982). This has been the standard approach in biodistance studies of metric traits for some time (Konigsberg and Buikstra 1995; Relethford 2003; Relethford and Blangero 1990; Relethford, Crawford, and Blangero 1997; Scherer 2007; Schillaci 2003; Schillaci and Stojanowski 2005; Steadman 1998; Steadman 2001; Stojanowski 2005a) and has the advantage of allowing estimation of three parameters useful for investigating regional population structure: relative genetic distances between samples; a measure of regional genetic variability known as Wright’s FST; and estimates of relative extra-regional gene flow. To maximize the number of variables and cases in the analysis I included all individuals for which at least four of the twelve measurements are available and estimated missing data using maximum-likelihood methods based on the EM algorithm (Allison 2001). I then controlled for sexual dimorphism by standardizing the raw data by sex before pooling the standardized data for analysis.

Figure 3.1. The study area in the U.S. Southwest and locations of regional samples. Pre–AD 1275 samples are gray, and post–AD 1275 samples are black. The Tewa Basin is the area where contemporary Tewa communities, and ancestral Tewa sites, occur. The VEP Area corresponds to the central Mesa Verde region (after Ortman 2012).

62 · Scott G. Ortman

Figure 3.2 presents the results of principal coordinates analysis of the R matrix derived from the craniometric dataset. This chart provides an overall summary of relationships among the regional samples in the analysis. I have outlined three groups of similar samples that likely comprise distinct biological lineages. There are many interesting details in this plot, but for the purposes of this chapter the most important pattern is that, in most cases, post–AD 1275 samples from ancestral Tewa districts are more similar to pre–AD 1275 samples from the Mesa Verde region than they are to any other pre–AD 1275 samples. The ancestral Tewa districts include the northern Pajarito Plateau, the Galisteo Basin (Tano), the Chama Valley, and the Santa Fe district; and the Mesa Verde region samples encompass southeastern Utah, the McElmo Drainage, and Mesa Verde proper. The only exception to this pattern is the Santa Fe sample, which is more similar to the Totah (Middle San Juan) sample than to Mesa Verde region samples. This analysis suggests that post–A.D. 1275 populations of the Pajarito Plateau, Galisteo Basin, and Chama Valley are primarily descendants of earlier Mesa Verde region populations. However, the ancestral Tewa population of the Santa Fe district appears to have a biological background different from the rest. Figure 3.3 presents an analysis of extra-regional gene flow following the model developed by Relethford and Blangero (1990). This figure summarizes the residual phenotypic variance of each regional sample. When this value is significantly less than zero, it indicates that the population in question experienced less-than-average gene flow from an unspecified external source; and when it is significantly greater than zero it indicates that the population in question experienced greater-than-average gene flow, relative to the overall pattern of genetic variation among the samples in the analysis. The results suggest two points. First, post–AD 1275 samples from the northern Pajarito, Galisteo, and Chama districts did not experience greater-than-average gene flow from other regional populations, including earlier Rio Grande populations. This suggests that the ancestral Tewa populations of these areas did not result from significant admixture of Mesa Verde immigrants with local populations. Second, the post–AD 1275 sample from the Santa Fe district appears to have recently experienced significant gene flow from one or more genetically distinct populations. This is important because the Santa Fe district turns out to be the portion of the Tewa Basin where the local population was largest prior to AD 1275 (Dickson 1979; McNutt 1969; Ortman 2016b; Scheick

Figure 3.2. Principal Coordinates Analysis of the R matrix of the craniometric dataset (after Ortman 2012). The first two eigenvectors have been scaled by the square roots of their eigenvalues. Down-facing, solid triangles indicate post–AD 1275 populations, and up-facing, open triangles indicate pre–AD 1275 populations. The group of samples that includes the ancestral Tewa lineage is circled.

Figure 3.3. Residual Variance from Relethford-Blangero analysis of the craniometric dataset. Regional samples are listed in order of residual variance by time period. Pre–AD 1275 samples have dark gray bars and post–AD 1275 samples have light gray bars. Samples with low residual variance had relatively isolated mating networks; and samples with high residual variance experienced greater than average gene flow.

Bioarchaeology and the Narrative Construction of Tewa Identity · 65

2007; Stubbs and Stallings 1953). Overall, the biodistance results suggest in-migrating Mesa Verde people settled relatively vacant landscapes in the Chama Valley, Pajarito Plateau, and Galisteo Basin. In contrast, the Santa Fe district population appears to have resulted from the blending of at least two previously distinct populations. Genetic Evidence To supplement these results, a group of collaborators and I recently completed a study of mitochondrial DNA extracted from domestic dog (Canis lupus familiaris) and turkey (Meleagris gallopavo) remains in the Mesa Verde region and Northern Rio Grande (Kemp et al. 2017). Given contemporary Pueblo views on destructive analysis of ancestral remains, we reasoned that genetic evidence from commensal species, which could have migrated in tandem with human populations, might provide a suitable proxy for tracing the migrations of both. Several studies have documented this tendency with other species in other parts of the world (e.g., Larson 2012; Ottoni et al. 2013; Thomson, Lebrasseur, Austin et al. 2014; Witt et al. 2015). For this study, we examined mitochondrial DNA haplogroups for turkeys and canids from twenty-seven sites in the Mesa Verde region and eleven sites in the Northern Rio Grande (Kemp et al. 2017: tables 1–2). Our results for dogs were inconclusive (in no small part because all the Northern Rio Grande canid burials for which we obtained mtDNA turned out to be of coyotes), but the results for turkeys replicate the findings from analyses of phenotypic traits discussed above. Our group extracted mitochondrial DNA from 173 turkey samples, obtaining complete or partial mtDNA sequences from 144. From these, we were able to define the haplogroup (aHap1, aHap2, other) for 127 samples. We then combined these results with those from a previous study of ancient turkey domestication in the U.S. Southwest (Speller et al. 2010). Table 3.1 summarizes these combined results, which show that aHap1 dominates over aHap2 in the pre–AD 1275 Mesa Verde region sample and the post–AD 1275 Northern Rio Grande sample, whereas in the pre–AD 1275 Northern Rio Grande sample aHap2 is co-dominant with aHap1. Due to small sample size in this third group, we assessed the statistical significance of these results using Fisher’s exact tests. The results, presented in table 3.2, suggest several points. First, mitochondrial haplogroup counts in the post–AD 1275 turkey populations of the Mesa

66 · Scott G. Ortman

Verde and Northern Rio Grande regions are significantly different and are highly unlikely to represent samples drawn from a single panmictic population. Second, mitochondrial haplogroup counts in the pre–AD 1275 and post–AD 1275 Northern Rio Grande turkey samples are also significantly different, thus reflecting substantial change in haplogroup frequencies within the Northern Rio Grande over time. Third, in contrast to these other comparisons, haplogroup frequencies among turkeys maintained in the pre–AD 1275 Mesa Verde region and the post–AD 1275 Northern Rio Grande are statistically indistinguishable. These results demonstrate that Northern Rio Grande turkey populations were initially distinct from Mesa Verde region turkey populations, but that by the end of the thirteenth century the Northern Rio Grande turkey population had changed to the point that it was indistinguishable from the Mesa Verde region population. Our results are consistent with the hypothesis that the post–AD 1275 turkey population of the Northern Rio Grande derives primarily from the Mesa Verde region. Our results do not rule out regions we did not sample as additional potential sources of the post–AD 1275 turkey population, but the close affinity between the pre–AD 1275 Mesa Verde region sample and the post–AD 1275 Northern Rio Grande sample is consistent with results from studies of human phenotypic traits that suggest the pre–AD 1275 Mesa Verde population and the post–AD 1275 Tewa Basin population comprise a single biological lineage. Patterns of domestic turkey use in the two areas add support to this inference. The relative frequency of turkey remains in Mesa Verde region trash deposits increased dramatically beginning in the AD 1000s as hunting pressure significantly depressed artiodactyl populations (Badenhorst and Driver 2009; Bocinsky et al. 2012; Schwindt et al. 2016). In the Northern Rio Grande, in contrast, domestic turkey remains are relatively rare in pre–AD 1275 contexts, but are much more common in post–AD 1275 contexts. For example, Akins (2013: 218) found that only six of twentythree tabulated pre–AD 1275 assemblages have turkey indices exceeding 0.2, but nine of eighteen tabulated post–AD 1275 assemblages from the same area have indices exceeding 0.5 (see Kemp et al. 2017; Ortman 2012: fig. 13.3). In combination with the genetic data, these findings suggest the turkey population of the post–AD 1275 Northern Rio Grande does not reflect an expansion of the pre–AD 1275 turkey population, but rather an influx of birds that accompanied human immigrants. Our results suggest

Bioarchaeology and the Narrative Construction of Tewa Identity · 67

Table 3.1. Summary of ancient mtDNA haplogroups for turkeys from ancestral Pueblo sites Group

aHap1

aHap2

Other

Total

Pre-migration Tewa Basin

6 (.60)

4 (.40)

0 (.00)

10 (1.0)

Pre-migration Central Mesa Verde Post-migration Tewa Basin

116 (.94)

7 (.05)

1 (.01)

124 (1.0)

66 (.94)

2 (.03)

2 (.03)

70 (1.0)

Total

188 (.92)

13 (.06)

3 (.01)

204 (1.0)

Table 3.2. Results of Fisher’s exact tests (P-values) comparing haplogroup frequencies between groups in Table 3.1 Pre-migration Central Mesa Verde Pre-migration Tewa Basin

0.0194

Post-migration Tewa Basin

1.000

Pre-migration Tewa Basin 0.0212

Note: The null hypothesis in these tests is that turkeys from the compared groups are samples from a single population with the same haplogroup frequencies.

the Mesa Verde region was a plausible source area for both the turkeys and, by proxy, their human keepers. Population History The third line of evidence related to the bioarchaeology of Tewa origins is a reconstruction of the population history of the Tewa Basin. These estimates are based on a database of approximately three thousand habitation sites located within the roughly 5,000 km2 area defined in figure 3.1; tree-ring dates and pottery assemblage data for more than 250 sites; and shapefiles representing survey coverage for the region. The details of this reconstruction are presented in Ortman (2016b). Here, I summarize the methods used and discuss the results. Using the local pottery typology, I developed a chronology of seventeen pottery periods dating between AD 900 and 1760 such that each period is defined by a distinct combination of pottery types. I then estimated the occupation span at each site based on the associated pottery and absolute dates. I also estimated the number of rooms present at each site using

68 · Scott G. Ortman

methods developed by Duwe and others (2016). I assumed a conversion of one person per room (following Brown 1987) and estimated the number of occupied rooms across the occupation spans of sites using simple rules for small sites, and a logistic growth function, calibrated from excavated sites, for sites with fifty or more rooms (Ortman 2012: chapter 4). Then, for those sites associated with a pottery tally, I used uniform probability distributions for pottery types in conjunction with Bayes’ Theorem to refine their population histories (Ortman 2016b). Finally, I assumed all sites with fifty or more rooms are known, but that smaller sites have only been identified systematically within surveyed areas. I used these assumptions in combination with the sample of small sites within surveyed areas to extrapolate total populations through time for each of five distinct sampling strata in the Tewa Basin. Figure 3.4 presents the spatial data in this analysis, and figure 3.5 summarizes the results by sampling stratum. The results illustrate that the population history of the Tewa Basin was quite dynamic. Starting around AD 900, the initial population was concentrated in the Santa Fe stratum, along the Tesuque and Pojoaque rivers east of the Rio Grande. Then, starting around AD 1200, a second population center developed in the Pajarito and Cochiti strata to the west of the river. Population growth on the west side of the river increased dramatically between 1250 and 1280, after which it also increased in the Santa Fe and Chama strata as it leveled off in the Pajarito and Cochiti strata by 1350. In subsequent periods, population declined in three of these areas, but was compensated for by continued growth in the Chama stratum. This suggests many inhabitants of the large fifteenth-century Chama towns derived from earlier settlements of the Pajarito, Chama, and Santa Fe areas. Recent sourcing work on ancestral Tewa pottery of this period supports this interpretation (Duwe 2011, 2019; Duwe and Anschuetz 2013). There are several areas where these results reinforce and extend those from phenotypic traits and ancient DNA. First, the population estimates show that the early Tewa Basin population was concentrated in the Santa Fe district east of the Rio Grande, and this is the area where the biodistance analysis shows the most evidence of admixture. Second, these results show that areas west of the Rio Grande, especially the Pajarito Plateau, was initially settled around AD 1200 and quickly developed into a new population center genetically indistinguishable from that of the Mesa Verde region. Third, these results demonstrate that the Chama stratum

Figure 3.4. Distribution of habitation sites, contemporary pueblo lands, and population strata for the Tewa Basin (after Ortman 2016).

Figure 3.5. Population history of the Tewa Basin (after Ortman 2016). The bold black line represents the total regional population; the other lines represent population histories for each sampling stratum.

Bioarchaeology and the Narrative Construction of Tewa Identity · 71

was largely empty until the late thirteenth century, at which time it was settled by people who were also indistinguishable from Mesa Verde people. Finally, the results suggest the population that resided in the Tewa Basin prior to AD 1200 was swamped by a much larger number of immigrants during the thirteenth century. Overall, the combined results from phenotypic traits, ancient DNA, and settlement demography suggest Tewa origins involved the unification of a relatively small group of locals and a much larger group of immigrants from the Mesa Verde region during the thirteenth century. East of the Rio Grande these immigrants blended with the local population, whereas to the west they settled in previously uninhabited areas. With this scenario in mind, I turn now to a discussion of oral traditions regarding Tewa origins to see how the events in these narratives relate to this picture. Tewa Oral Traditions

Most Tewa people today learn the story of their origin as they grow up. Through my work with Tewa people, I have discussed episodes in this story with many elders from several villages over many years. In general, present-day elders prefer that scholars stick to the published literature in sharing such information with others. I will therefore focus here on a summary of the published sources (Naranjo 2006; Ortiz 1969; Parsons 1929; Parsons 1994 [1926]; Yava 1978). No two versions of the narrative are exactly alike, but all present the same basic outline. The elements they share can be summarized as follows: 1. The people were living in Sip’opʰene beneath ʔOkhа̨ngep’okwinge “Sandy Lake Place” in the distant north. Supernaturals, humans, and animals all lived together in this place, including the first mothers of the Tewa, “Blue Corn Close to Summer” and “White Corn Close to Winter.” 2. “Then they were talking about it, how to go up from the water, how to get ready to go up” (Parsons 1994 [1926]). The corn mothers thus asked one of the men to go out and explore the way by which the people might leave the lake. This man went out to the north, the west, the south, and the east, but each time

72 · Scott G. Ortman

he returned, reporting that the world above was still ochu, or “unripe.” 3. The corn mothers then asked this man to go to the above. When he did so, he was attacked by all of the predatory animals and then magically healed and given a bow and arrow, buckskin clothes, and a headdress of feathers from carrion birds. He returned to the lake as “mountain-lion man,” or the hunt chief. 4. Following this, the hunt chief gave an ear of white corn to one of the men and told him to lead and care for the people during the summer, and gave a second ear of white corn to another man, telling him to lead and care for the people during the winter. Thus the dual chieftainship of the summer and winter moieties came into being, and the moiety chiefs joined the hunt chief as Pa:towa or “Made People.” 5. Next, the Made People told six pairs of brothers, known as Towa’e, to go out to each of the directions and scout the way for the people to leave the lake. Each pair shot arrows to determine the orientations of the directions. The blue brothers went to the north; the yellow brothers, to the west; the red, to the south; and the white, to the east, and all reported seeing a mountain on the horizon, and they slung mud toward each of these, creating the tsin, or cardinal flat-topped hills. The black brothers then went to the above and saw the morning star on the horizon, indicating that the dawn was near, and the all-colored brothers went to the below and saw a rainbow in the distance against the hardening ground. The Towa’e were thus added to the ranks of the Made People. 6. Based on these reports, the people prepared to leave the lake. Summer Chief went out first, but his feet sank into the mud, so Winter Chief led the way, freezing the ground before him so the people could walk. Soon some of the people began to get ill, so the Made People concluded the group was not yet complete; they needed something else before they could leave the lake. Accordingly, they returned to the lake and the Hunt Chief created the first leader of the medicine society. 7. The people then attempted to leave the lake three additional times, but each time the people discovered they were still not complete. Upon their second return to the lake, the K’ósa or “Clown” Society was established; upon the third return, the Scalp Society; and the

Bioarchaeology and the Narrative Construction of Tewa Identity · 73

fourth, the Women’s Society. The leaders of all these societies were added to the ranks of the Made People. 8. Finally, the people were ready to leave, and they proceeded southward in two groups. The first group followed the winter chief down the east side of the Rio Grande and subsisted by means of hunting. The second group followed the summer chief down the west side of the river, subsisting by means of farming and wild-plant collecting. After proceeding southward in twelve steps, they came back together and formed a village containing both groups. In the version recorded by Parsons, this village was TekʰeɁowîngeh, “Cottonwood-bud Pueblo,” located adjacent to present-day Pojoaque Pueblo in the center of the Tewa Basin (Duwe and Cruz 2019); in Ortiz’s version, the village was P’osíɁówîngeh, “Mossgreenness Pueblo,” located adjacent to present-day Ojo Caliente at the northern edge of the Basin (Harrington 1916). This narrative, widely shared among Tewa people today, makes several points of contact with the population processes suggested by the bioarchaeological evidence. First, it refers to the ancestors of Tewa people as having lived in the distant north, in another land, where they gradually developed the core institutions of Tewa society and obtained knowledge of the new land they would eventually occupy. This element of the story is consistent with the bioarchaeological evidence that suggests most Tewa ancestors lived in the Mesa Verde region prior to migrating to the Tewa Basin. Second, the narrative states that Tewa ancestors left the ancestral homeland in two groups, with the Winter people leading the way down along the east side of the Rio Grande, and the Summer people following along the west side. These details are broadly consistent with the demographic evidence that suggests the east side of the Rio Grande was settled first, well-prior to AD 1200, and the west side later, during the thirteenth century (see Duwe and Anschuetz 2013; Duwe and Cruz 2019; Ortman 2018). Third, the narrative states that the summer and winter people rejoined in the Rio Grande to found the first Tewa communities. The archaeology of Tewa origins is also consistent with this view. Tewa identity took shape around AD 1275, and the material culture of sites dating to this period clearly reflect the blending together of the material traditions of the initial settlers on the east side of the Rio Grande (the Winter People) with later settlers on the west side (the Summer People) (Duwe and

74 · Scott G. Ortman

Anschuetz 2013; Ortman 2012). This episode is also consistent with the bioarchaeological evidence, which suggests multiple groups merged in the Santa Fe district east of the Rio Grande in the late thirteenth century. However, these traditions also deviate from the bioarchaeological scenario in several ways. For example, the demographic evidence suggests the in-migrating Mesa Verde population was much larger than the local population; so much so that evidence of genetic admixture is limited to the Santa Fe district and there is little evidence of admixture in other areas. Therefore, the story seems to downplay the demographic differences between the two groups that came together to create ancestral Tewa society. Also, the story claims that Tewa ancestors were a single people who migrated in two groups before coming back together in the Rio Grande. In one sense this is true because Tewa ethnogenesis involved the coming together of Mesa Verde immigrants with a population that had migrated previously to the region. However, the eastern (winter) population entered the region around AD 900, some 250 years prior to the influx of western (summer) people from the Mesa Verde region. It therefore seems unlikely that the existing local population and the immigrant Mesa Verde-affiliated population viewed themselves as a single people when they first encountered each other. Discursive Strategies and Tewa Origins

The bioarchaeological evidence concerning Tewa origins suggests that, in the late thirteenth century, a small local population was swamped by a much larger immigrant population from the Mesa Verde region. Oral traditions surrounding Tewa origins, on the other hand, suggest Tewa society originated from a single group that split into two and rejoined in the Rio Grande following migration. How are we to make sense of these differences? I think the solution builds from two key points. First, the end result of this process was a single Tewa society characterized by a single language, identity, type of community organization, and material culture. Second, although Tewa origin narratives have a teleological character, they must have been constructed retrospectively to explain events that had already happened in ways that served the interests of the most powerful fractions of the emergent society. In other words, the most appropriate way to view Tewa oral traditions is not as eyewitness accounts of past events, but as post-hoc narrative constructions that encapsulate the

Bioarchaeology and the Narrative Construction of Tewa Identity · 75

perspectives of the dominant fraction of the resultant society and represent canonical discourses regarding the way that society came to be. From this perspective, the question is not whether oral traditions are accurate or not. Rather, the question is what the disjunctions between oral tradition and bioarchaeology reveal about the discursive practices of Tewa ancestors as they forged a new society. In this framing, it seems appropriate to view the canonical discourse regarding Tewa origins as one that enshrined the interests and priorities of the immigrants to a greater extent than those of the locals. Although the immigrants outnumbered the locals by about four to one, they were still newcomers in a land that had previously belonged to the locals alone. In situations like this, there is a real risk of the locals becoming hostile to the immigrants, and there is a need to integrate the two groups through a narrative that is accepted by both groups. The narrative constructed by ancestral Tewa leaders, which persists to this day, framed the situation as one where the locals and the immigrants were long-lost relatives who began in the same place but migrated separately to the Tewa Basin. One can imagine why this would serve the interests of both groups. The narrative supports the conclusion that the immigrants belonged despite being newcomers, and it gave them moral legitimacy as heirs to the Tewa Basin landscape. This narrative also supported the legitimacy of the locals by presenting them as equal contributors to the emergent society despite their numerical inferiority. Indeed, the locals may even have welcomed the immigrants and their social institutions, in a manner similar to the acceptance of “stranger kings” in other contexts (Sahlins 2008). Regardless, we know this discursive strategy worked because it became the canonical view among Tewa people. At the same time, the narrative frames Tewa origins as involving a single episode of migration from an ancient homeland in the northwest. Even if the locals and immigrants spoke related languages, and this fact was used to argue that they represented long-lost relatives, the bioarchaeology suggests it is quite a stretch to suggest that these two groups began their migrations to the Rio Grande from a single place and close together in time. Therefore, the canonical story downplays the historical precedence of the east side (winter) people. According to the story the winter people did lead the way down, but the summer people were right behind, and intended to follow all along. The bioarchaeological evidence suggests the prior histories of the two groups were quite different, as the east-side

76 · Scott G. Ortman

(winter) people had lived in the Tewa Basin for centuries before the coming of the west-side (summer) people, and archaeological evidence shows that the Mesa Verde (summer) people participated in the Chacoan regional system before collapsing during a period of drought and internecine conflict, while the local Rio Grande (winter) people avoided Chaco and did not collapse in its aftermath (Glowacki 2015; Lakatos and Post 2012; Lakatos and Wilson 2011). These differences were submerged in the narrative that took shape following the arrival of the summer people. Presumably this is at least in part because emphasizing this history would have de-legitimized the demographic takeover of the Tewa Basin by the immigrants. Indeed, the details of Tewa origin narratives, in light of the bioarchaeological evidence, suggest that an important discursive strategy in their construction involved a willful forgetting of certain details in the service of a larger interest (see Rieff 2016). This comparison of the bioarchaeology and oral traditions surrounding Tewa origins raises many questions that cannot be answered at present. How did the language of the locals relate to that of the immigrants? Did the locals accept the emergent narrative regarding Tewa origins or was it forced on them? Why does the material culture of the resultant Tewa society exhibit closer continuities with the numerically inferior local group than with the numerically dominant immigrant group from the Mesa Verde region? Did the adoption of “local” material culture also serve the interests of the dominant immigrant fraction? If so, in what way? Are there traces in the archaeological record of this coming together of summer and winter people in individual communities? These questions aside, careful attention to the disjunctions between bioarchaeology and oral tradition reveals the discursive strategies pursued by influential people of the day to construct a narrative that accounted for the “real” events surrounding Tewa origins in ways that supported the integration of immigrants and locals in the emerging Tewa society. The resultant narrative framed the process of Tewa ethnogenesis in a way that served the interests of the immigrants while acknowledging and honoring the locals as co-contributors. It is a testament to the political genius of ancestral Tewa leaders that the resultant narrative honors and incorporates the locals as equals in a single society. And perhaps most importantly, one last bit of bioarchaeology, in the form of skeletal trauma from interpersonal violence, shows that the prevalence of such trauma was at times quite high in the Mesa Verde region, but that it

Bioarchaeology and the Narrative Construction of Tewa Identity · 77

declined substantially in the post-migration Northern Rio Grande, despite all the social dynamics I have discussed in this chapter. This suggests that, over the long run, the new Tewa society became more peaceful than Mesa Verde society had ever been (Kohler et al. 2014; Ortman 2016a). So in this case at least, careful attention to the disjunctions between bioarchaeology and oral tradition reveals discursive practices that were highly successful in integrating immigrants and locals into a society that was much more successful at providing for human needs than its precursors had been. These conclusions raise an important point about the relationship between oral tradition and bioarchaeology that historical anthropologists should keep in mind as they construct their own historical narratives. In the early days of American archaeology researchers paid close attention to native oral traditions and interpreted them as relatively straightforward reflections of tribal histories (Cushing 1888; Fewkes 1900; Jeançon 1925; Mooney 1898). Robert Lowie (1915: 598) rightly noted that this approach was a dead end, explaining “we cannot know [oral traditions] to be true except on the basis of extraneous evidence, and in that case they are superfluous since the linguistic, ethnological, or archaeological data suffice to establish the conclusions in question.” In subsequent decades, researchers responded to such critiques by developing methods modeled after journalistic practices to recover more reliable information from oral tradition (Vansina 1961). Such methods do help if one’s purpose is to recover an accurate accounting of the tradition, and it is reasonable to argue that a more accurate accounting should bear a closer relationship to actual past events in a real society. But Lowie’s critique still applies—the only way to know if the result is “correct” is to compare it to other lines of evidence that have a more direct relationship to past behavior. And if this proves to be the case, one could rightly question what oral tradition adds to the discussion. My intervention here is to suggest that, in light of more recent studies of narrative (e.g. Hodges 2011), the right question to ask about oral traditions is not whether they are correct, but what discursive practices they reveal. Oral traditions are narratives constructed with an eye toward sociopolitical legitimacy, and people do care about real events, so one should expect such traditions to bear some relationship to actual past events that can be reconstructed through other means, including bioarchaeology. But at the same time, oral narratives are constructed by people

78 · Scott G. Ortman

living in a real society, where various factions can be expected to have had different interests and perceptions of sociopolitical reality. As a result, analysts should expect oral traditions to emphasize certain aspects of past events while glossing over others, regardless of how accurate our understanding of these traditions becomes. The key point, then, is the same one made by Schmidt (2006) in his study of African oral traditions: it is precisely those areas where oral traditions are “inaccurate” that the greatest interest, and potential for interpretation of past sociopolitical processes, lies. In other words, paying close attention to the disjunctions between oral tradition and other lines of evidence, including bioarchaeology, is the key to constructing narratives that plausibly relate to the histories of real societies, the formation of new social identities, and all that this entails. Works Cited Adams, E. Charles, and Andrew I. Duff, eds. 2004. The Protohistoric Pueblo World, A.D. 1275–1600. Tucson: University of Arizona Press. Adler, Michael, and Amber Johnson. 1996. “Appendix: Mapping the Puebloan Southwest.” In The Prehistoric Pueblo World, A.D. 1150–1350, edited by Michael A. Adler, 255–272. Tucson: University of Arizona Press. Akins, Nancy J. 1986. A Biocultural Approach to Human Burials from Chaco Canyon, New Mexico. Volume 9. Santa Fe, NM: Branch of Cultural Research, Department of the Interior, National Park Service. Akins, Nancy J. 2013. “Northern Rio Grande Faunal Exploitation: A View from the Pajarito Plateau, the Tewa Basin, and Beyond.” In From Mountain Top to Valley Bottom: Understanding Past Land Use in the Northern Rio Grande Valley, New Mexico, edited by Bradley J. Vierra, 215–229. Salt Lake City: University of Utah Press. Allison, Paul D. 2001. Missing Data. Quantitative Applications in the Social Sciences 07–136. Thousand Oaks, CA: Sage Publications. Anschuetz, Kurt F. 2005. “Landscapes as Memory: Archaeological History to Learn From and to Live By.” In Engaged Anthropology: Essays in Honor of Richard I. Ford, edited by Michelle Hegmon and B. Sunday Eiselt, 52–72. Ann Arbor: Museum of Anthropology, University of Michigan. Badenhorst, Shaw, and Jonathan C. Driver. 2009. “Faunal Changes in Farming Communities from Basketmaker II to Pueblo III (A.D. 1–1300) in the San Juan Basin of the American Southwest.” Journal of Archaeological Science 36(9): 1832–1841. Bernardini, Wesley. 2005. Hopi Oral Tradition and the Archaeology of Identity. Tucson: University of Arizona Press. Bocinsky, R. Kyle, Jason A. Cowan, Timothy A. Kohler and C. Davis Johnson. 2012. “How Hunting Changes the VEP World, and How the VEP World Changes Hunting.” In Emergence and Collapse of Early Villages: Models of Central Mesa Verde Archaeol-

Bioarchaeology and the Narrative Construction of Tewa Identity · 79

ogy, edited by Timothy A. Kohler and Mark D. Varien, 145–152. Berkeley: University of California Press. Brown, Barton McCaul. 1987. “Population Estimation from Floor Area: A Restudy of ‘Naroll’s Constant.’” Behavior Science Research 22 (1–4): 1–49. Bruner, Jerome. 1991. “The Narrative Construction of Reality.” Critical Inquiry 18(1): 1–21. Carson, E. Ann. 2006. “Maximum Likelihood Estimation of Human Craniometric Heritabilities.” American Journal of Physical Anthropology 131(2): 169–180. Cheverud, James M. 1988. “A Comparison of Genetic and Phenotypic Correlations.” Evolution 42(5): 958–968. Cordell, Linda S. 1979. “Prehistory: Eastern Anasazi.” In Handbook of North American Indians, vol. 9: Southwest, edited by Alfonso Ortiz, 131–151. Washington, DC: Smithsonian Institution. Corruccini, Robert S. 1972. “The Biological Relationships of Some Prehistoric and Historic Pueblo Populations.” American Journal of Physical Anthropology 37(3): 373–388. Cushing, Frank H. 1888. “Preliminary Notes on the Origin, Working Hypothesis and Preliminary Researches of the Hemenway Southwestern Archaeological Expedition.” Proceedings of the 7th International Congress of Americanists 151–194. Dickson, D. Bruce, Jr. 1979. Prehistoric Pueblo Settlement Patterns: The Arroyo Hondo, New Mexico, Site Survey. Santa Fe, NM: School of American Research. Duwe, Samuel G. 2011. “The Prehispanic Tewa World: Space, Time and Becoming in the Pueblo Southwest.” Doctoral dissertation, University of Arizona, Department of Anthropology. Duwe, Samuel G. 2019. “The Economics of Becoming: Population Coalescence and the Production and Distribution of Ancestral Tewa Pottery.” In Reframing the Northern Rio Grande Pueblo Economy, edited by Scott G. Ortman, 104–118. Anthropological Papers of the University of Arizona. vol. 80. Tucson: University of Arizona Press. Duwe, Samuel G., and Kurt F. Anschuetz. 2013. “Ecological Uncertainty and Organizational Flexibility on the Prehispanic Tewa Landscape: Notes from the Northern Frontier.” In From Mountain Top to Valley Bottom: Understanding Past Land Use in the Northern Rio Grande Valley, New Mexico, edited by Bradley J. Vierra, 95–112. Salt Lake City: University of Utah Press. Duwe, Samuel, and Patrick J. Cruz. 2019. “Tewa Origins and Middle Places.” In The Continuous Path: Pueblo Movement and the Archaeology of Becoming, edited by Samuel Duwe and Robert W. Preucel, 96–123. Tucson: University of Arizona Press. Duwe, Samuel, B. Sunday Eiselt, J. Andrew Darling, Mark D. Willis and Chester Walker. 2016. “The Pueblo Decomposition Model: A Method for Quantifying Architectural Rubble to Estimate Population Size.” Journal of Archaeological Science 65: 20–31. Eiselt, B. Sunday. 2012. Becoming White Clay: A History and Archaeology of Jicarilla Apache Enclavement. Salt Lake City: University of Utah Press. El-Najjar, Mahmoud Y. 1978. “Southwestern Physical Anthropology: Do the Cultural and Biological Parameters Correspond?” American Journal of Physical Anthropology 48(2): 151–157. El-Najjar, Mahmoud Y. 1981. “A Comparative Study of Facial Dimensions at Gran Quivira.” In Contributions to Gran Quivira Archaeology, edited by Alden C. Hayes, 157–

80 · Scott G. Ortman

159. Publications in Archeology, Vol. 17. Washington, DC: Department of the Interior, National Park Service. El-Najjar, Mahmoud Y. 1986. “The Biology and Health of the Prehistoric Inhabitants of Canyon de Chelly.” In Archaeological Investigations at Antelope House, edited by Don P. Morris, 206–220. Washington, DC: Department of the Interior, National Park Service. Fewkes, Jesse W. 1900. “Tusayan Migration Traditions.” In 19th Annual Report of the Bureau of American Ethnology for the Years 1897–1898, Part 2, 573–634. Washington, DC: Government Printing Office. Fowles, Severin M. 2010. “The Southwest School of Landscape Archaeology.” Annual Review of Anthropology 39: 453–468. Glowacki, Donna M. 2015. Living and Leaving: A Social History of Regional Depopulation in Thirteenth-Century Mesa Verde. Tucson: University of Arizona Press. Harrington, John Peabody. 1916. “The Ethnogeography of the Tewa Indians.” In 29th Annual Report of the Bureau of American Ethnology, 29–618. Washington, DC: Government Printing Office. Hill, J. Brett, Jeffery J. Clark, and William H. Doelle. 2010. “Depopulation of the Northern Southwest: A Macro-regional Perspective with Insights from the Hohokam.” In Leaving Mesa Verde: Peril and Change in the Thirteenth-Century Southwest, edited by Timothy A. Kohler, Mark D. Varien, and Aaron W. Wright, 34–52. Tucson: University of Arizona Press. Hodges, Adam. 2011. The “War on Terror” Narrative: Discourse and Intertextuality in the Construction and Contestation of Sociopolitical Reality. Oxford: Oxford University Press. Jeançon, Jean A. 1925. “Primitive Coloradoans.” The Colorado Magazine 2(1): 35–40. Kemp, Brian M., Kathleen Judd, Cara Monroe, Jelmer W. Eerkens, Lindsay Hilldorfer, Connor Cordray, Rebecca Schad, Erin Reams, Scott G. Ortman, and Timothy A. Kohler. 2017. “Prehistoric Mitochondrial DNA of Domesticated Animals Supports a 13th Century Exodus from the Northern US Southwest.” PLoS ONE 12(7): e0178882. Kidder, Alfred V. 1924. An Introduction to the Study of Southwestern Archaeology with a Preliminary Account of the Excavations at Pecos. New Haven, CT: Yale University Press. Kohler, Timothy A., Scott G. Ortman, Katie E. Grundtisch, Carly Fitzpatrick, and Sarah M. Cole. 2014. “The Better Angels of Their Nature: Declining Violence through Time among Prehispanic Farmers of the Pueblo Southwest.” American Antiquity 79(3): 444–464. Konigsberg, Lyle W., and Jane E. Buikstra. 1995. “Regional Approaches to the Investigation of Past Human Biocultural Structure.” In Regional Approaches to Mortuary Analysis, edited by Lane A. Beck, 191–219. New York: Plenum Press. Konigsberg, Lyle W., and Stephen D. Ousley. 1995. “Multivariate Quantitative Genetics of Anthropometric Traits from the Boas Data.” Human Biology 67(3): 481–498. Lakatos, Steven A., and Stephen S. Post. 2012. “Interaction, Accommodation, and Continuity among Early Communities in the Northern Rio Grande Valley, AD 200–900.” In Southwestern Pithouse Communities, AD 200–900, edited by Lisa C. Young and Sarah A. Herr, 123–140. Tucson: University of Arizona Press. Lakatos, Steven A., and C. Dean Wilson. 2011. “The Unexpected Stability of Rio Grande

Bioarchaeology and the Narrative Construction of Tewa Identity · 81

Communities During the Early Developmental Period.” In Crucible of Pueblos: The Early Pueblo Period in the Northern Southwest, edited by Richard H. Wilshusen, Gregson Schachner, and James R. Allison, 127–145. Los Angeles: Cotsen Institute of Archaeology. Larson, Greger. 2012. “Using Pigs as a Proxy to Reconstruct Patterns of Human Migration.” In Population Dynamics in Prehistory and Early History. New Approaches by Using Stable Isotopes and Genetics, edited by Elke Kaiser, Jaochim Burger, and Wolfram Schier, 31–40. Berlin, Boston: De Gruyter. Lekson, Stephen H. 2018. A Study of Southwestern Archaeology. Salt Lake City: University of Utah Press. Liebmann, Matthew J. 2012. Revolt: An Archaeological History of Pueblo Resistance and Revitalization in 17th Century New Mexico. Tucson: University of Arizona Press. Lowie, Robert H. 1915. “Oral Tradition and History.” American Anthropologist 17(3): 597–599. Mackey, James. 1977. “A Multivariate, Osteological Approach to Towa Culture History.” American Journal of Physical Anthropology 46(3): 477–482. Mackey, James. 1980. “Arroyo Hondo Population Affinities.” In Pueblo Population and Society: The Arroyo Hondo Skeletal and Mortuary Remains, edited by Ann M. Palkovich, 171–181. Arroyo Hondo Archaeological Series, Vol. 3. Santa Fe, NM: School of American Research Press. Mason, Ronald J. 2000. “Archaeology and Native North American Oral Traditions.” American Antiquity 65(2): 239–266. McNutt, Charles H. 1969. Early Puebloan Occupations at Tesuque By-Pass and in the Upper Rio Grande Valley. Volume No. 40. Anthropological Papers, Museum of Anthropology, University of Michigan, Ann Arbor. Mera, H.P. 1934. A Survey of the Biscuit Ware Area in Northern New Mexico. Santa Fe, NM: Laboratory of Anthropology. Mitchell, Mark D. 2013. Crafting History on the Northern Plains: A Political Economy of the Heart River Region, 1400–1750. Tucson: University of Arizona Press. Mooney, James. 1898. Calendar History of the Kiowa Indians. Washington, DC: Government Printing Office. Moore, John H. 1996. The Cheyenne. Malden, MA: Blackwell Publishers. Naranjo, Tessie. 2006. “We Came from the South, We Came from the North: Some Tewa Origin Stories.” In The Mesa Verde World, edited by David G. Noble, 49–57. Santa Fe, NM: School of American Research Press. Ortiz, Alfonso. 1969. The Tewa World: Space, Time, Being and Becoming in a Pueblo Society. Chicago: University of Chicago Press. Ortman, Scott G. 2009. “Genes, Language and Culture in Tewa Ethnogenesis, A.D. 1150– 1400.” Doctoral dissertation, Arizona State University, School of Human Evolution and Social Change. Ortman, Scott G. 2010. “Evidence of a Mesa Verde Homeland for the Tewa Pueblos.” In Leaving Mesa Verde: Peril and Change in the Thirteenth Century Southwest, edited by Timothy A. Kohler, Mark D. Varien, and Aaron M. Wright, 222–261. Tucson: University of Arizona Press.

82 · Scott G. Ortman

Ortman, Scott G. 2012. Winds from the North: Tewa Origins and Historical Anthropology. Salt Lake City: University of Utah Press. Ortman, Scott G. 2016a. “Discourse and Human Securities in Tewa Origins.” In The Archaeology of Human Experience, edited by Michelle Hegmon, 74–94. Archeological Papers of the American Anthropological Association, Vol. 27. Washington, DC: American Anthropological Association. Ortman, Scott G. 2016b. “Uniform Probability Density Analysis and Population History in the Northern Rio Grande.” Journal of Archaeological Method and Theory 23(1): 95–126. Ortman, Scott G. 2018. “The Historical Anthropology of Tewa Social Organization.” In Puebloan Societies: Homology and Heterogeneity in Time and Space, edited by Peter M. Whiteley, 51–74. Santa Fe, NM: School of Advanced Research Press. Ottoni, C., L.G. Flink, A. Evin, C. Georg, B. De Cupere, W. Van Neer, L. Bartosiewicz, et al. 2013. “Pig Domestication and Human-Mediated Dispersal in Western Eurasia Revealed through Ancient DNA and Geometric Morphometrics.” Molecular Biology and Evolution 30(4): 824–832. Parsons, Elsie C. 1929. The Social Organization of the Tewa of New Mexico. Memoirs of the American Anthropological Association, no. 36, Washington, DC. Parsons, Elsie C. 1994 [1926]. Tewa Tales. Tucson: University of Arizona Press. Peeples, Matthew. 2014. “Population History of the Zuni Region across the Protohistoric Transition: Migration, Gene Flow, and Social Transformation.” In Building Transnational Archaeologies, edited by Elisa Villapando and Randall H. McGuire, 93–109. Tucson: Arizona State Museum Archaeology Series. Relethford, John H. 2003. “Anthropometric Data and Population History.” In Human Biologists in the Archives: Demography, Health, Nutrition, and Genetics in Historical Populations, edited by D. Ann Herring and Alan C. Swedlund, 32–52. Cambridge: Cambridge University Press. Relethford, John H. 2004. “Boas and Beyond: Migration and Craniometric Variation.” American Journal of Human Biology 16(4): 379–386. Relethford, John H., and John Blangero. 1990. “Detection of Differential Gene Flow from Patterns of Quantitative Variation.” Human Biology 62(1): 5–25. Relethford, John H., Michael H. Crawford, and John Blangero. 1997. “Genetic Drift and Gene flow in Post-Famine Ireland.” Human Biology 69(4): 443–465. Relethford, John H., and Francis C. Lees. 1982. “The Use of Quantitative Traits in the Study of Human Population Structure.” Yearbook of Physical Anthropology 25: 113–132. Rieff, David. 2016. In Praise of Forgetting. New Haven, CT: Yale University Press. Sahlins, Marshall D. 2008. “The Stranger-King or, Elementary Forms of the Politics of Life.” Indonesia and the Malay World 36(105): 177–199. Scheick, Cherie L. 2007. “The Late Developmental and Early Coalition in the Northern Middle Rio Grande: Time or Process?” Kiva 73(2): 131–154. Scherer, Andrew K. 2007. “Population Structure of the Classic Period Maya.” American Journal of Physical Anthropology 132(3): 367–380. Schillaci, Michael A. 2003. “The Development of Population Diversity at Chaco Canyon.” Kiva 68(3): 221–245.

Bioarchaeology and the Narrative Construction of Tewa Identity · 83

Schillaci, Michael A., Erik G. Ozolins, and Thomas C. Windes. 2001. “Multivariate Assessment of Biological Relationships among Prehistoric Southwest Amerindian Populations.” In Following Through: Papers in Honor of Phyllis S. Davis, edited by Regge N. Wiseman, Thomas C. O’Laughlin, and Cordelia T. Snow, 133–149. Albuquerque, NM: Archaeological Society of New Mexico. Schillaci, Michael A., and Christopher M. Stojanowski. 2005. “Craniometric Variation and Population History of the Prehistoric Tewa.” American Journal of Physical Anthropology 126(4): 404–412. Schmidt, Peter R. 2006. Historical Archaeology in Africa: Representation, Social Memory, and Oral Traditions. Lanham, MD: Altamira Press. Schwindt, Dylan M., R. Kyle Bocinsky, Scott G. Ortman, Donna M. Glowacki, Mark D. Varien, and Timothy A. Kohler. 2016. “The Social Consequences of Climate Change in the Central Mesa Verde Region.” American Antiquity 81(1): 74–96. Sparks, Corey S., and Richard L. Jantz. 2002. “A Reassessment of Cranial Plasticity: Boas Revisited.” Proceedings of the National Academy of Sciences USA 99(23): 14636–14639. Speller, Camilla F., Brian M. Kemp, Scott D. Wyatt, Cara Monroe, William D. Lipe, Ursula M. Arndt, and Dongya Y. Yang. 2010. “Ancient Mitochondrial DNA Analysis Reveals Complexity of Indigenous North American Turkey Domestication.” Proceedings of the National Academy of Sciences USA 107(7): 2807–2812. Steadman, Dawnie W. 1998. “The Population Shuffle in the Central Illinois Valley: A Diachronic Model of Mississippian Biocultural Interactions.” World Archaeology 30(2): 306–326. Steadman, Dawnie W. 2001. “Mississippians in Motion? A Population Genetic Analysis of Interregional Gene Flow in West-Central Illinois.” American Journal of Physical Anthropology 114(1): 61–63. Stojanowski, Christopher M. 2005a. “The Bioarchaeology of Identity in Spanish Colonial Florida: Social and Evolutionary Transformation before, during, and after Demographic Collapse.” American Anthropologist 107(3): 417–431. Stojanowski, Christopher M. 2005b. Biocultural Histories in La Florida: A Bioarchaeological Perspective. Tuscaloosa: University of Alabama Press. Stubbs, Stanley A., and W.S. Stallings, Jr. 1953. The Excavation of Pindi Pueblo, New Mexico. Santa Fe, NM: Monographs of the School for American Research and the Laboratory of Anthropology, no. 18. Thomson, V.A., O. Lebrasseur, J.J. Austin, T.L. Hunt, D.A. Burney, T. Denham, N.J. Rawlence, et al. 2014. “Using Ancient DNA to Study the Origins and Dispersal of Ancestral Polynesian Chickens across the Pacific.” Proceedings of the National Academy of Science of the U.S.A. 111(13): 4826–4831. Turner, Christy G., II. 1993. “Southwest Indian Teeth.” National Geographic Research and Exploration 9(1): 32–53. Vansina, Jan. 1961. Oral Tradition: A Study in Historical Methodology. London: Routledge and Kegan Paul. Wendorf, Fred, and Erik K. Reed. 1955. “An Alternative Reconstruction of Northern Rio Grande Prehistory.” El Palacio 62(5–6): 131–173. Witt, Kelsey E., Kathleen Judd, Colin Grier, Timothy A. Kohler, Scott G. Ortman, Brian

84 · Scott G. Ortman

M. Kemp, and Ripan S. Malhi. 2015. “Analysis of Ancient Dogs of the Americas: Determining Possible Founding Haplotypes and Reconstructing Population Histories.” Journal of Human Evolution 79: 105–118. Yava, Albert. 1978. Big Falling Snow. A Tewa-Hopi Indian’s Life and Time, and the History and Traditions of His People. Albuquerque: University of New Mexico Press.

4 Negotiating Contact in the Periphery Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia Lesley A. Gregoricka

The construction and negotiation of identity within human groups represents a complex, dynamic entanglement of individual and social perception within one’s society further complicated by the management of external relations with neighboring societies. In particular, when considerable power differentials are present between groups engaged in trade, hegemonic influence inherent in these interactions may be met with resistance by those communities with less control over these exchanges in an attempt to maintain or mediate identities within broader social structures. Beginning in the Early Bronze Age (ca. 3100–2000 BCE), the emergence of a pan-Gulf interaction sphere linking southeastern Arabia with larger and more powerful neighbors including Mesopotamia and the Indus Valley brought with it the need for structurally diverse cultural systems to interact with one another (Carter 2003; Frank 1993; Potts 2001, 2009; Ratnagar 2001). Such interactions were laden with socioeconomic and ideological pressures from these core areas as demand for resources from the so-called periphery grew. Unlike the urbanized city-states of Mesopotamian and Indus Valley societies, Bronze Age communities in southeastern Arabia did not develop highly stratified social systems, but instead appear to have maintained tribal, kin-based systems of social organization. An analysis of skeletal remains from the region thus presents a unique opportunity for an indepth examination of the circumstances in which these local, “peripheral” populations resisted stratification and actively manipulated their own identities. Unfortunately, due to the commingled and fragmentary nature

86 · Lesley A. Gregoricka

of human remains in third millennium BCE tombs, traditional bioarchaeological methods for analysis are often not possible. Instead, and in conjunction with an assessment of mortuary treatment and social theory, biogeochemical analyses of teeth offer important insight into the nature of interactions between local Arabian communities and their trading partners, as well as what such interactions suggest about identity construction in areas peripheral to major city-states. Seeking Identity in Commingled Remains

Reconstructing identity among past peoples remains an elusive objective because of the complex ways in which individuals perceive themselves in light of participation in innumerable social roles (Knudson and Stojanowski 2008). Here, identity refers to a comprehensive set of personal attributes and self-perception coupled with metaperception (how a person views other people’s views of them), which in turn both shapes and constrains the actions performed by individuals. Perceptions of sex, gender and sexuality, age, ethnicity, and status all serve to holistically inform identity expression, which is itself fluid and may shift not only as one navigates within and between various social arenas but also through time (Díaz-Andreu et al. 2005). Because the human body contains a rich record shaped by these complex biological and social processes, bioarchaeologists use physical remains of the body as a sort of conduit from which to interpret these lived experiences (Agarwal and Glencross 2011). Such analyses require us to take a multifaceted approach informed by interdisciplinary lines of inquiry coupled with rigorous theoretical assessment (Knudson and Stojanowski 2009). Ideally, of course, the examination of intact and complete human skeletal remains (in conjunction with associated mortuary artifacts and detailed archaeological context) provides the most comprehensive foundation from which aspects of identity may be inferred. In reality, however, mortuary archaeology commonly reveals complex, secondary treatment of the dead that may involve the intentional and/or taphonomic commingling and fragmentation of bone (Panakhyo and Jacobi 2016; Tung 2016). This is particularly the case in the Near East, where practices of commingling were pervasive throughout time and affect many skeletal collections studied by bioarchaeologists today (see Sheridan 2017 for a detailed discussion). Sheridan (2017) emphasized the necessity of analyzing

Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia · 87

such assemblages alongside the remains of complete individuals to ensure a more inclusive and representative view of past peoples, especially given the ubiquity of commingling in so many regions. Fortunately, this need has begun to be recognized by bioarchaeologists over the last decade, and both methodological and theoretical work dedicated to explicating these complex assemblages has been undertaken (e.g., Adams and Byrd 2008, 2014; Osterholtz 2016; Osterholtz, Baustian, and Martin 2014). Inarguably, however, commingling and associated fragmentation make bioarchaeological analyses more challenging, limiting the application of standard methodologies (e.g., sex and age estimation) typically undertaken when examining the intact skeleton. Correspondingly, then, extracting evidence for identity from commingled bones requires skill and resolve, but is perhaps most usefully aided by engaging with modern social and mortuary theory. Commingled human remains within the context of mortuary ritual can perhaps best be perceived as both representative of those individuals’ identities and experiences, but also as manipulated objects that continue to interact with the living. Interactions with commingled remains may thus be employed as a means of redefining or enhancing social ties between the dead and the living, facilitating transitions to ancestorhood, defining access rights to territories or resources, and/or reproducing or reinforcing social order (Brück 2006; Chapman 2000; Moutafi and Voutsaki 2016; Saxe 1970; Tung 2014). In this sense, secondary burial treatment resulting in commingling and fragmentation represents the end result of a multi-stage process comprised of actions performed by the living in which bones become imbued with particular meaning that may be disassociated with the sentient individual to whom the remains initially belonged. This kind of object agency may thus transform these remains into cultural tools used by the living to maneuver social, political, economic, or religious spheres of influence (Hodder 2003; Tung 2016). In particular, Tung (2014) compellingly extends Gell’s (1998) argument of the ritual, secondary agency of art objects to human remains—what she refers to as “postmortem agency”—and while rightly asserting that such secondary agency (or effective agency; see Robb 2004) is not equivalent to the primary agency held by living individuals, contends that portions of the body serve to prolong and reassert the agency of once-living people. Moreover, the agential power of these remains to reinforce or redefine social structures relies not on these bones as objects of influence in and of

88 · Lesley A. Gregoricka .

themselves, but on meaning continuously generated as the living interact with these body parts (Tung 2014). Fragmentation similarly offers a useful theoretical lens when applied to commingled assemblages and identity studies. Initially conceived by Chapman (2000; see also Chapman and Gaydarska 2007), fragmentation is used to explicate the purposeful breakage of objects in order to facilitate enchainment, or connecting individuals to one another by means of object fragment exchange within or between places. Concomitantly, the concept of fragmentation has been applied to explain the intentional breaking, manipulating, reassembling, or otherwise reusing of human remains prior to their final deposition in the landscape (Brittain and Harris 2010; Chapman 2000, 2010; Chapman, Wallduck, and Triantaphyllou 2014; Duncan and Schwarz 2014). As a result of the human fragment exchanges embedded in mortuary performances—often part of a long-term process intended to transmute the dead into ancestors, in lieu of a more static, singular event—enchained social relationships are created and solidified (Chapman et al. 2014). While care should be taken to avoid viewing mortuary contexts with a decidedly Westernized, capitalist lens—in other words, as simplistic, nonoverlapping dichotomies contrasting individual versus dividual identities (see Boutin [2016], Fowler [2004], and Moutafi and Voutsaki [2016] for a more in-depth examination of this perspective)—the communal and commingled nature of these interments may generally speak to a deliberate, symbolic removal of individual personhood in favor of the construction of a collective identity (Baustian, Osterholtz, and Cook 2014; Hertz 1906; Shanks and Tilley 1982). As such, studies examining commingled assemblages must necessarily focus on social and not individual identity construction. Such perspectives can offer valuable insight into perceptions of group belonging and, like the maintenance of individual identity, remains a process that must be constructed and continuously reaffirmed by engaging in shared practices at a collective level. Identity at this level can be analyzed by examining not only the communally interred physical bodies themselves, but also the specific mortuary decisions that living communities made. Such decisions have great power to communicate changes in community identity over time, which may be reflective of broader societal transitions in social organization and complexity (e.g., Binford 1971; Peebles 1971).

Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia · 89

Exchange Systems and Mortuary Practices in Bronze Age Arabia

At the end of the fourth millennium BCE, Mesopotamia was in the midst of considerable socioeconomic change as it grappled with a devastating drought and the corresponding collapse of trade routes along the Euphrates (Parker and Goudie 2007; Weiss and Bradley 2001). In particular, the loss of its copper supply from Anatolia eventually led to a revival in interregional trade with the Oman Peninsula—today, the United Arab Emirates and the Sultanate of Oman—and its abundant natural copper reserves after a long hiatus in contact (Carter 2003; Matthews 2002; Potts 1986). The appearance of Mesopotamian ceramics in the Hafit period (ca. 3100–2700 BCE) of southeastern Arabia coincides with archaeological evidence for small-scale mining and ingot production in the mountainous interior of the peninsula, and together, are indicative of intensified trade relations between these two regions (Cleuziou 1996; Potts 2001). By the Umm an-Nar period (ca. 2700–2000 BCE), along with the reemergence and expansion of these exchange networks, local communities throughout the United Arab Emirates became increasingly sedentary with the adoption of oasis agriculture centered around date palm cultivation, while additional cereals and other garden crops were grown in the shade these palms provided (Méry and Tengberg 2009; Potts 1990). Massive fortification towers up to forty meters in diameter constructed around freshwater wells sometimes accompanied these settlements at both inland and coastal sites (Blau 1999; Potts 2001, 2009). Monumental tombs, which in the previous Hafit period were positioned atop high places and in most cases held relatively few individuals, transformed into massive communal mortuary structures 4–14.5 meters in diameter (figure 4.1) often located on low-lying ground in close proximity to Umm an-Nar settlements (Blau 2001; Potts 1990, 1997). Hundreds of individuals of all ages and both sexes were interred together over the course of 200–300 years, with no apparent selective mortuary bias toward any portion of the population (al-Tikriti and Méry 2000; McSweeney, Méry, and Macchiarelli 2008). Primary interments in these circular tombs appear to have been intentionally disturbed over time by the living community as additional corpses were inserted into the tomb’s entrance, which contributed to the commingling observed in the assemblages found within (Blau 2001; Frifelt 1991). Nevertheless, discoveries of partially intact segments

90 · Lesley A. Gregoricka

Figure 4.1. The Umm an-Nar–period tomb of Unar 2 (2300–2100 BCE), located in the Shimal Necropolis in the Emirate of Ras al-Khaimah (photo courtesy of Christian Velde).

of the skeleton within Umm an-Nar tomb contexts suggest that in at least some cases, decomposition was not complete before new bodies introduced into this space pushed aside older interments. Additionally, at sites such as Unar 2 at Shimal, Hili, and Mleiha, it appears that primary interments (perhaps isolated in lower tomb levels) may have been left to decompose and become desiccated as part of excarnation rituals before their removal for additional processing and disarticulation—including cremation—before finally being reinserted into the upper levels of the tomb (alTikriti and Méry 2000; Benton 2006; Blau 2001, 2007; Bondioli et al. 1998; McSweeney et al. 2008). Further evidence of secondary burial practices includes cut-marks identified on human bone recovered from Umm anNar tombs at Tell Abraq and Hili, indicative of postmortem defleshing prior to final interment (Baustian and Martin 2010; Bondioli et al. 1998). A wide variety of both local and nonlocal grave goods also accompanied the commingled dead (Carter 2003; Hellyer 1998; Potts 2009). The presence of nonlocal offerings is perhaps unsurprising given that indigenous communities inhabiting southeastern Arabia during the Early Bronze Age were positioned directly along paths of long-distance trade

Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia · 91

across the Persian Gulf. This strategic geographic location, coupled with high demand for copper, led to the development of the region as a sort of entrepôt engaged in commercial import and export trade activities with regions hundreds or even thousands of kilometers apart—including Mesopotamia, Dilmun, Elam, Central Asia, and the Indus Valley—that thrived until the end of the third millennium BCE (Possehl 1996). Recently, archaeologists have begun linking the appearance of Umm an-Nar culture to a fundamental conflict in social ideology associated with the growing involvement of southeastern Arabia in interregional exchange systems. With the intensification of trade with powerful citystates in Mesopotamia and the Indus Valley, the necessary management of increasingly complex exchange relationships, concomitant with the supervision of labor forces for monument construction and the emergence of craft specialization and oasis agriculture, may have resulted in the emergence of social stratification that posed a direct threat to kinbased forms of social organization and traditional conceptualizations of identity (Cleuziou 2007; Gregoricka 2016; Magee 2014). Interestingly, however, while the archaeological record indicates that Early Bronze Age communities in the United Arab Emirates certainly became more complex throughout the third millennium BCE, such stratification did not burgeon into the development of large, urban centers characterized by rigid social hierarchies (Mouton 2009). Here, tomb membership and the meaning imbued in changing practices of mortuary collectivity were examined together in order to better comprehend identity formation—and potentially, resistance to stratification—among local populations negotiating changing social circumstances amidst external influence. While multiple lines of bioarchaeological evidence would ideally be employed to elucidate such questions, extensive fragmentation and commingling limit traditional methodologies. Nevertheless, utilizing a multi-isotopic approach (as opposed to relying on ratios from a single isotope) enables us to examine in a more nuanced fashion the complexities of human interaction and can offer insight into the forces shaping local identity, particularly when contextualized by data from the mortuary archaeological record and interpreted using social theory.

92 · Lesley A. Gregoricka

Identity and Isotopic Analysis

Analysis of radiogenic strontium (87Sr/86Sr) and stable oxygen (δ18Oc(VPDB)) and carbon (δ13Cap(VPDB)) isotope values from human dental enamel recovered from Umm an-Nar tombs represents an effective strategy to address complex questions about human behavior and adaptation in the past when dealing with commingled assemblages. Strontium isotopes vary in natural abundance based on bedrock geology and exogenous inputs (e.g., rain, sea spray, windblown dust) that generate ratios incorporated into teeth during childhood that are indicative of geographic residence in early life (Bentley 2006; Ericson 1985) (figure 4.2). Conversely, oxygen isotope values integrated into enamel carbonate (CO3) primarily reflect the isotopic signatures of consumed water (Longinelli 1984; Luz and Kolodny 1985; Luz, Kolodny, and Horowitz 1984), the composition of which is determined by a range of geographically specific factors including temperature, altitude, and distance from the coastline (Dansgaard 1964; Gat 1996; White et al. 2000). Finally, while carbon isotope values in human skeletal material are typically utilized to reconstruct paleodietary intake by differentiating between the ingestion of C3 and C4 resources—thus hinting at sex-or status-based access to certain goods—the consumption of isotopically nonlocal foods may also be indicative of culturally disparate dietary practices and the presence of nonlocal individuals (DeNiro 1987; Schoeninger and Moore 1992). Human dental enamel from the Early Bronze Age tombs of Umm anNar Island (Tombs I, II, and V; n = 33), Tell Abraq (n = 29), Mowaihat Tomb B (n = 13), and Unar 1 at Shimal (n = 23) were analyzed. The commingled and fragmentary nature of Umm an-Nar human remains resulted in an inability to match both in situ and isolated teeth with other portions of the skeleton that might otherwise be assessed as part of bioarchaeological investigations into identity formation in the past. Further, poor preservation and diagenetic alteration of bone due in part to arid environmental conditions meant that bone collagen and apatite could not be reliably sampled as part of isotopic analyses. In the absence of individual interments, the tooth type from a single dental quadrant most commonly recovered from tomb contexts (e.g., left mandibular first molars from Tell Abraq) was selected to determine a sampling MNI (minimum number of individuals) and to ensure that the same individual was not re-sampled and thus not represented in the dataset more than once.

Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia · 93

Figure 4.2. Geologic map of southeastern Arabia with Umm an-Nar site/tomb locations discussed in this study (adapted from Gregoricka 2013a: Figure 1).

While previous biogeochemical research in southeastern Arabia has elucidated aspects of residential mobility, diet, nonlocal identity, and fictive kinship (Gregoricka 2013a, 2013b, 2014, 2016, 2019; Schrenk et al. 2016), strontium, oxygen, and carbon isotopic ratios have not been examined in conjunction with one another on a regional scale. Nor have these methods been used to examine social issues and the relationship between social stratification and identity during this time period. Nevertheless, it is important to recognize that identity cannot be solely defined by, nor strictly equated with, geographic origins as determined by stable isotope analysis, but instead is constructed through self-reflection and performance, as well as by the external perceptions of these performances as experienced by others (Knudson and Stojanowski 2009). Correspondingly, then, isotopic information must be incorporated as part of a broader, aggregated dataset to more effectively assess identity as a social construct. In particular, because mortuary practices represent social performances as constructed, dynamic tableaus arranged and edited over time by living

94 · Lesley A. Gregoricka

communities, these funerary presentations—when elucidated using modern social theory—can not only illuminate and contextualize isotopic data, but can elevate it so that interpretations related to identity may be undertaken. Results

Of the individuals sampled from Umm an-Nar mortuary contexts, the vast majority (n = 95) from all six tombs exhibited local strontium isotope ratios (‒x = 0.70887±0.00007, 1σ) ranging from 87Sr/86Sr = 0.70865 to 87Sr/86Sr = 0.70904, a difference of only 0.00039 (figure 4.3). Oxygen isotope values from the same ninety-eight individuals displayed similarly clustered, local values (x‒ = -2.5±0.8‰, 1σ; n = 96) (figure 4.4), although a slightly broader range (-4.7 to -1.1‰) may be attributed to the utilization of different local sources of fresh water depending on site location. While inhabitants of coastal sites like Tell Abraq and Mowaihat obtained water from underground wells (-3.2 to -1.1‰), the people at Shimal (Unar 1) living along the foothills of the Hajjar Mountains likely collected 18O-depleted mountain run-off from wadis (-4.7 to -2.5‰). Carbon isotope values from both local fauna (-9.0 to +5.6‰) and humans (-13.6 to +0.9‰) are consistent with the consumption of a wide variety of C3, C4, and marine resources that varied strongly based on site location (island vs. coastal vs. foothills), with the population on Umm an-Nar Island relying most heavily on marine products as evidenced by the consumption of 13C-enriched foods, the people of Shimal (Unar 1) more dependent on oasis agriculture and 13C-depleted C3 domesticates, and the coastal sites of Tell Abraq and Mowaihat falling between the two. Three outliers—two from Tell Abraq and one from Mowaihat—were identified using strontium isotope analysis and corroborated by oxygen and/or carbon isotope values (table 4.1, figures 4.3 and 4.4). Two individuals (TA 165 and MW 197) exhibited nonlocal values for all three isotopes; a third (TA 161) produced an outlier ratio for strontium, although associated oxygen and carbon isotope values fell with local ranges.

Figure 4.3. Strontium and oxygen isotope ratios from Umm anNar–period human dental enamel across southeastern Arabia. Local 87Sr/86Sr isotope ranges (designated by the area between the dotted lines) were derived from mean archaeological Bronze Age faunal ratios ±2 standard deviations.

Figure 4.4. Oxygen and carbon isotope values from individuals interred in six Umm an-Nar tombs from the United Arab Emirates.

TA 165

MW 197

Tell Abraq

Mowaihat

Local Human Range at Site 0.70882 to 0.70891 0.70882 to 0.70891 0.70884 to 0.70888

Specimen Ratio 0.710661 0.708179 0.708582

87Sr/86Sr

-0.7

-6.0

-1.6

Specimen Value

-3.1 to -1.8

-3.2 to -1.2

-3.2 to -1.2

Local Human Range at Site

δ18Oc(VPDB) (‰)

Note: Italicized values represent those falling outside of ranges defined as local at each site.

TA 161

Outlier Specimen ID

Tell Abraq

Site

-13.0

-13.6

-7.5

Specimen Value

-11.3 to -6.2

-11.6 to -5.4

-11.6 to -5.4

Local Human Range at Site

δ13Cap(VPDB) (‰)

Table 4.1. Strontium, oxygen, and carbon isotope data for three outliers interred in two different Umm an-Nar tombs in the United Arab Emirates.

98 · Lesley A. Gregoricka

Geographic Origins and Mortuary Identity

Nonlocal Interment in Local Mortuary Spaces The presence of at least three nonlocal individuals placed in two Umm an-Nar tombs (Tell Abraq and Mowaihat Tomb B) in part substantiates the Oman Peninsula’s role in Gulf interregional exchange systems as suggested by the large number of foreign grave goods interred alongside the dead. These nonlocals may represent merchants, diplomats, marriage partners as part of exogamous practices tied to maintaining economic relationships, or other immigrants to the region. Regardless of their socioeconomic or political roles, their individual identities have been subsumed within the larger collective through the commingling of their bodies and grave goods with those of locals (Shanks and Tilley 1982). This inclusive fragmentation process may suggest that at least some nonlocals were privy to processes of transmutation into ancestors and played a role in the maintenance and arbitration of enchained social relationships (Chapman 2010; Chapman et al. 2014)—perhaps not only among locals but between local and nonlocal communities. Such inclusivity may also reveal a pointed mortuary strategy in which these nonlocals (who may have originated from core areas where mortuary practices differed substantially from those in southeastern Arabia) were not given special or individualized treatment characteristic of their native cultures. This may reflect the autonomy of this peripheral region relative to core centers in Mesopotamia and the Indus Valley, the active resistance against the adoption of nonlocal mortuary practices by the local inhabitants of these sites, and the inability of core regions to ideologically influence mortuary performances among local Arabian communities (see Gregoricka [2013a] for an expanded discussion), even among the nonlocal dead. To derive a deeper understanding of this relationship, an examination of Stein’s (1999) distance-parity model may be useful here. This economicdriven framework contends that the dominance of the core over the periphery and the resultant inequality and underdevelopment inherent in this relationship will decrease with distance; in other words, the further away the periphery is from the core, the less power the core has over it. Practically, the cost of transportation to and from the core increases as peripheries become increasingly remote, resulting in lessened economic

Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia · 99

pressure and hegemonic control over the periphery as a whole (Stein 1999). While the core might continue to influence peripheral elites to a certain extent, no major changes in local production or social complexity should occur in the periphery. Subsequently, unlike world-systems theory (see Wallerstein [1974, 2000] as well as archaeological applications of world-systems theory by Algaze [1993], Edens and Kohl [1993], and Frank [1993]), power dynamics would remain variable between the core and the periphery, and may even be manipulated to work for the periphery’s advantage (Stein 1999). Correspondingly, freedom of mortuary expression may have in part stemmed from the distance between the Oman Peninsula and core centers elsewhere in the region. For instance, while Mesopotamia sent representatives to establish colonial outposts to places like Anatolia, Syria, and Iran in order to oversee the acquisition of valued natural resources (Oates 1993; Stein 1999, 2002), no evidence of such outposts exist in southeastern Arabia. Reduced economic influence may have more broadly encouraged social and ideological agency among the local inhabitants of the region. Local Identity and Resistance to Stratification Although a handful of nonlocal individuals were identified in Umm anNar tombs, members of the local community clearly dominated these assemblages. As such, Early Bronze Age primary and secondary burial rites involving these remains likely emphasized the reproduction of local forms of social order by means of interacting with, and therefore, imbuing these bones with agential power (Tung 2014). At the same time, such rites may have conferred ancestor status upon deceased members of the local community once commingling facilitated the incorporation of skeletal material into the collective (Shanks and Tilley 1982). However, despite the communal nature of interment during the third millennium BCE, social organization among local Umm an-Nar communities likely underwent considerable change as a result of the emergence and intensification of interregional trade relationships tied to copper production, oasis agriculture, monumental construction projects, and craft specialization. The concomitant development of social complexity and a growing managerial elite to oversee these local and interregional endeavors may have led to the emergence of social hierarchies incongruous with more traditional forms of social organization centered on kin relations.

100 · Lesley A. Gregoricka

As such, increased emphasis on collective funerary practices from the earlier Hafit to the Umm an-Nar period that incorporated all members of the community may have represented a sort of “push back” by the nonelite, exhibiting resistance and agency against growing social hierarchies through the active suppression of stratification in highly visible mortuary spaces. Nonlocal prestige goods also underwent breakage and became commingled alongside the dead, concealing their original placement with particular individuals. This act might similarly reflect resistance against the accumulation of wealth by one or a few elite individuals as mortuary practices caused the movement, breakage, and thus transformation of these items into communal objects disassociated from their original owners (Magee 2014). These funerary performances subsequently served to compensate for changing social dynamics among the living through acceptance of a higher community principle in death, a perspective that perhaps sought to refocus local social systems back to a more egalitarian way of life. Because social order among the living can be continuously reinforced, reaffirmed, but also reinterpreted through treatment of the deceased, the dead exerted considerable influence. In this way, as agential objects (Robb 2004; Tung 2014, 2016) these remains may have become a powerful political tool to suppress stratification, thereby preventing the extreme class differences, urban growth, and overall social complexity seen in Mesopotamia and elsewhere. Conclusions

Examining identity and social change among past populations represented by intentionally commingled and fragmentary remains presents a particular challenge that requires not only careful bioarchaeological analyses but that also takes into account the funerary practices and processes that led to said commingling as informed by mortuary theory. The biogeochemical analyses of radiogenic strontium and stable oxygen and carbon isotopes from the dental enamel of those interred in Umm an-Nar tombs were used to evaluate geographic origins among the tombs’ Bronze Age inhabitants, and from this, to examine social identity and agency in relation to both external interaction with core centers as part of interregional exchange systems, as well as internal social dynamics between the local population and the appearance of local elites. Results indicate that local communities maintained autonomy and resisted outside influence

Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia · 101

as reflected in the performance of local mortuary traditions unaffected by the presence of nonlocals or the region’s involvement in trade with powerful neighbors from core areas like Mesopotamia and the Indus Valley. Moreover, although social change inevitably followed a more sedentary lifestyle in the Umm an-Nar period, including the adoption of oasis agriculture and the appearance of monumental architecture, the actions and influence of a growing faction of local elites may have been actively resisted and even suppressed, facilitated by the political manipulation of the dead and the meaning inherent in interactions with bodies and body parts as the living facilitated a transition to collective ancestorhood. In this way, the value of a more traditional and kin-based social organization, way of life, and social identity may have been preserved. Acknowledgments

This research was funded by the National Science Foundation (BCS0961932), the Philanthropic Education Organization (PEO) Scholar Award, the Ruggles-Gates Fund for Biological Anthropology, and a Sigma Xi Grant-in-Aid of Research. I am indebted to Deb Martin, Christian Velde, Imke Moellering, Dan Potts, Sabah Jasim, Flemming Højlund, Margarethe and Hans-Peter Uerpmann, Richard Meadow, Michele Ziolkowski, Johanna Olafsdotter, and Kim Aaris-Sørensen for their assistance in gaining access to human and faunal skeletal collections as well as logistical support in the United Arab Emirates, Denmark, and the United States. Thanks are also extended to Drew Coleman at the University of North Carolina at Chapel Hill Isotope Geochemistry Laboratory, as well as Andrea Grottoli and Yohei Matsui at The Ohio State University Stable Isotope Biogeochemistry Laboratory. Works Cited Adams, Bradley J., and John E. Byrd, eds. 2008. Recovery, Analysis, and Identification of Commingled Human Remains. Totowa, NJ: Human Press. Adams, Bradley J., and John E. Byrd, eds. 2014. Commingled Human Remains: Methods in Recovery, Analysis, and Identification. Boston: Academic Press. Agarwal, Sabrina C., and Bonnie A. Glencross, eds. 2011. Social Bioarchaeology. Malden, MA: Wiley-Blackwell. Algaze, Guillermo. 1993. The Uruk World System: The Dynamics of Expansion of Early Mesopotamian Civilization. Chicago: University of Chicago Press.

102 · Lesley A. Gregoricka

al-Tikriti, W., and S. Méry. 2000. “Tomb N at Hili and the Question of the Subterranean Graves during the Umm an-Nar Period.” Proceedings of the Seminar for Arabian Studies 30: 205–219. Baustian, Kathryn M., Anna J. Osterholtz, and Della C. Cook. 2014. “Taking Analyses of Commingled Remains in to the Future: Challenges and Prospects.” In Commingled and Disarticulated Human Remains: Working Toward Improved Theory, Method, and Data, edited by Anna J. Osterholtz, Kathryn M. Baustian, and Debra L. Martin, 265– 274. New York: Springer. Baustian, Kathryn M., and Debra L. Martin. 2010. “Patterns of Mortality in a Bronze Age Tomb from Tell Abraq.” In Death and Burial in Arabia and Beyond: Multidisciplinary Perspectives, edited by Lloyd Weeks, 55–59. Oxford: Archaeopress. Bentley, R. Alexander. 2006. “Strontium Isotopes from the Earth to the Archaeological Skeleton: A Review.” Journal of Archaeological Method and Theory 13(3): 135–187. Benton, Jodie. 2006. Burial Practices of the Third Millennium BC in the Oman Peninsula: A Reconsideration. Sydney: University of Sydney. Binford, Lewis. 1971. “Mortuary Practices: Their Study and Their Potential.” In Approaches to the Social Dimensions of Mortuary Practices, edited by James A. Brown, 6–29. Washington, DC: Society for American Archaeology. Blau, Soren. 2001. “Fragmentary Endings: A Discussion of 3rd-Millennium BC Burial Practices in the Oman Peninsula.” Antiquity 75(289): 557–570. Blau, Soren. 2007. “Skeletal and Dental Health and Subsistence Change in the United Arab Emirates.” In Ancient Health: Skeletal Indicators of Agricultural and Economic Intensification, edited by Mark N. Cohen and Gillian M.M. Crane-Kramer, 190–206. Gainesville: University Press of Florida. Bondioli, L., A. Coppa, and R. Macchiarelli. 1998. “From the Coast to the Oasis in Prehistoric Arabia: What the Human Osteodental Remains Tell Us about the Transition from a Foraging to the Exchange Economy? Evidence from Ra’s al-Hamra (Oman) and Hili North (U.A.E.).” In Proceedings of the XIII Congress, Forli, Italy, 229–234. Boutin, Alexis. 2016. “Exploring the Social Construction of Disability: An Application of the Bioarchaeology of Personhood Model to a Pathological Skeleton from Ancient Bahrain.” International Journal of Paleopathology 12: 17–28. Brittain, Marcus, and Oliver Harris. 2010. “Enchaining Arguments and Fragmenting Assumptions: Reconsidering the Fragmentation Debate in Archaeology.” World Archaeology 42(4): 581–594. Brück, Joanna. 2006. “Fragmentation, Personhood and the Social Construction of Technology in Middle and Late Bronze Age Britain.” Cambridge Archaeological Journal 16(3): 297–315. Carter, Robert. 2003. “Restructuring Bronze Age Trade: Bahrain, Southeast Arabia and the Copper Question.” In The Archaeology of Bahrain: the British Contribution, edited by Harriet Crawford, 31–42. Oxford: BAR International Series 1189. Chapman, John. 2000. Fragmentation in Archaeology: People, Places and Broken Objects in the Prehistory of Southeastern Europe. London: Routledge. Chapman, John. 2010. “‘Deviant’ Burials in the Neolithic and Chalcolithic of Central and South Eastern Europe.” In Body Parts and Bodies Whole, edited by Katharina

Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia · 103

Rebay-Salisbury, Marie Louise Stig Sørensen, and Jessica Hughes, 30–45. Oxford: Oxbow Books. Chapman, John, and Bisserka Gaydarska. 2007. Parts and Wholes: Fragmentation in Prehistoric Context. Oxford: Oxbow Books. Chapman, John, Rosalind Wallduck, and Sevi Triantaphyllou. 2014. “Disarticulated Human Bone Dispersal during the Mesolithic, Neolithic and Chalcolithic in the Balkans and Greece.” In Archaeothanatology: An Interdisciplinary Approach on Death from Prehistory to the Middle Ages, edited by Mihai Gligor, 11–45. Annales Universitatis Apulensis, Series Historica 18/II. Alba Iulia, Romania: Editura Mega. Cleuziou, Serge. 1996. “The Emergence of Oases and Towns in Eastern and Southern Arabia.” In The Prehistory of Asia and Oceania, edited by G. Afanas’ev, S. Cleuziou, J. Lukacs, and M. Tosi, 159–165. Forli: ABACO Edizioni. Cleuziou, Serge. 2007. “Evolution toward Complexity in a Coastal Desert Environment: The Early Bronze Age in the Ja’alan, Sultanate of Oman.” In The Model-based Archaeology of Socionatural Systems, edited by Timothy Kohler and Sander van der Leeuw, 209–227. Santa Fe, NM: School for Advanced Research. Dansgaard, W. 1964. “Stable Isotopes in Precipitation.” Tellus 16(4): 436–468. DeNiro, Michael J. 1987. “Stable Isotopy and Archaeology.” American Scientist 75(2): 182–191. Díaz-Andreu, Margarita, Sam Lucy, Staša Babić, and David N. Edwards. 2005. The Archaeology of Identity: Approaches to Gender, Age, Status, Ethnicity, and Religion. London: Routledge. Duncan, William N., and Kevin R. Schwarz. 2014. “Partible, Permeable, and Relational Bodies in a Maya Mass Grave.” In Commingled and Disarticulated Human Remains: Working Toward Improved Theory, Method, and Data, edited by Anna J. Osterholtz, Kathryn M. Baustian, and Debra L. Martin, 149–170. New York: Springer. Edens, C., and P. Kohl. 1993. “Trade and World Systems in Early Bronze Age Western Asia.” In Trade and Exchange in Prehistoric Europe, edited by Christopher Scarre and Frances Healy, 17–34. Oxford: Oxbow Monographs No. 33. Ericson, Jonathon E. 1985. “Strontium Isotope Characterization in the Study of Prehistoric Human Ecology.” Journal of Human Evolution 14(5): 503–514. Fowler, Chris. 2004. The Archaeology of Personhood: An Anthropological Approach. New York: Routledge. Frank, Andre Gunder. 1993. “Bronze Age World System Cycles.” Current Anthropology 34(4): 383–405. Gat, J. 1996. “Oxygen and Hydrogen Isotopes in the Hydrologic Cycle.” Annual Review of Earth and Planetary Sciences 24: 225–262. Gell, Alfred. 1998. Art and Agency: An Anthropological Theory. New York: Clarendon Press. Gregoricka, Lesley A. 2013a. “Residential Mobility and Social Identity in the Periphery: Strontium Isotope Analysis of Archaeological Tooth Enamel from Southeastern Arabia.” Journal of Archaeological Science 40(1): 452–464. Gregoricka, Lesley A. 2013b. “Geographic Origins and Dietary Transitions during the Bronze Age in the Oman Peninsula.” American Journal of Physical Anthropology 152(3): 353–369. Gregoricka, Lesley A. 2014. “Assessing Life History from Commingled Assemblages: The

104 · Lesley A. Gregoricka

Biogeochemistry of Inter-tooth Variability in Bronze Age Arabia.” Journal of Archaeological Science 47(1): 10–21. Gregoricka, Lesley A. 2016. “Human Response to Climate Change during the Umm anNar/Wadi Suq Transition in the United Arab Emirates.” International Journal of Osteoarchaeology 26(2): 211–220. Gregoricka, Lesley A. 2019. “Temporal Trends in Mobility and Subsistence Economy among the Tomb Builders of Umm an-Nar Island.” In Life and Death in Ancient Arabia: Mortuary and Bioarchaeological Perspectives, edited by Kimberly D. Williams and Lesley A. Gregoricka. Gainesville: University Press of Florida. Hellyer, Peter. 1998. Hidden Riches: An Archaeological Introduction to the United Arab Emirates. Abu Dhabi: Union National Bank. Hertz, Robert. 1960 [1906]. Death and the Right Hand. Translated by R. Needham and C. Needham. Glencoe, IL: The Free Press. Hodder, Ian. 2003. “The ‘Social’ in Archaeological Theory: An Historical and Contemporary Perspective.” In A Companion to Social Archaeology, edited by Lynn Meskell and Robert Pruecel, 23–42. Malden, MA: Blackwell. Knudson, Kelly J., and Deborah E. Blom. 2009. “The Complex Relationship between Tiwanaku Mortuary Identity and Geographic Origin in the South Central Andes.” In Bioarchaeology and Identity in the Americas, edited by Kelly J. Knudson and Christopher M. Stojanowski, 194–211. Gainesville: University Press of Florida. Knudson, Kelly J., and Christopher M. Stojanowski. 2009. “The Bioarchaeology of Identity.” In Bioarchaeology and Identity in the Americas, edited by Kelly J. Knudson and Christopher M. Stojanowski, 1–23. Gainesville: University Press of Florida. Longinelli, Antonio. 1984. “Oxygen Isotopes in Mammal Bone Phosphate: A New Tool for Paleohydrological and Paleoclimatogical Research?” Geochimica et Cosmochimica Acta 48(2): 385–390. Luz, Boaz, and Yehoshua Kolodny. 1985. “Oxygen Isotope Variations in Bone Phosphate of Biogenic Apatites. IV. Mammal Teeth and Bones.” Earth and Planetary Science Letters 75(1): 29–36. Luz, Boaz, Yehoshua Kolodny, and Michael Horowitz. 1984. “Fractionation of Oxygen Isotopes between Mammalian Bone-Phosphate and Environmental Drinking Water.” Geochimica et Cosmochimica Acta 48(8): 1689–1693. Magee, Peter. 2014. The Archaeology of Prehistoric Arabia. Cambridge: Cambridge University Press. Matthews, Roger. 2002. Secrets of the Dark Mound: Jemdet Nasr 19261928. London: British School of Archaeology in Iraq. McSweeney, Kathleen, Sophie Méry, and Roberto Macchiarelli. 2008. “Rewriting the End of the Early Bronze Age in the United Arab Emirates through the Anthropological and Artefactual Evaluation of Two Collective Umm anNar graves at Hili (Eastern Region of Abu Dhabi).” Arabian Archaeology and Epigraphy 19(1): 114. Méry, Sophie, and Margareta Tengberg. 2009. “Food for Eternity? The Analysis of a Date Offering from a 3rd Millennium BC Grave at Hili N, Abu Dhabi (United Arab Emirates).” Journal of Archaeological Science 36(9): 2012–2017. Moutafi, Ioanna, and Sofia Voutsaki. 2016. “Commingled Burials and Shifting Notions

Commingled Mortuary Practices and Identity Construction in Bronze Age Arabia · 105

of the Self at the Onset of the Mycenaean Era (1700–1500 BCE): The Case of the Ayios Vasilios North Cemetery, Laconia.” Journal of Archaeological Science: Reports 10: 780–790. Oates, Joan. 1993. “Trade and Power in the Fifth and Fourth Millennia BC: New Evidence from Northern Mesopotamia.” World Archaeology 24(3): 403–422. Osterholtz, Anna J., ed., 2016. Theoretical Approaches to Analysis and Interpretation of Commingled Human Remains. New York: Springer. Osterholtz, Anna J., Kathryn M. Baustian, and Debra L. Martin, eds. 2014. Commingled and Disarticulated Human Remains: Working toward Improved Theory, Method, and Data. New York: Springer. Panakhyo, Maria, and Keith Jacobi. 2016. “Limited Circumstances: Creating a Better Understanding of Prehistoric Peoples through the Reanalysis of Collections of Commingled Human Remains.” In Theoretical Approaches to Analysis and Interpretation of Commingled Human Remains, edited by Anna J. Osterholtz, 75–96. New York: Springer. Parker, A., and A. Goudie. 2007. “Development of the Bronze Age Landscape in the Southeastern Arabian Gulf.” Arabian Archaeology and Epigraphy 18(2): 132–138. Peebles, Christopher S. 1971. “Moundville and Surrounding Sites: Some Structural Considerations of Mortuary Practices II.” In Approaches to the Social Dimensions of Mortuary Practices, edited by James Brown, 68–91. Washington, DC: Society for American Archaeology. Possehl, Gregory. 1996. “Meluhha.” In The Indian Ocean in Antiquity, edited by Julian Reade, 133–208. London: Keagan Paul. Potts, Daniel. 1986. “Eastern Arabia and the Oman Peninsula during the Late Fourth and Early Third Millennium B.C.” In Gamdat Nasr: Period or Regional Style?, edited by Uwe Finkbeiner and Wolfgang Rollig, 121–170. Wiesbaden: Tubinger Atlas des Vorderen Orients Beiheft B 62. Potts, Daniel. 1990. The Arabian Gulf in Antiquity, Vol. I: From Prehistory to the Fall of the Achaemenid Empire. Oxford: Clarendon Press. Potts, Daniel. 2001. “Before the Emirates: An Archaeological and Historical Account of Developments in the Region c. 5000 BC to 676 AD.” In United Arab Emirates: A New Perspective, edited by Ibrahim Al Abed and Peter Hellyer, 28–69. London: Trident Press Ltd. Potts, Daniel. 2009. “The Archaeology and Early History of the Persian Gulf.” In The Persian Gulf in History, edited by Lawrence G. Potter, 27–53. New York: Palgrave Macmillan. Ratnagar, Shereen. 2001. “The Bronze Age: Unique Instance of a Pre-Industrial World System?” Current Anthropology 42(3): 351–365. Robb, John. 2004. “The Extended Artefact and the Monumental Economy: A Methodology for Material Agency.” In Re-thinking Materiality: The Engagement of Mind with the Material World, edited by Elizabeth DeMarrais, Chris Gosden, and Colin Renfrew, 131–139. Cambridge: McDonald Institute for Archaeological Research. Saxe, Arthur A. 1970. “Social Dimensions of Mortuary Practices.” Doctoral dissertation, University of Michigan, Department of Anthropology. Schoeninger, Margaret, and K. Moore. 1992. “Bone Stable Isotope Studies in Archaeology.” Journal of World Prehistory 6(2): 247–296.

106 · Lesley A. Gregoricka

Schrenk, Alecia, Lesley A. Gregoricka, Debra L. Martin, and Daniel T. Potts. 2016. “Differential Diagnosis of a Progressive Neuromuscular Disorder using Bioarchaeological and Biogeochemical Evidence from a Bronze Age Skeleton in the UAE.” International Journal of Paleopathology 13: 1–10. Shanks, Michael, and Christopher Tilley. 1982. “Ideology, Symbolic Power, and Ritual Communication: A Reinterpretation of Neolithic Mortuary Practices.” In Symbolic and Structural Archaeology, edited by Ian Hodder, 129–154. Cambridge: Cambridge University Press. Sheridan, Susan G. 2017. “Bioarchaeology in the Ancient Near East: Challenges and Future Directions for the Southern Levant.” Yearbook of Physical Anthropology 162(S63): 110–152. Stein, Gil. 1999. Rethinking World-Systems: Diasporas, Colonies, and Interaction in Uruk Mesopotamia. Tucson: University of Arizona Press. Stein, Gil. 2002. “Colonies without Colonialism: A Trade Diaspora Model of Fourth Millennium B.C. Mesopotamian Enclaves in Anatolia.” In The Archaeology of Colonialism, edited by Claire L. Lyons and John K. Papadopoulos, 27–64. Los Angeles: Getty Research Institute. Tung, Tiffiny A. 2014. “Agency: Til Death Do Us Part? Inquiring about the Agency of Dead Bodies from the Ancient Andes.” Cambridge Archaeological Journal 24(3): 437–452. Tung, Tiffiny A. 2016. “Commingled Bodies and Mixed and Communal Identities.” In Theoretical Approaches to Analysis and Interpretation of Commingled Human Remains, edited by Anna J. Osterholtz, 243–251. New York: Springer. Wallerstein, Immanuel. 1974. The Modern World-System: Capitalist Agriculture and the Origins of the European World Economy in the Sixteenth Century. New York: Academic Press. Wallerstein, Immanuel. 2000. The Essential Wallerstein. New York: The New Press. Weiss, Harvey, and Raymond S. Bradley. 2001. “What Drives Societal Collapse?” Science 291(5504), 609–610. White, Christine, Michael W. Spence, Fred J. Longstaffe, and Kimberley R. Law. 2000. “Testing the Nature of Teotihuacan Imperialism at Kaminaljuyu using Phosphate Oxygen-Isotope Ratios.” Journal of Anthropological Research 56(4): 535–558.

5 Collective Bodies, Collective Identities The Development of Identity in Bronze Age Cyprus Anna Osterholtz

The development of a unique Cypriot identity developed over time and in relation to the complex interactions occurring on Cyprus beginning in the Neolithic. The small island has long held a strategic place in the eastern Mediterranean. It has been a stopping place in trading missions between the western and eastern Mediterranean (figure 5.1), partly due to the nature of the currents within the Mediterranean basin, which allow easy access to natural ports and, later, to constructed harbors (Simmons 2014). Coastal trading centers enabled the exchange of goods including copper, olive oil, and wine throughout the region. In addition, dye works have been excavated (Karageorghis 1976), suggesting that textile production and trade may have been important. This process began in the Neolithic, as evidenced by sites like Ais Giorkis, an early Neolithic period site where nonlocal goods have been documented. The expression of social identity on Cyprus has always reflected the specifics of its geographic location, which combines elements of both insularity and connectivity (e.g., Knapp 1990a; Muhley, Madden, and Stech 1988; Sherratt and Sherratt 1993). In the Bronze Age, the cultural contacts with regional trading partners within the eastern and western Mediterranean shaped the development of what it meant to be Cypriot. This chapter examines the roles of tomb use, location, and reuse in the development and renegotiation of Cypriot identity from the Early to the Late Bronze Ages through major population growth and increasing social complexity, which may have been partially the result of external trade relationships (e.g., Kelder 2009).

Figure 5.1. Map of Cyprus with sites discussed in the text. Inset shows directions and paths of currents within the eastern Mediterranean.

Collective Bodies, Collective Identities: Development in Bronze Age Cyprus · 109

Previous research on Cypriot identity used archaeological datasets that can show an individual’s social identity (e.g., Frankel 2000). Models developed using archaeological data are presented here, followed by assessment of the models based on analysis of several skeletal assemblages from throughout the Bronze Age. While previous large-scale analyses of Cypriot identity did not include skeletal material as an integrated dataset, mortuary ritual and the human skeletal remains that result from this ritual are an invaluable resource. Manipulation of the body and the choices made by the living with respect to treatment and placement of the dead establish the relationship between the living and the dead. The treatment of the dead, therefore, serves as a mechanism for the formation and renegotiation of social identities both for the individual and society at large. Colonization versus Migration as Identity Formation Processes

This chapter examines two distinct periods of increasing population size on the island of Cyprus—one at the beginning of the PreBA (prehistoric Bronze Age, 2400–1700 BC) and another at the beginning of the ProBA (protohistoric Bronze Age, 1700–1100 BC), following Knapp’s (2013) chronology. Both of these population increases occurred at times of technological innovation and/or social change. Technological innovation and its identification through material culture has long been associated with social identity. Frankel refers to this as “habitus” (2000). Following Bourdieu (1977), habitus is made up of society as lived experience. This must include elements of material culture as well as social interaction and structure. Using this understanding, mortuary ritual (including the treatment of the dead) is an element of habitus and should be examined as an important aspect of understanding lived experience in the past. While population increases can occur through a variety of methods, two are the most germane in the formation of Cypriot identity. The first is colonization and the second is the postcolonial concept of integration and creation of a third-space phenomenon (see figure 5.2 for visual representations of these models). These models will be discussed individually below, with reference to the expected skeletal changes for each model.

110 · Anna Osterholtz

Figure 5.2. Migration/integration and colonization models of population movement discussed in the text.

Colonization Colonization can be defined as the social process through which newcomers mandate social change upon the indigenous inhabitants of the land where they settle (figure 5.2). Primary examples of colonization can be found in the New World, and can lead to forced displacement of indigenous populations, significant health impacts, and changes to the social use of violence. Primarily this seems to be due to the disruption of social ties among forcibly displaced groups; social ties are an important buffer against nutritional deficiency, used to protect vulnerable groups and to provide medical care (Osterholtz 2017). Colonization is also associated with the concept of virgin soil epidemics (Crosby 1976; Jones 2003), a condition where indigenous populations have no prior exposure or immunity to the diseases of the colonizers. Colonialism is inferred archaeologically through multiple methods, including changes in architecture and material culture, as well as the introduction of new foods, new agricultural methods, and/ or new manufacturing techniques. It may also be visible in the introduction of new

Collective Bodies, Collective Identities: Development in Bronze Age Cyprus · 111

mortuary monumentation or practices associated with the colonizer’s homeland. Colonization is expected to present osteologically with decreased health scores after the introduction of a colonizing population (and the associated forced social changes), as well as changing patterns of violence (e.g., changes in the location and/or frequency of cranial depression fractures in different sexes or age cohorts). When the lived experience of indigenous groups suffers under colonialism, it is often inscribed on the bones in the form of poorer health scores and different patterns of trauma. Migration and Integration, or Third-Space Phenomena The concept of the creation of a third-space phenomenon grows out of postcolonial theory, and is predicated on the assumption that change cannot extend only one way when two groups interact with each other (figure 5.2). This leads to a hybridization of the migrant group and the indigenous group. The resulting third-space will contain elements of both previous cultures and social identities, but will also contain new elements as the combination of old and new leads to the identification of a new social identity with close relationships to both past groups. Hall (1997: 30) defines social identity as “the knowledge, value and significance attached to membership in a social group.” Practice theory teaches us that activity and behavior are fundamental to cultural affiliation; in effect culture does, not is (Nielsen and Walker 2009). If trade and exposure to other groups is a habitual part of life, then identity must be (at least in part) constructed around interaction. Boardman (2001) and Sommer (2007) both note that familiarity with outside groups increases the susceptibility of one group to be open to new ideas or immigrants from external sources. Increased exposure to other groups within a Mediterranean regional trading sphere would also have exposed populations to each other’s microbiomes and diseases, creating a common immunological background between the groups and lessening the likelihood of virgin soil disease transfer. This shared disease exposure is due to significant contact, implying extensive cultural interaction. In trading groups, the exchange of goods often goes hand in hand with the exchange of ideas and mating partners. In this model, migrants are incorporated into existing social structure and elements of their home cultures are added to the indigenous

112 · Anna Osterholtz

group’s culture, creating a new identity that hybridizes both old and new traditions. The hybridization of populations containing both newcomers and existing groups requires there to be significant previous contact for a smooth transition to a new third-space. While this chapter focuses primarily on prehistoric periods on Cyprus, examining the modern experience of Cypriots living within Cyprus and those living within the Turkish Republic of Northern Cyprus may be informative as well. In discussing the interaction of modern “Greek” (those living on Cyprus) and Turkish Cypriot (those living in the Turkish Republic of Northern Cyprus) interactions, Stevens and colleagues (2014: 1738) note that “national identifications and the constructed physical and cultural boundaries change according to social context.” Social remittances are at play with integration and hybridization as well, and Levitt and Lamba-Nieves (2010: 2) write that, “migrants carry ideas, practices and narratives which enable mobility and different forms of membership and belonging.” Migrants may also stay in contact with their previous places of origin and transmit back cultural information from this new third-space to their place of origin. In this way, cultural exchange can be viewed as a constantly changing, continual feedback loop. This loop helps to create continuity with the parent populations while creating a new identity through interactions (thus creating a third-space). Migration theory examines movement in terms of circulation and diffusion (Mains et al. 2013; Vertovec 2001). In essence, social bonds are formed in their new homelands, but the social bonds that existed in their places of origin may not have been severed. Migrants may “link their cross-cutting belongingness with complex attachments and multiple allegiances to issues, peoples, places and traditions beyond the boundaries of their resident nation-states” (Çaglar 2001: 610). This interconnectedness likely produced cultural and biological buffering. While different cultural traditions of violence (both physical and structural, explored in detail below) may evolve within the new third-space, this is not necessarily the case. In effect, where colonization breaks social ties, migration and integration promotes social ties between the migrant and the indigenous populations. Therefore, health scores are not expected to differ after a migration incident, assuming that other social and environmental factors remain constant. In short, migration and integration present a different

Collective Bodies, Collective Identities: Development in Bronze Age Cyprus · 113

osteological signature. We should see secular changes (the normal shifts that are expected with cultural changes in grain processing, aggregation and increased population sizes), but overall patterns of health should be more consistent over time. This consistency is indicative of cultural buffering mechanisms that incorporate groups through mutual social changes as opposed to enforced social changes that break cultural buffering mechanisms (as would be seen in colonization). Movements and Migrations on Cyprus

From the Neolithic period onward, visitors came to Cyprus bringing foreign goods and interacting with local populations (Simmons 2014). Janes (2010: 144) describes the island as a “stepping stone,” forming a focal point for interaction and trade with southwest Asia, Egypt, and the western Mediterranean. Broodbank (2013: 373) describes the eastern Mediterranean as a “theatre of interaction” with goods and skilled craftspeople moving between islands. The sections below explore population structure and mortuary practices throughout the PreBA and ProBA periods, with special emphasis on population size increases and major cultural changes for each time period. PreBA Population Structure and Mortuary Practices This chapter focuses on the population size increases that occurred during the PreBA (prehistoric Bronze Age, 2400–1700 BC) and ProBA (protohistoric Bronze Age, 1700–1100 BC). The first population size increase at the beginning of the PreBA is associated with the reintroduction of cattle onto Cyprus, the expansion of agricultural lands, and the development of agricultural surplus with redistribution centers. The early PreBA marks the beginning of true social stratification on the island. Material culture changes associated with this time period include the introduction of Anatolian design elements on locally produced ceramic spindle whorls, which Frankel (2002) interprets as the imposition of Anatolian habitus onto a local population. Knapp (1990b) describes a positive feedback loop regarding cattle and copper, noting that copper tools were likely used to clear forested land for agriculture, which allowed for the use of cattle in increased agricultural lands. This allowed for the creation of surplus, visible through large

114 · Anna Osterholtz

storage areas at numerous coastal and inland sites (Keswani 1993; Steel 2004). The creation of agricultural surplus and new farming techniques made the expansion of sites into previously unavailable areas possible. Population size increases that occurred at the beginning of the PreBA have historically been explained as resulting from either colonization or integration. Frankel (2000) views this population increase as evidence of Anatolian colonization, in which individuals from Anatolia settle on the island and bring a dominant cultural model that swamped the local etioCypriot identity. Alternatively, Knapp uses essentially the same dataset to argue for integration of external populations into an already existing etio-Cypriot world through the creation of a new third-space identity (Knapp 2012: 33). This model of hybridization is in line with postcolonial theory (Voskos and Knapp 2008), one that seeks to understand the “amalgamation” of “materialities” (Knapp 2012: 32), or the blending of material cultures into a new understanding of what it means to be Cypriot in the PreBA. PreBA mortuary customs revolved around the use of rock-cut chamber tombs. The Kalavasos village tombs are an excellent example of the early PreBA mortuary complex. Chamber tombs were constructed away from associated settlements and consisted of three basic elements (the dromos, the stomion, and the burial chamber itself) (Pearlman 1985; Todd 1986; Todd and Flourentzos 2007; Todd and Pearlman 2007). Chamber tombs may be simple or complex but tended to have a significant number of single interments. Early PreBA tombs at Lapithos Vrysi tou Barba show distinct signs of social differentiation and hierarchy, with larger burial chambers containing more metal artifacts than those without grave goods (Keswani 2004). Burials tend to be laid in a flexed position on one side. Most have single individuals interred with grave goods consisting of ceramics, spindle-whorls, jewelry, and metal implements (Swiny and Herscher 2003). Over time, there is an increase in the number of small, plain bowls with a decrease in the use of jugs within the burial contexts. Ceramics typically associated with feasting decrease in number with time (Knapp 2013). Skeletal analysis showed that the PreBA interments from the Kalavasos village tombs are those of primary burials (Osterholtz 2015). Knapp (2013) argues that, for some PreBA cemeteries, burials become a focal point for competitive display. The location of the tombs may also be important, as Keswani (2005) links the use of mortuary ritual to lineage-based

Collective Bodies, Collective Identities: Development in Bronze Age Cyprus · 115

land claims, alluding to the feedback loop between copper and cattle. As Knapp (2013: 321) notes, mortuary display and luxury imports may have attracted the attention of foreign visitors and traders, “broadening the exposure Cypriot culture and its rich copper resources to the wider eastern Mediterranean world.” The importance of external trade goods and contacts in the creation of Cypriot identity here implicitly calls for mortuary processes to be both public and performative. This performance, and its reliance on external goods and contacts, helped to create a Cypriot identity that emphasized relationships external to Cyprus, setting up the island as a good place for migration and the continuation of the migrant identity diffusion cycle discussed above. ProBA Population Structure and Mortuary Practices Population size continued to increase throughout the PreBA due to expanded agricultural lands. Expanding industrialization of production of trade goods is evident at such sites as Kalavasos Ayios Dhimitrios, which has a structure filled with large pithoi for storing olive oil. Building X at Kalavasos Ayios Dhimitrios had enough pithoi to store at least 35,000 liters of olive oil (South 1988, 1996; South and Russel 1989; South et al. 1989). Production of ceramics on the island also became more standardized in the ProBA with the introduction of the potter’s wheel. There is significant debate about the nature of pottery production, particularly whether the standardization represents a unified centralized production or not (Crewe 2007; Sherratt 1991). However, the introduction of the same technology indicates that, at the very least, communities were in communication with each other and sharing technology. Significant social hierarchy is inferred by Keswani (2004), who argues that coastal centers were in control of inland copper resources and controlled the exchange of these goods. In this way, access to foreign goods and contacts was important for high status and social differentiation. Hirschfeld (1992, 2000) notes that Cyprus likely played the role of middle-man in exchange, providing a common location for dissemination of goods. Hirschfeld (1992, 2000) found that pottery marks on Mycenaean ceramics indicate that Mycenaean ceramics were likely disseminated throughout the eastern Mediterranean by Cypriots, not Mycenaeans. Elites controlled access to highly prized foreign trade goods.

116 · Anna Osterholtz

From a biological perspective, this intensification of social hierarchy may have led to increased levels of structural violence. Structural violence can take the form of differential access to high-quality nutrition, particularly in childhood. Individuals lacking that access may have decreased overall health and increased rates of opportunistic infection (Goodman et al. 1988). Sites had become specialized in the ProBA, meaning that some may have better access to high-quality nutrition than others. Keswani (1993, 2004) argues that urbanization is the outcome of this site specialization and increased social complexity. Also important are systems of physical violence, or violence that leaves direct evidence of trauma on the skeleton. While physical violence is present from the earliest assemblages on Cyprus, the intensification of social hierarchy may lead to different exposure to physical violence, with some groups (e.g., sex or age cohorts) exposed to different patterns of physical trauma. During the ProBA, population size increases are interpreted as reflecting a hybridization model and an internal development model. In the hybridization model, Mycenaean colonists, who were familiar with Cyprus due to generations of trade, settled in Cyprus after the fall of the palatial system in the western Mediterranean (Karageorghis 1982; Leriou 2002). This model is difficult to view archaeologically, and is typically inferred based on the presence of nonlocal ceramics that have been recovered from high-status contexts (Catling and Karageorghis 1960; Karageorghis 1982, 1976; Knapp 1997; Manning and Hulin 2005) and the influence of Mycenaean tomb architecture on Cyprus. Iacovou (1998, 2008) points to the introduction of the Greek language as the best sign of significant Mycenaean contact and cultural exchange. In essence, the hybridization model argues for the integration of Mycenaean colonists into the larger Cypriot social structure, with higher-status individuals taking on elements of Mycenaean identity. Iacovou (1999, 2008) and Karageorghis (1982) do not distinguish between colonization and migration in terms of the construction or destruction of social ties implicit in these different types of population movements. Therefore, even though these researchers define this as colonization, it falls into the category of migration presented in this chapter. The influx of populations and their integration of significant elements of Mycenaean social identity into an existing Cypriot hierarchy allows for the renegotiation of social roles and statuses, with access to the newcomers and their ties to the western Mediterranean highly prized. The

Collective Bodies, Collective Identities: Development in Bronze Age Cyprus · 117

third-space created by this population influx honored both internal ties (particularly those in control of natural resources and agricultural goods coming from the inland regions of the island) and external ties (including those with access to trade goods or trade relationships). In contrast, the internal development model examines the changes from a primarily Cypriot viewpoint. Iacovou (2008) emphasizes the importance of the Greek language, and deemphasizes the role of influences from Anatolia and the Levant. The internal development model looks westward and so doesn’t completely account for the importance of the overall Mediterranean regional sphere. It isn’t just the influence of the western Mediterranean, but the overall importance of Cyprus and its multicultural outward facing identity that is important. The internal development model tends to treat Greek and Cypriot identity as equivalent. Mortuary practices of the ProBA are diverse and extensive. External contacts are evident through the incorporation of tomb types identified with both Aegean and Anatolian cultural affinities (Keswani 2004). Though significant variability occurs, there are also some broad patterns. Burials move from extramural locations to intramural locations and may include either single or multiple interments. Grave goods also tend to incorporate more prestige goods, including artifacts made from gold, ivory, glass, faience, and ostrich eggs, as well as imported goods. Finally, the political climate also changed during the beginning of the ProBA. With the breakdown of the palatial system, the economic relationships between Cyprus and the west were in flux, which may have created economic opportunities for traders on Cyprus. The relative prosperity of Cyprus may be seen as evidence of political autonomy from their trading partners despite the cultural similarities. Cyprus remained a stable political entity in an unstable sea. Keswani (2004) argues that ProBA burial practices reflect new social realities and hierarchies, including the use of burial space as an important marker of administrative control over land or resources. Both Knapp (2013) and Keswani (2004) argue that the use of monumental architecture, luxury goods, and intramural locations is important for expression of status. The manipulation of mortuary space, the movement of tomb space into and the association of tombs with administrative structures indicates that the bodies and the structures in which they are housed are used to negotiate social status and maintain control over resources.

118 · Anna Osterholtz

Tombs and Identity

The movement of tombs into city spaces can be seen as a mechanism for the creation of lineage-based control over economic resources, whether they be access to trade routes or control of olive oil surplus. In addition to using the tombs as economic anchors, the tying of lineages to certain monumental structures and the destruction of earlier burial plots may be seen as the superimposition of new elites or lineages over old elites or lineages. The destruction of burial locations in antiquity at the sites of Vournes and Tsarouklas can be interpreted as competition between groups for political legitimacy (Manning 1998; Webb 1999). Keswani (2004) argues that these destructions were tied to newcomers from mixed descent groups (embodying the third-space) who were not closely tied culturally to the human remains. Were these remains disturbed and the tomb spaces destroyed as a mechanism of destroying another’s claims to the area, allowing for a redefinition of history? In areas with continuous habitation, such as the Vasilikos Valley, residents built tombs over the existing ones, but earlier tombs were not destroyed, suggesting continuity between newly constructed tombs and the older tombs or a desire by new lineages to associate with previous lineages. The current chapter seeks to add an additional line of evidence to this discussion, namely the analysis of the inhabitants of the tombs and communities. Detailed below are the methods employed (including a discussion of the use of techniques specifically designed for commingled and fragmentary assemblages), and the expectations for both the colonization and migration/integration/third-space phenomenon models. Skeletal Analysis

Bioarchaeological analysis of archaeological human remains from six sites throughout the southern coast of Cyprus was conducted to test the nature of population movement on Cyprus (table 5.1). Colonization is expected to result in poorer overall health scores and changing patterns of violence after population influx. In contrast, migration and the creation of a third-space is expected to present a more static appearance of health and violence, while still allowing for social change to occur. The analysis is fundamentally biocultural in nature, seeking to put bioarchaeological

Collective Bodies, Collective Identities: Development in Bronze Age Cyprus · 119

Table 5.1. Sites examined and their associated time periods and calibrated radiocarbon dates Refined Time Time Period Period

Calibrated Radiocarbon Dates

PreBA

PreBA 1

2400–2000 BC

PreBA 2

2000–1700 BC

ProBA 1

1700–1450 BC

ProBA 2

1450–1300 BC

ProBA 3

1300–1100 BC

ProBA

Site Name Sotira Kaminoudhia Kalavasos Village Tombs Sotira Kaminoudhia Kalavasos Village Tombs Episkopi Phaneromeni Kalavasos Village Tombs Kalavasos Ayios Dhimitrios Kalavasos Mangia Kalavasos Village Tombs Hala Sultan Tekke

data back into a larger archaeological discussion of the nature of social change and the development of Cypriot identity. The biocultural model employs the use of a population perspective that provides a view into differential access by groups and the health of society as a whole. The model also recognizes culture as an adaptive force within human environments and is therefore linked to biological adaptation, as culture can either inhibit or encourage disease (Armelagos 2011; Zuckerman and Armelagos 2011). In addition, a biocultural approach is useful for organizing data that can then be examined through various theoretical lenses, including those of landscape, identity, and the concept of collective kin. The analysis presented here employs specialized techniques associated with the analysis of commingled and fragmentary remains. Typically, bioarchaeological research begins with the analysis of individuals whose data are then aggregated to understand population-level changes. With commingled and fragmentary remains, however, the individual is no longer visible, but skeletal elements are. Elements, therefore, form the basis of analysis upon which the population-level inferences are drawn. Because of the lack of discrete individuals, research questions must be adapted to address this research challenge. It is, for example, impossible to trace the extent of pathological processes across multiple elements when commingling occurs, and so specific differential diagnoses are difficult

120 · Anna Osterholtz

to perform. General indicators of adult and childhood nutrition can be examined through frequencies of cribra orbitalia and porotic hyperostosis (Stuart-Macadam 1985), the number and timing of linear enamel hypoplasias (Goodman and Rose 1990), and the presence and location of periosteal reactions, particularly on the anterior surface of the tibiae (see complete discussion in Osterholtz 2015). Trauma analysis was also completed for both fragmentary and intact crania. Analysis of entheses was completed as well, in order to understand changing activity patterns through time, although specific activities were not inferred (Mariotti, Facchini, and Belcastro 2007; Milella, Mariotti, and Belcastro 2011; Milella et al. 2012).1 Data were collected on more than one thousand bone fragments and teeth representing at least one hundred individuals. These archaeological human remains are currently curated at various district and national museums throughout Cyprus and were accessed during 2013 and 2014. They were analyzed by site, each of which has a single-period occupancy. This chapter will focus on temporal comparisons as opposed to intraisland regional comparisons. The only multiperiod site included in the analysis was the Kalavasos Village Tombs. The tombs associated with the later ProBA period all contained multiple individuals with burials of subadults and adults. Earlier burials from this site contained both individual and multiple interments. The later burials contained more multiple interments than the earlier burials, which is a pattern mirrored in the later ProBA period sites of Kalavasso Ayios Dhimitrios, Kalavasos Mangia, and Hala Sultan Tekke. The inclusion of multiple individuals in these later tombs argues for lineage-based places of burial, where inclusion within the community is the only requirement for inclusion in the tomb. These tombs were located under the streets adjacent to the administrative structures (such as Building X at Kalavasso Ayios Dhimitrios); the association of the tombs with the administrative structure suggests the importance of the tomb for administrative control, especially when the grave goods of these tombs are some of the wealthiest on the island. Most health indicators show no differences over time or by region, suggesting that the subsistence base is relatively stable and that populations weren’t subjected to the structural violence that would have been expected during colonization. Some health indicators do show general trends, as discussed below, but very few were statistically significant.

Collective Bodies, Collective Identities: Development in Bronze Age Cyprus · 121

Table 5.2. Linear enamel hypoplasia frequency by site and time period Site

Time Period

Sotira Kaminoudhia

PreBA 1 (2400–2000 Cal BC) ProBA 1 (1700–1450 Cal BC) ProBA 1 and ProBA 2 (1700–1300 Cal BC) ProBA 2 (1450–1300 Cal BC) ProBA 2 (1450–1300 Cal BC) ProBA 3 (1300–1100 Cal BC)

Episkopi Phaneromeni Kalavasos Village Tombs Kalavasos Ayios Dhimitrios Kalavasos Mangia Hala Sultan Tekke

MNI (from all contexts) 12

LEH frequency* 3%**

21

29%

63

7%

45

55%

17

20%

20

52%

Notes: * n teeth with LEH/total n teeth scored. ** Because of poor dental preservation, this frequency likely underestimates the true LEH frequency.

Linear enamel hypoplasias (LEH) increase in frequency through time. Two assemblages stand out with high percentages of teeth with LEH: Kalavasos Ayios Dhimitrios (ProBA 2) and Hala Sultan Tekke (ProBA 3) (table 5.2). Both sites are large administrative centers that appear to have held administrative roles over the accumulation of goods and had a role in the distribution of trade goods. This increase in LEH frequency could also be the result of social change associated with the intensification of agricultural production. In a large-scale analysis of precontact North America, Martin and Osterholtz (2016) found childhood stress increases with intensification of agricultural production; importantly, this period was a time of intensification and specialization in Cyprus. Finally, Hala Sultan Tekke is located on the coast, specializing in trade, and possibly having a different exposure to disease vectors than the inland sites. Caries rates and antemortem tooth loss show no discernable patterns, likely indicating no major changes in diet through time even though grain processing methods appear to have undergone changes. Dental wear patterns differ through time, with more severe wear present in earlier periods. For the analysis of dental wear, each tooth was scored independently due to the high number of isolated teeth at all sites. Sexes were also pooled for this reason. Small sample sizes may skew these data (Episkopi

122 · Anna Osterholtz

Figure 5.3. Boxplot of changes in dental wear through time.

Phaneromeni, n = 1; Kalavasos Ayios Dhimitrios, n = 9; Kalavasos Mangia, n = 4; Kalavasos village tombs, n = 8; and Sotira Kaminoudhia, n = 2), but the only tooth that yielded a significant result for dental wear was the maxillary right central incisor (p = .026) (figure 5.3). The reduced levels of tooth wear in adults in later periods is likely related to differences in how grain was processed. Different patterns of extra-masticatory tooth use are also present (see figure 5.4 for an example), although the sample sizes are too small for statistical testing. Regardless of time period, with a single exception, only individuals who died as nonadults exhibited unhealed or healing porotic hyperostosis and/or cribra orbitalia. Adults, regardless of time period or site, had only healed lesions. Frequency data are not given here due to the small and skewed samples. Preservation also inhibited a thorough analysis, and so trends can be noted but not tested with these samples (table 5.3). This suggests that the stresses leading to cribra orbitalia were common in childhood, and that adults exhibiting cribra orbitalia had survived those

Collective Bodies, Collective Identities: Development in Bronze Age Cyprus · 123

Figure 5.4. Examples of extramasticatory tooth use.

stresses or infections. No temporal patterning is visible in the postcranial periosteal deposition recorded. Both adults and nonadults had active periosteal deposition as well as healed lesions. Two cases of possible perinatal/infantile scurvy were identified based on pathognomonic lesions of the greater wing of the sphenoid. The term possible is used intentionally here as a thorough differential diagnosis cannot be conducted due to the commingled nature of the assemblage. Common elements to a differential diagnosis for scurvy would include normal appositional bone growth, systemic infectious processes (such as treponematosis and viral infection), and rickets (Alesandri et al. 1995; Cantey, Pritchard, and Sanchez 2013; Jaffe 1972; Ortner 2003; Ortner et al. 2001). Increased abnormal porosity and bony deposition on the external surface of the greater wing of the sphenoid are a diagnostic criterion according to Snoddy and colleagues (2018), so the possible identification of perinatal/ infantile scurvy can be made for the individuals at Kalavasos Mangia and Kalavasos Ayios Dhimitrios. Both date to the ProBA 2 period (one from Kalavasos Ayios Dhimitrios and the other from Kalavasos Mangia). These were the youngest individuals examined in this analysis and the only perinatal individuals in any of the samples. This is suggestive of some degree of maternal insufficiency during pregnancy or very early natal life, but not

Adult (18+)

Sotira Kaminoudhia

Kalavasos Village Tombs

Y. Adult M. Adult (35–50 y)

Adult

Child Adolescent + (12+ y)

Y. Adult (18–35 y) Episkopi Phaneromeni Child (2–12 y) Adult

AAD

Site Female Male Indeterminate Female N/A Female Male Indeterminate N/A Female Indeterminate Female Ambiguous Male Indeterminate Male Male

Sex

Table 5.3. Cribra orbitalia and porotic hyperostosis frequencies CO

2 1 1 6 1 6 5

1

3 1 1 1 1

1 1

1 13 2 7

5 1 3 1

1E

CO NS No CO Unhealed

Left

1 Mi

2 Mi

1 Mi

1 Mo

Healed

CO

1 7 2 3 1 1

1

1

6 1 9

1 1E

CO NS No CO Unhealed

Right

1 Mi

Healed

31 1 1 1

Indeterminate Female Indeterminate Male

2 2

1 1 1

1

1

2 1

2 2 5 8

1 12 1 Mi

1 Mo 1E

1 Mi

1 Mi 2 Mo

1 Mo

31 1 1 1

2 2 2 5 1 17 1 16

1

2

1

1 1

1

3

2E 1 Mi 1 Mo

1E

1E

Mi

1 Mo

Notes: AAD = Age at Death; CO NS = cribra orbitalia not scorable; No CO = cribra orbitalia not present; Mi = minimal expression; Mo = moderate expression; E = extensive expression.

M. Adult

Y. Adult

Subadult Infant Child Adolescent Adult

Hala Sultan Tekke 5 1 17 1 15

Nonadult Preterm (